You are on page 1of 14

Handout-7

Course 201N
1st Semester 2006-2007
Inorganic Chemistry
Instructor: Jitendra K. Bera

Contents
3. Organometallic Chemistry

Introduction
Chelate effect, hapticity and typical ‘organometallic’ ligands
Carbonyl
Organometallic Chemistry
What are organometallic compounds?
An organometallic compound is broadly defined as one that involves one or more carbon
atoms of an organic group or molecule and a transition, lanthanide, actinide, or main
group metal atom. In 1848, E. C. Frankland synthesized dimethylzinc, and then
subsequently discovered Zn(C2H5)2, Hg(CH3)2, Sn(C2H5)4 and B(CH3)3. Among the
oldest and well-used organometallic compounds include the Grignard reagents RMgX,
the alkyl lithiums, RLi. They have found extensive use in the laboratory-scale synthetic
chemistry. Since the 1960s, exploratory synthetic research in organometallic compounds
has been dominated by studies of d-block elements, but a revival of interest in main-
group organometallic compounds has produced new classes of compounds. In this course
we will be focusing primarily on transition metal based organometallic compounds.

1760 Paris, Cadet works on inks based on Co salts. Prepared from cobalt minerals containing Arsenic.
As2O3 + 4 CH3COOK → [(CH3)2)As]2O. Considered as first organometallic compound.
1827 Zeise’s salt Na[PtCl3C2H4] first olefin complex
1863 Friedel and Crafts prepare organochlorosilanes
1900-20 Grignard RMgX synthetic application
1909 Pope Me3PtI
1912 Nobel prize to Grignard and Sabatier
1917 Schlenk prepares Li alkyls via transalkylation from R2Hg
1930 Ziegler and Gilman simplify organolithium preparation, using ether cleavage and alkyl halide
metallation respectively
1920-50 Ziegler, Gilman, Wittig RNa, RK, RLi
1931 Hieber H2Fe(CO)4
1953 Wittig discovers the reaction bearing his name
1953 Ziegler TiCl4/R3Al polymerization reaction
1956 Brown develops hydroboration
1959 Smidt, Hafner [(C3H5)PdCl2]2 allyl complex
1963 Nobel prize to Ziegler and Natta
1964 Fischer (CO)5W=C(OMe)Me carbene
1965 Wilkinson and Coffey (PPh3)3RhCl homogeneous catalyst
1968 (C8H8)2U First organouranium compound
1973 Wilkinson, Fisher Nobel prize
1979 Nobel prize to Brown and Wittig, use of organoboranes and methylenephosphoranes respectively
in organic synthesis
2005 Grubbs and Schrock Noble Prize

The information box above lists some of the landmark achievements in the area of
organometallic chemistry.

Chelate Effect “Chelate” is from the Greek meaning “claw” or to grab on to. The
chelate effect or chelation is one of the most important ligand effects in transition metal
coordination chemistry. Since most metal-ligand bonds are relatively weak compared to
C-C bonds, M-L bonds can often be broken rather easily, leading to dissociation of the
ligand from the metal. Consider the two metal ligand complexes shown below:
L L

M L M + L

L L
L
M L M
The second metal complex is much less likely to lose one of the ligands due to the
bridging group that holds the ligands in proximity to the metal center. From a kinetic
viewpoint, if one of the ligands dissociates, it will remain close enough to the metal
center to have a high probability of re-coordinating before another ligand can bind. From
a thermodynamic viewpoint, by tethering two donor ligands together, one removes the
entropic driving force for dissociating a ligand and thus making more particles in solution
(more disorder). The chelate effect can be extremely dramatic. There are cases known
where the presence of a chelate will change the equilibrium constant by a factor of 1010
favoring the coordinated form of the ligand to the metal over the dissociated form.
Naturally, the longer and more flexible the bridging group in a chelating ligand, the less
dramatic the chelating effect.

Hapticity “Eta-x” was originally developed to indicate how many carbons of a π-


system were coordinated to a metal center. It is now used to describe the bonding mode
of a ligand to a metal center. An η5-cyclopentadienyl ligand, for example, has all five
carbons of the ring bonding to the transition metal center.
ηx values for carbon ligands where the x value is odd usually indicate anionic carbon
ligands (e.g., η5-Cp, η1-CH3, η1-allyl or η3-allyl, η1-CH=CH2). The # of electrons
donated by the ligand is usually equal to x + 1. When x value is even, ηx usually indicate
neutral carbon π-system ligands (e.g., η6-C6H6, η2-CH2=CH2, η4-butadiene, η4-
cyclooctadiene). The # of electrons donated by the ligand in the even (neutral) case is
usually just equal to x.

M
M M M
η5-Cp η3-Cp η3-allyl η1-allyl

The nomenclature, however, has been generalized by most in the organometallic


field to include non-carbon ligands when there is some question as to the bonding mode
(or hapticity) of the ligand donor atoms to the metal. For example, the bisphosphine
Ph2PCH2CH2PPh2 (dppe) is normally a chelating ligand, but there are metal complexes
known where only one of the phosphine atoms is
coordinated to the metal center and the other is
“dangling.” The nomenclature for such a singly
coordinated bisphosphine ligand would be: M(η1- M PPh2 PPh2
dppe) -- meaning that only one of the two possible
phosphorus atoms is bonded to the metal. Note that in η1-dppe
cases like this, having an odd hapticity does NOT mean
that the ligand is anionic (it is, however, a 2 electron donor!). When dppe is coordinated
in its normal chelating mode, one usually omits the η2-designation, as this is considered
redundant.

Typical Ligands Used


Unlike the typical Lewis bases that were found in the classical complexes, most of the
ligands encountered in Organometallic chemistry are of the type known as π-acid ligands.
All of these have some type of higher energy orbitals, which can receive back electron
density from the metals. Consequently we find that these ligands stabilize metals in low
oxidation states. In contrast to classical complexes, organometallic compounds contain
metal usually in low oxidation states (in some situations the metal is found in a negative
oxidation state also!). The table given below summarizes a few more ligands that are
mostly employed in Organometallic chemistry.

Ligand Name Bonding Type Formal Electrons


Charge Donated

Molecular Hydrogen: H2 H 0 2
M
H

Hydride: H− M-H -1 2
Amine, phosphine, arsine: NR3, M-NR3 0 2
PR3, AsR3
M-PR3

Nitrosyl: N≡≡ O+ M N O +1 2
linear form

≡O−
Nitrosyl: N≡ M N -1 2
bent form
O
Carbonyl: C≡
≡O M C O 0 2
Carbonyl: C≡
≡O O 0 2 to each
η2 mode
µ,η C metal
M M
Carbonyl: C≡
≡O O 0 1 to each
µ bridging
metal
C
M M
Carbonyl: C≡≡O O 0 1 to each
µ3 bridging
metal
C
M M
M
Alkene: R2C=CR2 R R 0 2
η2 bonding mode R C C R

M
Alkyne: RC≡
≡CR R C C R 0 2 or 4
2
η bonding mode

M
Dienes: R2C=CH-CH=CR2 0 4
η4 bonding mode
M
Benzene: C6H6 0 6
η6 bonding mode
M

Aryl: C6H5− -1 2
η1 terminal M

Alkenyl (vinyl), alkynyl: M R -1 2


CH=CH2−, C≡ ≡CH−
η1 terminal M
Alkenyl (vinyl): -1 4
CH=CH2− M
η2 terminal

Allyl: C3H5− -1 4
η3 terminal
M

Allyl: C3H5− M -1 2
η1 terminal

Cyclopentadienyl: C5H5− -1 6
η5

M
M

Cyclopentadienyl: C5H5− -1 4
η3

M
Carbene: =CYR R 0 2
where Y is a substituent capable of π
interaction with the carbene carbon M C
atom Y

Alkylidene: =CR22−
− R -2 4
where no substituents capable of π M C
bonding to the carbene carbon atom
are present R
No need to memorize the above table. Here is a summary that will help you.


Cationic 2e- donor: NO (nitrosyl)
Neutral 2e- donors: PR3 (phosphines), CO (carbonyl), R2C=CR2 (alkenes),
RC≡CR (alkynes, can also donate 4 e-), N≡CR (nitriles)
Anionic 2e- donors: Cl- (chloride), Br- (bromide), I- (iodide), CH3- (methyl),
CR3- (alkyl), Ph- (phenyl), H- (hydride)

Anionic 4e- donors: C3H5- (allyl), O2- (oxide), S2- (sulfide), NR2- (imide),
2-
CR2 (alkylidene)

Anionic 6e- donors: Cp- (cyclopentadienyl)

18-Electron Rule and the stability of organometallic complexes


It is generally observed that the stability of many organometallic complexes, particularly
those involving the first row transition metals can be predicted by using the so-called 18-
electron rule. This rule can be stated as, when a metal achieves an outer electron
configuration of ns2(n-1)d10np6 , there will be 18 electrons in the valence orbitals and a
closed stable configuration is the result. It should be pointed here that for many second
and third row transition metals 16 electron complexes are also found. The 18-electron
rule is not sacrosanct. Rather, it may be used quite advantageously to predict the stability
and existence of organometallic compounds. Let us see how to apply this to a few
examples.
Electron counting is the process of determining the number of valence electrons about a
metal center in a given transition metal complex.
To determine the electron count for a metal complex:
 Determine the oxidation state of the transition metal center(s) and the metal
centers resulting d-electron count. To do this one must:
o note any overall charge on the metal complex
o know the charges of the ligands bound to the metal center
o know the number of electrons being donated to the metal center from each
ligand
 Add up the electron counts for the metal center and ligands

[Many students made mistakes in d-electron counting. In many instances, the


concept was correct but student didn’t make it right as the electron counting was
wrong! If you are confused with metal d-electron counts, practice well and check
with the table provided below]
Examples

CH3
R3P CO
Re
PR3
CO
1) There is no overall charge on the complex
-
2) There is one anionic ligand (CH3 , methyl group)
3) Since there is no overall charge on the complex (it is neutral), and since we have one
anionic ligand present, the Re metal atom must have a +1 charge to compensate for the
one negatively charged ligand. The +1 charge on the metal is also its oxidation state. So
the Re is the in the +1 oxidation state. We denote this in two different ways: Re(+1),
Re(I), or ReI.
Now we can do our electron counting:

Re(+1) d6
2 PR3 4e-
4e-
2 CO
CH3
2e-
CH2=CH2 2e-
Total: 18e-
Examples of 18-electrons systems

[V(CO)6]-; Cr(C6H6)2; [Mn(CO)5]-; [Fe(OH2)6]2+; [Co(CN)6]3-; Ni(CO)4; [Cu(NH3)4]+

Cr (CO)6 Cr 6e Fe(CO)5 Fe 8e
6CO 12e 5CO 10e
Total 18 e Total 18e
Ni(CO)4 Ni 8e Fe(CO)4PPh3 Fe 8e
4CO 10e 4CO 10e
Total 18e 1PPh3 2e
Total 18e
Fe(Cp)2 Fe(II) 6e Cr(C6H6)2 Cr 6e
2Cp 12e 2C6H6 12e
Total 18e
Cr(NO)4 Cr 6e Fe(CO)2(NO)2 Fe 8e
4NO 12e 2CO 4e
Total 18e 2NO 6e
Total 18e

There are multiple examples of transition metal compounds with less (group 3, 4 and 10)
or more than 18 electrons in the metal shell:

CH2Ph OAr Me
Me Me H3N Cl
Ti W W Pd Co Ni
CH2Ph Me Me Cl NH3
PhH2C ArO OAr
CH2Ph Me

8e 9e 12 e 16 e 19 e 20 e

HW: These are 18 electron complexes. Do the counting yourself.

Ru
CO
W
CO

Metal Carbonyl Compounds


Transition metal carbonyl complexes (metal carbonyls) are among the most well studies
organometallic compounds. Almost all transition metals form carbonyls. Many of them
comply with the 18 electron rule. For examples:
V(CO)6, Cr(CO)6, Mn2(CO)10, Fe(CO)5, Co2(CO)8, Ni(CO)4
 Because of the π-acceptor ability of CO, a number of stable anionic metal
carbonyls exists: Ti(CO)62-, V(CO)6-, Cr(CO)42-, Mn(CO)5-, Fe(CO)42-, Co(CO)4-.

 Some of them can be protonated to form metal carbonyl hydrides (HCo(CO)4,


H2Fe(CO)4, HMn(CO)5 etc.) which often behave as strong protic acids in aqueous
solutions.

 Since CO is a weak σ-donor, cationic metal carbonyls are rare and very
electrophilic: Mn(CO)6+, Fe(CO)62+, Co(CO)5+, Ir(CO)63+, Hg(CO)22+

 There is a variety of polynuclear cluster carbonyl compounds containing M-M


bonds.
(CO)3
Co
CO
OC Os CO
CO CO Co(CO)3
CO
OC Os Os CO
OC CO
CO CO (OC)3Co Co(CO)3

O (CO)3
2-
Os

OC C CO (OC)3Os Os(CO)3

OC Co Co CO (OC)3Os

OC C CO Os
(CO)3

O
HW: Do electron counting for individual Co atom in Co2(CO)8. If you have paid enough
attention in this course, you will be able to explain the existence of a Co-Co bond in this
complex.

Molecular Orbital Diagram of CO


A simple MO diagram for CO is shown below. It would be sufficient if you concentrate
on the highest occupied molecular orbital (HOMO) and lowest un-occupied molecular
orbital (LUMO) at this level. The HOMO is the 5σ lone pair orbital mainly centered on
the carbon and weakly anti-bonding with respect to the C-O bond. The weak C-O
antibonding nature of this MO, however, is clearly seen in the experimental data
presented below. The LUMO is strongly π* antibonding and is low enough in energy to
act as a good acceptor orbital for interacting with filled d-orbitals on metals.
[The terms are symmetry designations, π molecular orbitals being anti-symmetric with respect to a defining
plane containing at least one atom, and σ molecular orbitals symmetric with respect to the same plane. In
the case of two-centre bonds, a π bond has a nodal plane that includes the internuclear bond axis, whereas a
σ bond has no such nodal plane ]

LUMO

HOMO
Experimental Data Supporting Nature of MO’s in CO
Species Config C-O Å νCO cm−1 Comment

CO (5σ)2 1.13 2143

CO+ (5σ)1 1.11 2184 5σ MO is weakly antibonding


CO* (5σ)1(2π)1 T 1.21 1715 2π MO is strongly antibonding

Carbonyl ligand is considered as a weak 2-electron σ-donor and very strong π-acceptor.
Two types of interactions are involved in the complexation of carbonyl with transition
metal ion. The carbonyl donates its lone pair to vacant metal d(σ) orbital and back-
donation occurs metal dπ orbital to C-O π* orbital.

Carbonyl IR Stretching Frequencies


 The position of the carbonyl bands in the IR depends mainly on the bonding mode of
the CO (terminal, bridging) and the amount of electron density on the metal being π-
backbonded to the CO.
 The number (and intensity) of the carbonyl bands one observes depends on the
number of CO ligands present and the symmetry of the metal complex.

Bonding Modes
As one goes from a terminal CO-bonding mode to µ2-bridging and finally µ3-bridging,
there is a relatively dramatic drop in the CO stretching frequency seen in the IR.
As an example the IR spectrum of [(η5-Cp)Fe (CO)2]2 is shown
to illustrate the clear difference in the region of absorptions of
the two types of carbonyls. Note that the electron count cannot
tell us about the existence (absence) of the bridging carbonyls.
This has to found from the experiment.

Effect of Electron Density on Metal:


As the electron density on a metal center increases, more π-
backbonding to the CO ligand(s) takes place. This further
weakens the C-O bond by pumping more electron density into
the formally empty carbonyl π* orbital. This increases the M-CO bond strength making it
more double-bond-like, i.e., the resonance structure M=C=O assumes more importance.
This can clearly be seen on the table that illustrates the effect of charge and
electronegativity on the amount of metal to CO π-
backbonding and the CO IR stretching frequency.
The variation in the νCO for the complexes given
right can be readily explained by looking at
charge on the central metal ion. As the positive
charge increases, the metal is less likely (less
effective) in the back bonding involving π-
overlap. If there is less of π-bonding, the C-O
bond has greater bond order; a higher value for
the stretching frequency is consequently observed.
Similarly, the higher negative charge on the metal
increases the extent of π-backbonding, and hence
the CO frequencies decreases.

Shown below is another example of the dramatic effect on the νCO IR stretching
frequencies on reducing Fe2(µ-PPh2)2(CO)6 by 2 electrons to form the dianionic
complex [Fe2(µ-PPh2)2(CO)6]2-. The average νCO frequency shifts almost 150 cm-1 to
lower energy on reduction.
Ligand Donation Effects
The ability of the ligands on a metal to donate electron density to the metal center
certainly has considerable effect on the absolute amount of electron density on that metal.
This, in turn, naturally effects the νCO IR stretching frequencies in metal carbonyl
complexes. Ligands that are trans to a carbonyl can have a particularly large effect on the
ability of the CO ligand to effectively π-backbond to the metal. For example, two trans
π-backbonding ligands will
partially compete for the same
d-orbital electron density,
weakening each others net M-L
π-backbonding. If both trans
ligands are carbonyls, they will
make equally strong M-C
bonds.
When the trans ligand is a σ-donating ligand, this can increase the M-CO bond strength
(more M=C=O character) by allowing unimpeded metal to CO π-backbonding. Pyridine
and amines are not that strong σ-donors, but they are even worse π-backbonding ligands.
So the CO has virtually no competition for π-backdonation. Look at the IR stretching
frequency of the CO group that is opposite to L in a series of complexes W(CO)5L. The
idea is as follows: if L
is a weak π-accepting
ligand, most of the
electron density of the
metal will be taken by
the CO trans to it, and
hence relatively low C-
O stretching frequency. This is what happens when pyridine(py) is opposite CO. When
PPh3 is the opposite ligand the competition for π-electron density increases and a slightly
higher value is observed. With PF3 is the opposite ligand, the value obtained for this
compound is nearly the same as for W(CO)6. It indicates that the π-acceptor properties
are nearly equal for CO and PF3.

CO CO CO
CO
CO CO CO
OC OC CO OC OC
W W W W
OC CO OC OC CO OC CO
CO
CO PF3 PPh3 Py
2010 cm-1 2007 cm-1 1942 cm-1 1895 cm-1

Based on CO IR stretching frequencies, the following ligands can be ranked from best π-
acceptor to worst:

NO+ > CO > PF3 > RN≡


≡C > PCl3 > P(OR)3 > PR3 > RC≡
≡N > NH3

You might also like