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

Full

This document is an introduction to the textbook "Introduction to Organometallic Chemistry" by M. S. Balakrishna and Prasenjit Ghosh. It provides an overview of the topics that will be covered in the textbook, including bonding concepts in main group chemistry, organometallic chemistry of s-block and p-block elements, transition metal organometallics, and applications of organometallic compounds. The introduction emphasizes that organometallic chemistry bridges organic chemistry and main group/transition metal chemistry by studying compounds containing metal-carbon bonds. It also outlines the organization of the textbook and lists some relevant literature.
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)
172 views

Full

This document is an introduction to the textbook "Introduction to Organometallic Chemistry" by M. S. Balakrishna and Prasenjit Ghosh. It provides an overview of the topics that will be covered in the textbook, including bonding concepts in main group chemistry, organometallic chemistry of s-block and p-block elements, transition metal organometallics, and applications of organometallic compounds. The introduction emphasizes that organometallic chemistry bridges organic chemistry and main group/transition metal chemistry by studying compounds containing metal-carbon bonds. It also outlines the organization of the textbook and lists some relevant literature.
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/ 151

INTRODUCTION TO

ORGANOMETALLIC
CHEMISTRY

M. S. Balakrishna & Prasenjit Ghosh


Indian Institute of Technology Bombay
Book: Introduction to Organometallic
Chemistry (Ghosh and Balakrishna)
This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.

The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by NICE CXOne and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact [email protected]. More information on our
activities can be found via Facebook (https://facebook.com/Libretexts), Twitter (https://twitter.com/libretexts), or our blog
(http://Blog.Libretexts.org).
This text was compiled on 01/09/2024
TABLE OF CONTENTS
Licensing

1: Introduction
1.1: Bonding Concepts in Main Group Chemistry
1.2: VSEPR Theory and its Utility

2: Organometallic Chemistry of s- and p-block Elements


2.1: General Methods of Preparation
2.2: Organometallic Compounds of Alkali Metals (Sodium and Lithium)
2.3: Organometallic Compounds of Alkaline Earth Metals (Beryllium and Magnesium)
2.4: Structure and Bonding

3: Organometallic Chemistry of p-block Elements


3.1: Reactions of Organometallic Compounds
3.2: Organometallic Compounds of Boron and Aluminium
3.3: Organometallic Compounds of Gallium and Indium
3.4: Zeigler Natta Polymerization Catalysts
3.5: Organosilicon and Organogermanium Compounds
3.6: Organotin and Organolead Compounds

4: Organoelement Compounds of Group 15


4.1: Organometallic Compounds of As(V) and Sb(V)
4.2: Organometallic Compounds of As(III) and Sb(III)
4.3: Phosphines

5: Group 12 Elements
5.1: Organometallic Compounds of Zinc and Cadmium
5.2: Organometallic Compounds of Mercury

6: General Properties of Transition Metal Organometallic Complexes


6.1: 18 Valence Electron Rule
6.2: Synthesis and Stability

7: Metal Alkyls and Metal Hydrides


7.1: Transition Metal Alkyl Complexes
7.2: Metal Hydrides

8: Carbonyls and Phosphine Complexes


8.1: Metal Carbonyls
8.2: Metal Phosphines

1 https://chem.libretexts.org/@go/page/191330
9: Complexes of π−bound Ligands
9.1: Metal Alkene Complexes
9.2: Metal Allyl and Diene Complexes
9.3: Metal Cyclopentadienyl Complexes

10: Reaction Mechanisms


10.1: Oxidative Addition and Reductive Elimination
10.2: Insertion and Elimination Reactions
10.3: Nucleophilic and Electrophilic Addition and Abstraction

11: Applications
11.1: Homogeneous Catalysis - I
11.2: Homogeneous Catalysis - II

12: Physical Methods in Organometallic Chemistry


12.1: Characterization of Organometallic Complexes

13: Multiply-Bonded Ligands


13.1: Metal-Carbenes
13.2: Metal-Carbynes

14: Metathesis
14.1: Catalytic Applications of Organometallic Compounds- Alkene Metathesis
14.2: Credits

Index
Index

Glossary
Detailed Licensing

2 https://chem.libretexts.org/@go/page/191330
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.

1 https://chem.libretexts.org/@go/page/417155
CHAPTER OVERVIEW

1: Introduction
1.1: Bonding Concepts in Main Group Chemistry
1.2: VSEPR Theory and its Utility

This page titled 1: Introduction is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S. Balakrishna &
Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1
1.1: Bonding Concepts in Main Group Chemistry
Learning Objectives
In this lecture you will learn the following:
Lewis structure.
Limitations of Lewis model.

Organometallic Chemistry of Main Group Elements


Our aim in this course is to introduce one of the important disciplines of inorganic chemistry which bridges the organic compounds
with main group and transition elements; organometallic chemistry. Organometallic chemistry is defined as the chemistry of
chemical compounds containing one or more metal—carbon bonds which are essentially polar (Mδ+—Cδ-) in nature. To begin with
it is necessary to understand the bonding concepts that explain the structures of both main group and transition elements. The
objectives of first few lectures are to give some insight into the various metal—ligand interactions that would later help in planning
the synthesis and also eventually looking into their electronic and kinetic stability. The course is divided into four major sections
and the first section deals with the various bonding concepts employed for both main group and transition metal compounds. The
second and third sections deal with the organometallic chemistry of main group and transition elements, respectively, while the
fourth section would be devoted to applications of organometallic compounds with special emphasis to catalysis.
A set of problems along with solutions are presented at the end and also in Chapter 2. In future, it is planned to provide interactive
sections as well.
The pertinent literature and books in this and related areas are listed at the end.

Molecular structure and bonding


The chemical properties of the molecules can be directly correlated to their electronic structures. In this lecture attempts are being
made to give an overall view of how bonding concept evolved starting from Lewis approach to the development of molecular
orbital theory.

1.1.1: Lewis structures


Lewis proposed that when two atoms come close to each other to establish a bond by sharing an electron pair, a covalent bond will
be established. One pair of electrons would give a single bond X—Y; two or three pairs of electrons would leads to the formation
of double (X=Y) and triple bonds (X≡Y), respectively. The pairs of valence electrons that are not utilized in bonding are called lone
pairs of electrons or simply lone pairs. The lone pair of electrons does not participate in bonding; however, they do influence the
shape and the geometry of the molecule and their chemical properties as well.
Lewis introduced octet rule which states that each atom shares its valence electrons with neighboring atoms to have a total of eight
(s2p6) electrons in its valence shell to have noble-gas configuration. An exception to this rule is hydrogen as it can have only two
valence electrons in its only shell, 1s.
By simply counting the number of valence electrons present on the central atom and its neighbors, Lewis structures can be written
in just three easy steps.
i. Consider the valence electrons of all participating atoms; add an electron for each negative charge and subtract one electron for
each positive charge.
ii. Identify the central atom and write the symbols of the atoms around central atom. In majority of polyatomic molecules, the least
electronegative one will be the central atom with an exception of hydrides, for example, H2O, NH3 or H2S.
iii. Distribute the electron pairs throughout the molecule to satisfy the octet of all atoms present in the molecule starting from the
most electronegative one. Each pair of singly bonded atoms requires one pair of electrons.
iv. Each bonding pair should be represented by a single bond and the net charge is assumed to be possessed by the ion (cation or
anion) as a whole and not by an individual atom.
For some molecules, Lewis dot structure differs from the experimentally determined structural observations. For example, in
acetate ion both the C—O bonds are identical as per X-ray structure determination but the prediction by the Lewis structure is
incorrect. The reason is due to resonance.

1.1.1 https://chem.libretexts.org/@go/page/172717
Limitations of Lewis model
Molecules with odd number of electrons can never satisfy the octet rule.

Example: NO.

Some atoms with fewer valence electrons can never complete octet without formal charges.

A central atom can have more than 8 electrons. Example: SF6

Lewis model does not explain paramagnetic nature of O2.

Geometry and molecular shapes cannot be explained by Lewis model.

Worked examples
Example 1 : ClO2-

Solution :
ClO2- Total number of electrons 7 + 2 x 6 + 1 = 20 = 10 pairs

Identify the central atom and connect the peripheral atoms with a pair of electrons as two dots and count the remaining electron
pairs
O : Cl : O 20-4 = 16 electrons left (8 pairs)

Complete the octet of oxygen atoms and count the remaining electron pairs

16-12 = 4 electrons

Complete the octet of chlorine atom,

Since all the electrons are utilized and octet is satisfied, no need of any multiple bonds.

The structure is or or
Example 2 : CO
Solution :
CO (carbon monoxide) Total number of electrons 4 + 6 = 10

Connect the two atoms with a pair of electrons as two dots and count the remaining electron pairs
C : O 10-2 = 8 electrons left

Complete the octet of oxygen atoms and count the remaining electron pairs

8-6 = 2 electrons

1.1.2 https://chem.libretexts.org/@go/page/172717
Place the remaining electron pair on carbon atom,

All the electrons are utilized and octet is not satisfied for carbon as there is a shortage of four more electrons.

Drag two electron pairs on oxygen atom in between carbon and oxygen to establish two more bonds so that a triple bond
exists between C and O which satisfies the octet of both C and O.

The structure is :C:::O: or :C≡O: or |C≡O|

Problems:
Work out Lewis structures for: BF4-, PCl3, PO43-, SO42-, O3, N2 and SO2

This page titled 1.1: Bonding Concepts in Main Group Chemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

1.1.3 https://chem.libretexts.org/@go/page/172717
1.2: VSEPR Theory and its Utility
Learning Objectives
In this lecture you will learn the following
Definition of VSEPR.
Hybridization concept.
Prediction of shapes and geometries of molecules.
Bent’s rule and its application in understanding the structural parameters.

The Valence Shell Electron Pair Repulsion Theory (VSEPR)


VSEPR theory is an improved and extension of Lewis model but predicts the shapes of polyatomic molecules. This model was first
suggested by Nevil Sidgwick and Herbet Powell in 1940 and later improved by Ronald Gillespie and Ronald Nyholm.
Prediction of molecular shapes and geometries was made easy by this model through the following simple steps.
1. Draw the Lewis structure.
2. Count the total number of bonds and lone pairs around the central atom. (Each single bond would involve one pair of electrons).
3. Arrange the bonding pairs and lone pairs in one of the standard geometries to minimize the electron-electron repulsion.
i. Lone pair electrons stay closer to the nucleus and also they spread out over a larger space than bond pairs and hence large
angles between lone pairs.
ii. The repulsion follows the order LP—LP > LP—BP > BP—BP.
4. Multiple bonds should be considered as a single bonding region.

Steric numbers:
Another term called steric number is often used in VSEPR theory.

Steric number (SN) = No. of attached atom + No. of lone pairs. Since the lone pair—lone pair repulsions are maximum, the most
stable geometry can be obtained by maximizing the distance between steric numbers on the central atom.

Molecular shapes are eventually determined by two parameters: Bond distance, separation between the nuclei of two bonded atoms
in a straight line and the bond angle, the angle between any two bonds containing a common atom.

While mentioning the molecular shapes lone pairs may be ignored, however, while defining the geometry both the lone pairs and
bond pairs should be considered.

For example: in water molecule the central oxygen atom is in tetrahedral environment with two lone pairs and two O—H bonds (or
two bond pairs). The shape of the water molecule is therefore bent (two lone pairs are ignored).
Similarly, in ammonia, the nitrogen atom is in tetrahedral environment with three bonded pairs (three N—H bonds) and one lone
pair. The shape of NH3 molecule is pyramidal.

Predicting the molecular geometries


To begin with, draw the Lewis structure.

Count the number of bonding pairs and lone pairs around the central atom.

Arrange the bonding pairs and the lone pairs in one of the standard geometries thereby minimizing electron—electron repulsion.

Multiple bonds count as a single bonding region.


What is Bent’s rule:
More electronegative substituents ‘prefer’ hybrid orbitals having less s-character, and more electropositive substituents ‘prefer’

1.2.1 https://chem.libretexts.org/@go/page/172718
orbitals having more s-character.

The bond angles in CH4, CF4 and CH2F2 can be explained using Bent’s rule. While a carbon in CH4 and CF4 uses four identical sp³
hybrids for bonding, in CH2F2 the hybrids used are not identical.

The C-F bonds are formed from sp3 + x hybrids, with slightly more p-character and less s-character than an sp³ hybrid, and the
hydrogen are bonded by sp3 - x hybrids, with slightly less p-character and slightly more s-character. Increasing the amount of p-
character in the C-F bonds decreases the F-C-F bond angle, because for bonding by pure p-orbitals the bond angle would be
decreased to 90°.

Molecular shapes determined by VSEPR theory


Steric Number
Molecule Geometry Example
(Number Electron Pairs) (SN)

MA2 2 Linear BeCl2

MA3 3 Trigonal planar BF3

MA4 4 Tetrahedral SiF4

MA5 5 Trigonal bipyramidal PF5

MA6 6 Octahedral SF6

MA7 7 Pentagonal bipyramidal IF7

Molecule SN Number of lone pairs Geometry shape Example

MA2 2 0 Linear CO2

3 0 Trigonal planar SO3


MA3 Trigonal planar
3 1 angular SO2

0 Tetrahedral CH4
MA4 4 1 Tetrahedral Trigonal pyramidal NH3
2 Angular H2O

0 Trigonal bipyramidal AsF5


1 Seesaw SF4
MA5 5 Trigonal bipyramidal
2 T-shaped ClF3
3 linear XeF2

0 Octahedral SF6
MA6 6 1 Octahedral Square pyramidal BrF5
2 Square planar XeF4

Problems:
1. Draw the structure and depict the geometry around Se atoms in [Se3O6F3]3-, which is a symmetric ionic molecule with cyclic
structure, using VSEPR model.
Solution

1.2.2 https://chem.libretexts.org/@go/page/172718
Apply VSEPR theory to the structure.
Se has six valence electrons. One Se—F and three Se—O (one terminal and two bridging) will add four more electrons to the
valence shell of Se, so, 10 e = 5 electron pairs out of which four are bonded pairs and one is a lone pair. A TBP environment is
expected.
2. Predict and draw the structure of I3+ using VSEPR model.
Solution

Total number of electron at the central I atom is: 7 + 2 -1 (charge) = 8; 2BP + 2LP, should be tetrahedral and angular shape.
3. Which of H2O and F2O will have the larger X-O-X bond angle?
Solution:

F is more electronegative than H; therefore the space occupied by the O-H bonding pair in the O valence shell will be greater.
Hence, H2O will have the larger X-O-X bond angle.
4. Explain why the X-P-X bond angles for the series of POX3 molecules decrease from X = Br (104.1°) to X = Cl (103.3°) to X = F
(101.3°)
Solution:

Fluorine is the most electronegative halogen, so it will draw electron density in the P-F bond away from P atom; repulsion of the
P-F bonding pairs will be less than the repulsion of P-Cl and P-Br bonding pairs, so the F-P-F bond angle will be the smallest.
5. Show that the following molecules and their corresponding shapes are correct, using VSEPR theory.

BCl3, trigonal planar

[IF5]2-, pentagonal planar

[NH4]+, tetrahedral

SF6; octahedral

XeF4; square planar

AsF5; trigonal bipyramidal

Xe(O)F4, square pyramid

IF7, pentagonal bipyramidal

[H3O]+ tetrahedral

H2Se tetrahedral

6. Work out the geometry of I3- ion using VSEPR model.


Solution

1.2.3 https://chem.libretexts.org/@go/page/172718
This page titled 1.2: VSEPR Theory and its Utility is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M.
S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

1.2.4 https://chem.libretexts.org/@go/page/172718
CHAPTER OVERVIEW

2: Organometallic Chemistry of s- and p-block Elements


2.1: General Methods of Preparation
2.2: Organometallic Compounds of Alkali Metals (Sodium and Lithium)
2.3: Organometallic Compounds of Alkaline Earth Metals (Beryllium and Magnesium)
2.4: Structure and Bonding

This page titled 2: Organometallic Chemistry of s- and p-block Elements is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

1
2.1: General Methods of Preparation
Learning Objectives
In this section you will learn the following
Various synthetic methodologies to make M—C bonds.
How to choose an appropriate synthetic method.
Reaction conditions and the role of solvents

An organometallic compound contains one or more metal-carbon bonds.

Synthesis
General Methods of Preparation
Most organometallic compounds can be synthesized by using one of four M-C bond forming reactions of a metal with an organic
halide, metal displacement, metathesis and hydrometallation.
a. Reaction with metal and transmetallation
The net reaction of an electropositive metal M and a halogen-substituted hydrocarbon is

2 M + RX(alkyl or aryl) ⟶ MR + MX

For example

8 Li + 4 CH Cl ⟶ Li (CH ) + 4 LiCl
3 4 3 4

Mg + CH Br ⟶ CH MgBr(organometal halide with Mg, Al, Zn)


3 3

If, one metal atom takes the place of another, it is called transmetallation
′ ′
M + M R ⟶ M + MR

2 Ga + 3 CH −Hg−CH ⟶ 3 Hg + 2 Ga (CH )
3 3 3 3

Transmetallation is favorable when the displacing metal is higher in the electrochemical series than the displaced metal.
b. Metathesis
The metathesis of an organometallic compound MR and a binary halide EX is a widely used synthetic route in organometallic
chemistry.

MR + EX ⟶ ER + MX

Li (CH ) + SiCl ⟶ 4 LiCl + Si (CH )


4 3 4 4 3 4

Al (CH ) + 2 BF ⟶ 2 AlF + 2 B(CH )


2 3 6 3 3 3 3

Metathesis reaction can frequently be predicted from electronegativity or hard and soft acid-base considerations.

Hydrocarbon groups tends to bond to the more electronegative element; the halogen favors the formation of ionic compounds with
the more electropositive metal.
In brief, the alkyl and aryl group tends to migrate from the less to the more electronegative element [χ = electronegativity].

2.1.1 https://chem.libretexts.org/@go/page/172724
When the electronegativities are similar, the correct outcome may be predicted, with care*, by considering the combination of the
softer element with organic group and harder element with fluoride or chloride.
*An insoluble product or reactant may change the outcome, e.g.;

SnPh (THF) + HgBr (THF) ⟶ HgPhBr(s) + PhSnBr(THF)


4 2

HgPhBr turns out to be insoluble in THF


Metathesis reactions involving the same central element are often referred to as redistribution reactions.

SiCl + SiMe ⟶ Me SiCl + Me SiCl +⋅ ⋅ ⋅


4 4 3 2 2

3 GeCl + 2 AlMe ⟶ 3 GeMe + 4 AlCl


4 6 4 3

Al is more electropositive than Ge, this reaction occurs as it is thermodynamically favorable.


c. Hydrometallation
The net outcome of the addition of a metal hydride to an alkene is an alkylmetal compound.

EH + H C=CH ⟶ E−CH −CH


2 2 2 3

The reaction is driven by the high strength of E-C bond relative to that of most E-H bonds, and occurs with a wide variety of
compounds that contain E-H bonds.
Hydroboration

Hydrosilylation

Ionic and electron-deficient compounds of Group 1, 2


Organometallic derivatives of all Group 1 metals are known. Amongst, the alkyllithium compounds are most thoroughly studied
and useful reagents.
Many of them are commercially available.
MeLi is generally handled in ether solution, but RLi compounds with longer chains are soluble in hydrocarbons.
Commercial preparation:

M + RX ⟶ MR(often contaminated with halide)

The best method would be:

HgR + 2 Li ⟶ 2 LiR + Hg
2

MeLi exists as a tetrahedral cluster in the solid state and in the solution. Many of its higher homologs exist in solution as hexamers
or equilibrium mixture of aggregates ranging up to haxamers.
The larger aggregates can be broken down by Lewis bases, such as, TMEDA.
Common organolithium compounds have one Li per organic group.
Several polylithiated organic molecules containing several lithium atoms per molecule are known.
The simplest example is Li2CH2, which can be prepared by the pyrolysis of MeLi which crystallizes in a distorted antifulorite*
structure. However, the finer details of the orientation of the CH2 groups are yet to be established.

2.1.2 https://chem.libretexts.org/@go/page/172724
*the antifluorite structure is the inverse of the fluorite structure in which the locations ofcations and anions are reversed. Look into
the structures of CaF2 (fluorite structure) and K2O (antifluorite structure). An fcc array of cations and all the tetrahedral holes are
filled with anions.

Radical anion salts


Sodium naphthalide is an example of an organometallic salt with a delocalized radical anion, C10H8-.
Such compounds are readily prepared by reacting an aromatic compound with an alkali metal in a polar aprotic solvent.
Naphthalene dissolved in THF reacts with Na metal to produce a dark green solution of sodium naphthalide.

Na(s) + C H (THF) ⟶ Na[ C H ](THF)


10 8 10 8

EPR spectra show that the odd electron is delocalized in an antibonding orbital of C10H8.
Formation of radical anion is more favorable when the π of LUMO of the arene is low in energy.
Simple MOT predicts that the energy of LUMO decreases steadily on going from benzene to more extensively conjugated
hydrocarbons.
Sodium naphthalide and similar compounds are highly reactive reducing agents.
They are preferred to sodium because unlike sodium, they are readily soluble in ethers.
The resulting homogeneous reaction is generally faster and easier to control than a heterogeneous reaction between one reagent in
solution and pieces of sodium metal, which are often coated with unreactive sodium oxide or with insoluble reaction products.
The additional advantage is that by proper choice of the aromatic group the reduction potential of the reagent can be chosen to
match the requirements of a particular synthetic task.
Alternative route to delocalized anion is the reductive cleavage of acidic C—H bonds by an alkali metal or alkylmetallic
compound.
Example:

Problems:
1. Classify the following reactions into, (i) hydrometallation, metal displacement, metathesis OR transmetallation reactions; (ii)
give an example for each case in the form of a balanced chemical equation.
Solution
a.

2.1.3 https://chem.libretexts.org/@go/page/172724
′ ′
M + Mx R ⟶ M + MR

….Transmetallation
e.g.:

2 Ga + 3 CH −Hg−CH ⟶ 3 Hg + 2 Ga (CH )
3 3 3 3

b.

MR + EX → ER + MX

….Metathesis
e.g.:

Li (CH ) + SiCl ⟶ 4 LiCl + Si (CH )


4 3 4 4 3 4

or

Al (CH ) + 2 BF ⟶ 2 AlF + 2 B(CH )


2 3 6 3 3 3 3

c. \[\ce{EH + H2C=CH2 -> E—CH2—CH3 } \nonumber \] ….Hydrometallation


e.g.:

Problems:
2. For each of the following compounds, indicate those that may serve as
(1) a good carbanion nucleophile reagent,
(2) a mild Lewis acid,
(3) a mild Lewis base at the central atom,
(4) a strong reducing agent. (A compound may have more than one of these properties)
(a) Li4(CH3)4, (b) Zn(CH3)2, (c) (CH3)MgBr, (d) B(CH3)3, (e) Al2(CH3)6, (f) Si(CH3)4, (g) As(CH3)3.
Solution
a. (MeLi)4 - good carbanion nucleophile and strong reducing agent
b. ZnMe2 - reasonable carbanion nucleophile, mild Lewis acid, reducing agent
c. MeMgBr - good carbanion nucleophile
d. BMe3 - mild Lewis acid
e. Al2Me6 - good carbanion nucleophile, strong reducing agent
f. SiMe4 - mild Lewis acid
g. AsMe3 - mild Lewis base

This page titled 2.1: General Methods of Preparation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

2.1.4 https://chem.libretexts.org/@go/page/172724
2.2: Organometallic Compounds of Alkali Metals (Sodium and Lithium)
Learning Objectives
In this lecture you will learn the following
How to synthesize and handle sodium, lithium compounds.
Structural features.

Organometallic compounds of alkaline metals


Organic compounds such as terminal alkynes which contain relatively acidic hydrogen atoms form salts with the alkali metals.
+ −
2 Et−C ≡ CH + 2 Na ⟶ 2 Na [Et−C ≡ C] +H
2

+ −
Me−C ≡ C−H + K[ NH ] ⟶ K [Me−C ≡ C] + NH
2 3

1
C H + Na ⟶ NaCp + H
5 6 2 2

NaCp is pyrophoric in air, but air-sensitivity can be lessened by complexing the Na+ with dme.
In the solid state, [Na(dme)][Cp] is polymeric
*Pyrophoric material: is one that burns spontaneously when exposed to air.

Transmetallation:

HgMe + Na ⟶ 2 NaMe + Hg
2

Organolithium compounds:
H ydrocarbon Solvent

nBuCl + 2 Li −−−−−−−−−−−−→ nBuLi + LiCl

Organolithium compounds are of particular importance among the group 1 organometallics.


Many of them are commercially available as solutions in hydrocarbon solvents.
Solvent choices for reactions involving organometallics of the alkali metals are critical. For example, nBuLi is decomposed by
Et2O to give nBuH, C2H4 and LiOEt.

2.2.1 https://chem.libretexts.org/@go/page/172725
Alkali metal organometallics are extremely reactive and must be handled in air- and moisture-free environments; NaMe, for
example, burns explosively in air.
Lithium alkyls are polymeric both in solution and in the solid state.

NMR is very useful in understanding the solution structures; 6Li (I = 1), 7Li (I = ½), 13C (I = ½)

The structures of (tBuLi)4 and (MeLi)4 are similar. nBuLi when mixed with TMEDA, gives a polymeric chain. TMEDA link
cubane units together through the formation of Li-N bonds.

Alkyllithium compounds are soluble in organic solvents whereas Na and K salts are insoluble, but are solubilized by the chelating
ligand TMEDA. Addition of TMEDA may break down the aggregates of lithium alkyls to give lower nuclearity complexes. E.g.
[nBuLi.TMEDA]2

However, detailed studies have revealed that the system is far from simple, and it is possible to isolate crystals of either
[nBULi.TMEDA]2 or [(nBuLi)4.TMEDA]∞.
In the case of (MeLi)4, the addition of TMEDA does not lead to cluster breakdown, and the X-ray structure confirms the
composition (MeLi)4.2TMEDA, the presence of both tetramers and the amine molecules in the crystal lattice.

N. D. R. Barnett et al. J. Am. Chem. Soc., 1993, 115, 1573.

Lithium alkyls and aryls are very useful reagents in organic synthesis and also in making corresponding carbon compounds of main
group elements.
Lithium alkyls are important catalysts in the synthetic rubber industry for the stereospecific polymerization of alkenes.

This page titled 2.2: Organometallic Compounds of Alkali Metals (Sodium and Lithium) is shared under a CC BY-NC-SA 4.0 license and was
authored, remixed, and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the
LibreTexts platform; a detailed edit history is available upon request.

2.2.2 https://chem.libretexts.org/@go/page/172725
2.3: Organometallic Compounds of Alkaline Earth Metals (Beryllium and
Magnesium)
Objectives
In this lecture you will learn the following
Organometallic compounds of beryllium and magnesium.
Structural features of alkyl lithium and beryllium sandwich compounds.

Beryllium

HgMe + Be ⟶ Me Be + Hg(at 383 K)


2 2

2 PhLi + BeCl ⟶ Ph Be + 2 LiCl(in diethyl ether)


2 2

In vapor phase, Me2Be is monomeric with a linear C—Be—C (Be-C = 170 pm).
The solid state structure is polymeric and resembles that of BeCl2.

2 NaCp + BeCl ⟶ Cp Be + 2 NaCl


2 2

The X-ray diffraction at 128 K suggested [(η1-Cp)(η5-Cp)Be].


However, 1H NMR spectrum shows that all protons environments are equivalent even at 163 K.
Also, solid state structure shows the Be atom is disordered over two equivalent sites and NMR data can be interpreted in terms of
fluxional process in which the Be atom moves between these two sites.
However, Cp*2Be possesses a sandwich structure with both the rings are coplanar.

Magnesium
Alkyl and aryl magnesium halides (Grignard reagents, RMgX) are extremely well-known on account of their uses in synthetic
chemistry.

Mg + Rx ⟶ RMgX(in diethyl ether)

Transmetallation is useful means of preparing pure Grignard reagents.

Mg + RHgBr ⟶ Hg + RMgBr

Mg + R Hg ⟶ Hg + R Mg
2 2

Two-coordination at Mg in R2Mg is observed only when the R groups are sufficiently bulky, e.g. Mg{C(SiMe3)3}2.
RMgX are generally solvated and Mg centre is typically tetrahedral.
e.g. EtMgBr.2Et2O; PhMgBr.2Et2O.
Cp2Mg has a staggered sandwich structure.

2.3.1 https://chem.libretexts.org/@go/page/172726
Solutions of Grignard reagent may contain several species, e.g. RMgX, R2Mg, MgX2, RMg(μ-X)2MgR, which are further
complicated by solvation. The position of equilibrium between these species is markedly dependent on concentration, temperature
and solvent; strongly donating solvents favour monomeric species in which they coordinate to the metal centre.

Treatment with dioxane results in the precipitation of MgCl2(dioxane) leaving behind pure R2Mg in the solution.
Problems:
1. The compound (Me3Si)2C(MgBr)2.nTHF is monomeric. Suggest a value of ‘n’ and propose a structure for this Grignard reagent.
Solution

2. If a typical Grignard reagent exists as an equilibrium mixture of dialkylmagnesium and magnesium halide, give a method of
isolating pure dialkyl magnesium. Your answer should be in the form of balanced chemical equations only.
Solution

2 RMgX −
↽⇀
− R Mg + MgX
2 2

Treatment of equilibrium mixture with dioxane results in the precipitation of, say, MgCl2(dioxane) (if, X = Cl), leaving behind pure
R2Mg in the solution.

2.3.2 https://chem.libretexts.org/@go/page/172726
This page titled 2.3: Organometallic Compounds of Alkaline Earth Metals (Beryllium and Magnesium) is shared under a CC BY-NC-SA 4.0
license and was authored, remixed, and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and
standards of the LibreTexts platform; a detailed edit history is available upon request.

2.3.3 https://chem.libretexts.org/@go/page/172726
2.4: Structure and Bonding
Learning Objectives
In this lecture you will learn the following
Solid state structures of nickel-arsenide, alkyl lithium and alkyl aluminium compounds

tructure and bonding


The slight differences that arise between organometallic compounds and binary hydrogen compounds are mainly due to the
tendency of alkyl groups to avoid ionic bonding.
The molecular structures of AlMe3and MeLi differ from AlH3and LiH.
Even the more ionic MeK crystallizes in the nickel-arsenide structure rather than the rock-salt structure adopted by KCl.
Nickel-arsenide structure is typical of soft-cation, soft-anion combinations.
Electron deficient compounds such as AlMe3contain 3c-2e bonds analogous to the B—H—B bridges in diborane.

The Nickel-Arsenide, NiAs, Structure

MeLi in nonpolar solvents consists of tetrahedron of Li atoms with each face bridged by a methyl group. Similar to Al2Me6, the
bonding in MeLi consists of a set of localized molecular orbitals. The symmetric combination of three Li 2s orbitals on each face of
the Li4 tetrahedron and one sp3 hybrid orbital from CH3 gives an orbital that can accommodate a pair of electron to form a 4c-2e
bond.

2.4.1 https://chem.libretexts.org/@go/page/172727
The lower energy of the C orbital compared with the Li orbitals indicates that the bonding pair of electrons will be associated
primarily with the CH3 group, thus supporting the carbanionic character of the molecule. Some analysis has indicated that about
90% ionic character for the Li-CH3 interaction.

The interaction between an sp3 orbital from a methyl group and the three 2s orbitals of the Li atoms in a triangular face of
Li4(CH3)4 to form a totally symmetric 4c,2e bonding orbital. The next higher orbital is non-bonding and the uppermost is
antibonding.
Me2Be and Me2Mg exist in a polymeric structure with two 3c,2e-bonding CH3 bridges between each metal atom.

This page titled 2.4: Structure and Bonding is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

2.4.2 https://chem.libretexts.org/@go/page/172727
CHAPTER OVERVIEW

3: Organometallic Chemistry of p-block Elements


3.1: Reactions of Organometallic Compounds
3.2: Organometallic Compounds of Boron and Aluminium
3.3: Organometallic Compounds of Gallium and Indium
3.4: Zeigler Natta Polymerization Catalysts
3.5: Organosilicon and Organogermanium Compounds
3.6: Organotin and Organolead Compounds

This page titled 3: Organometallic Chemistry of p-block Elements is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

1
3.1: Reactions of Organometallic Compounds
Learning Objectives
In this section you will learn the following
Chemical properties of organometallic compounds.
Reactivity patters and their electronic properties.
Electronic properties verses chemical reactivity.
Application of alkyl lithium in organic transformations.

Reactions of organometallic compounds


The reactions of organometallic compounds of electropositive elements are dominated by factors such as the carbanion character of
the organic moiety and the availability of a coordination site on the central metal atom. Reaction patterns:

a. Oxidation
All organometallic compounds are potentially reducing agents.
Those of electropositive elements are in fact very strong reducing agents (many of them are pyrophoric in nature).
The strong reducing character also presents a potential explosion hazard if the compounds are mixed with larger amount of
oxidzing agents.
Why is it so?
All organometallic compounds of the electropositive metals that have unfilled valence orbitals, or that readily dissociate into
fragments with unfilled orbitals are pyrophoric,e.g., Li (CH ) , Zn(CH ) , B(CH ) and Al (CH )
4 3 4 3 2 3 3 2 3 6

Volatile pyrophoric compounds, such as B(CH ) , may be handled in vacuum line and, inert atmosphere techniques are used for
3 3

less volatile but air-sensitive compounds. Compounds such as Si(CH ) and Sn(CH ) which do not have low-lying empty
3 4 3 4

orbitals, require elevated temperatures to initiate combustion, and can be handled in air.
The combustion of many organometallic compounds takes place by a radical chain mechanism.

b. Nucleophilic character
The partial negative charge of an organic group attached to an electropositive metal makes it a strong nucleophile and Lewis base.
This is referred to as its carbanion character even though the compound itself is not ionic. Alkyllithium and alkylaluminium
compounds and Grignard reagents are the most common carbanion reagents in laboratory-scale synthetic chemistry. The carbanion
character diminishes for the less metallic boron and silicon.
The carbanion character finds many synthetic applications.

3.1.1 https://chem.libretexts.org/@go/page/172731
Where X = halide, E = B, Si, Ge, Sn, Pb, As and Sb

c. Lewis acidity
Due to the presence of unoccupied orbitals on the metal atom, electron-deficient organometallic compounds are observed to be
Lewis acids.
e.g

B(C H ) + LiC H ⟶ Li[B(C H ) ]


6 5 3 6 5 6 5 4

This reaction may be viewed as the transfer of the strong base C6H5-from the weak Lewis acid Li+to the stronger acid B(III).
Organometallic species that are bridged by organic groups can also serve as Lewis acids and, in the process, bridge cleavage can
take place.

Al (CH ) + 2 N(C H ) ⟶ 2 (CH ) AlN(C H )


2 3 6 2 5 3 3 3 2 5 3

Electron deficient organometallic species are Lewis acids.

3.1.2 https://chem.libretexts.org/@go/page/172731
Problems
1. Name each of the following compounds and classify them: (a) SiH(CH3)3, (b) BCl(C6F5)2, (c) Al2Cl2(C6H5)4, (d) Li4(C4H9)4,
(e) Rb(CH2H3).
Solution
a. trimethylsilane (tetrahedral monomer, electron-precise);
b. bis(pentafluorophenyl)chloroborane (trigonal monomer, electron-deficient);
c. tetraphenyldichlorodialuminum (two Al-Cl-Al bridges, in this structural form it is electron –precise);
d. butyllithium or more precisely, tetrabutyltetralithium (tetrahedral Li4 array with a phenyl carbon bridging each face, electron
deficient);
e. ethylrubidium, salt-like.
2. Sketch the structures of: (a) methyl lithium, (b) trimethyl boron, (c) hexamethyldialuminum, (d) tetramethylsilane, (e)
trimethylarsane, and (f) tetraphenylarsonium.
Solution
a. methyl lithium: Li tetrahedron with each face capped by CH3 [see chapter 6]
b. trimethyl boron: planar triangular array of B and C
c. hexamethyldialuminum:four terminal CH3 and two CH3 bridges in a diborane-like structure
d. tetramethylsilane: tetrahedral
e. trimethylarsane: pyramidal
f. tetraphenylarsonium: pseudotetrahedral

This page titled 3.1: Reactions of Organometallic Compounds is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

3.1.3 https://chem.libretexts.org/@go/page/172731
3.2: Organometallic Compounds of Boron and Aluminium
Learning Objectives
In this lecture you will learn the following
Preparation and reactivity of organoboron and organo aluminium compounds.
Influence of Lewis acidity on structural features

Organoboron Compounds
BMe3 is colorless, gaseous ( b.p. -22 °C), and is monomeric. It is pyrophoric but not rapidly hydrolyzed by water.
Alkylboranes can be synthesized by metathesis between BX3 and organometallic compounds of metals with low electronegativity,
such as RMgX or AlR3.

BF + 3 CH MgBr ⟶ B(CH ) + 3 MgBrF(solvent used : dibutyl ether)


3 3 3 3

Why dibutyl ether as a solvent: Has much lower vapor pressure than BMe3 and as a result the separation by trap-to-trap distillation
on a vacuum line is easy.
Also, there is a very weak association between BMe3 and OBu2(Me3B:OBu2).
Although, trialkyl- and triarylboron compounds are mild Lewis acids, strong carbanion reagents lead to anions of the type [BR4]-.
Example, Na[BPh]4: The bulky anion hydrolyses very slowly in neutral or basic water and is useful for the preparation of large
positive cations.
+
Na [BPh] +K ⟶ K[BPh]
4 4

K[BPh]4 is insoluble, used for the gravimetric estimation (determination) of potassium, an example of the low solubility of large-
cation and large-anion salts in water/

Organohaloboron compounds are more reactive than simple trialkylboron compounds.


Preparation:

2 BCl + 6 AlR ⟶ 3 R BCl + 6 AlR Cl(metathesis)


3 3 2 2

(diborane)

2 BCl + BMe −−−−−−→ 3 BMeCl (redistribution reaction)


3 3 2

Reactions: (Protolysis reactions with ROH, R2NH and other reagents)

3 BMeCl + 2 HNR ⟶ BMe (NR ) + [ R NH ]Cl


2 2 2 2 2 2

BMe Cl + Li(C H ) ⟶ BMe (C H ) + LiCl


2 4 9 2 4 9

Organoaluminium compounds
With less bulky alkyl groups, dimerization occurs and one of the distinguishing features of alkyl bridge is the small Al-C-Al angle,
which is ~ 75°.
The 3c,2e bonds are very weak and tend to dissociate in the pure liquid which increases with increase in the bulkiness of the alkyl
group.

3.2.1 https://chem.libretexts.org/@go/page/172732
Al (CH ) −
↽⇀
− 2 Al (CH )
2 3 6 3 3

K= 1.52 x 10-8

Al (C H ) −
↽⇀
− 2 Al (C H )
2 4 9 6 4 9 3

K= 2.3 x 10-4
Perpendicular orientation of pheynl groups in Al2Ph6
Triphenylaluminium exists as a dimer with bridging η1-phenyl groups lying in a plane perpendicular to the line joining the two Al
atoms.

This structure is favored partly on steric grounds and partly by supplementation of the Al-C-Al bond by electron donation from the
phenyl π-orbitals to the Al atoms.
Tendency for bridging: X > Ph > alkyl
3c,2e bonds formed by a symmetric combination of Al and C orbitals

An additional interaction between the pπ orbital on C and an antisymmetric combination of Al orbitals.

Synthesis

3.2.2 https://chem.libretexts.org/@go/page/172732
Very useful as alkene polymerization catalysts and chemical intermediates.
Expensive carbanion reagents for the replacement of halogens organic groups by metathesis.
Laboratory scale preparations involves:

2 Al + 3 Hg (CH ) ⟶ Al (CH ) + 3 Hg
3 2 2 3 6

Commercial method:

2 Al + CH Cl ⟶ Al Cl (CH )
3 2 2 3 4

Al Cl (CH ) + 6 Na ⟶ Al (CH ) + 2 Al + 6 NaCl


2 2 3 4 2 3 6

Commercial method for ethylaluminium and higher homologs:


60−110°C

2 Al + 3 H + 6 RHC=CH −−−−−−−→ 2 Al (CH CH R)


2 2 2 2 2 6
110−200atm

The reaction probably proceeds by the formation of a surface Al—H species that adds across the double bond of the alkene in a
hydrometallation reaction.

Reactions:
Alkylaluminum compounds are mild Lewis acids and form complexes with ethers, amines and anions. When heated, often β-
hydrogen elimination is responsible for the decomposition of ethyl and higher alkylaluminium compounds. E.g. Al(iC4H9)3
Tendency towards bridging structure is: PR2-> X -> H -> Ph-> R-.

Problems:
1. Propose a structure for Al2(Me)4Cl2.
Solution:
Similar to diborane:

2. For these compounds (H3Si)2O and (CH3CH2)2O, which do you expect to have the lower force constant, Si-O-Si bending or C-
O-C bending?

3.2.3 https://chem.libretexts.org/@go/page/172732
Solution:
Lower force constant for Si-O-Si bending.
3. Explain how the difference in reactivity between Al-C and Si-C bonds with O-H groups leads to the choice of different strategies
for the synthesis of aluminum and silicon alkoxides.
Solution

Al Me + 6 MeOH ⟶ 2 Al (OMe) + 6 CH
2 6 3 4

For reaction of Al2Me6 with alcohols, see the text book by Shriver and Atkins.
Tetramethylsilane does not react with methyl alcohol. Therefore, the appropriate reagent is tetrachlorosilane and the reaction is:

SiCl + 4 MeOH ⟶ Si (OMe) + 4 HCl


3 4

4. Compare formulas of the most stable hydrogen compounds of germanium and arsenic with those of their methyl compounds.
Can the differences be explained in terms of the relative electronegativities of C and H?
Solution
GeH4, GeR4; AsH3, AsR3
The stability of hydrides and alkyls are very similar for each element. This may due to similar H and C electronegativity.
5. To buy from a chemical company, the price of trimethylaluminum is higher than that of triethylaluminum. Is it due to the
methods of synthesis? Rationalize the price difference.
Solution:
Triethylaluminum can be made in larger quantities by direct reaction of aluminum, hydrogen gas and ethane gas which is a cheaper
method.
60−110°C

2 Al + 3 H + 6 RHC=CH −−−−−−−→ 2 Al (CH CH R)


2 2 2 2 2 6
110−200atm

Preparation of trimethyaluminum involves a more expensive route such as MeCl and aluminum to form Al2Me4Cl2 followed by
treatment with sodium metal. The sodium metal and MeCl are not cheap as compared to ethane and hydrogen gases.

2 Al + CH Cl ⟶ Al Cl (CH )
3 2 2 3 4

Al Cl (CH ) + 6 Na ⟶ Al (CH ) + 2 Al + 6 NaCl


2 2 3 4 2 3 6

This page titled 3.2: Organometallic Compounds of Boron and Aluminium is shared under a CC BY-NC-SA 4.0 license and was authored,
remixed, and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts
platform; a detailed edit history is available upon request.

3.2.4 https://chem.libretexts.org/@go/page/172732
3.3: Organometallic Compounds of Gallium and Indium
Learning Objectives
In this section you will learn the following
Chemistry of gallium and indium.
How to stabilize M—M multiple bonds.

3 Li (C H ) + 4 GaCl ⟶ 2 LiCl + 4 Ga (C H )
4 2 5 4 3 2 5 3

Trialkylgallium compounds are mild Lewis acids, so the corresponding metathesis reaction in ether produces the complex
(C2H5)2OGa(C2H5)3. Similarly excess use of C2H5Li leads to the salt, Li[Ga(C2H5)4].

Li (C H ) + GaCl ⟶ 3 LiCl + Li[Ga (C H ) ]


4 2 5 4 3 2 5 4

Alkylindium and alkylthalium compounds may be prepared similar to gallium analogs. InMe3 is monomeric in the gas phase and in
the solid the bond lengths indicate that association is very weak. Partial hydrolysis of TlMe3 yields the linear (MeTiMe]+ion, which
is isoelectronic and isostructural with HgMe2.
CpIn and CoTl exist as monomers in the gas phase but are associated in solids {Inert-pair effect is displayed for In and Tl}. CpTl is
useful as a synthetic reagent in organometallic chemistry because it is not as highly reducing as NaCp.

Species of the type R4E2(single E-E bond) and [R4E2] - (with E-E bond order of 1.5) can be prepared for Ga and In with bulky R
groups (R = (Me3Si)2CH, 2,4,6-iPr3C6H2), and reduction of [(2,4,6-iPr3C6H2)4Ga2] to [(2,4,6-iPr3C6H2)4Ga2] - is accompanied by a
shortening of the Ga—Ga bond from 252-234 pm.

3.3.1 https://chem.libretexts.org/@go/page/172733
Using even bulkier substituents, it is possible to prepare gallium(I) compounds, RGa starting from GaI. No structural data are yet
available for these monomers
(We are working on it).

Crystallized as dimer but reverts to monomer when dissolved in cyclohexane.

Interest in organometallic comounds of Ga, In and Tl is mainly because of their potential use as precursors to semiconducting
materials such as GaAs and InP. Volatile compounds can be used in the growth of thin films by MOCVD (metal organic chemical
vapor deposition) or MOVPE (metal organic vapor phase epitaxy) techniques. Precursors include appropriate Lewis base adducts
of metal alkyls, e.g. Me3Ga.NMe3 and Me3In.PEt3. Thermal decomposition of gaseous precursors result in semiconductors (III-V
semiconductors) which can be deposited in thin films.
1000−1150K

Me Ga(g) + AsH (g) −−−−−−−→ GaAs(s) + 3 CH (g)


3 3 4

III-V semiconductors: Derive their name from the old groups 13 and 15, and include AlAs, AlSb, GaP, GaAs, GaSb, InP, InAs and
InSb. Off these GaAs is of the greatest commercial interest. Although Si is probably the most important commercial
semiconductor, a major advandage of GaAs over Si is that the charge carrier mobility is much greater. This makes GaAs suitable
for high-speed electronic devices.
Another important difference is that GaAs exhibits a fuly allowed electronic transition between valence and conduction bands (i.e.
it is direct band gap semiconductor) whereas Si is an indirect band gap semiconductor. The consequence of difference is that GaAs
(also other III-V types) are more suited than Si for use in optoelectronic devices, since light is emitted more efficiently. The III-Vs
have important applications in light-emitting diodes (LEDs).

3.3.2 https://chem.libretexts.org/@go/page/172733
Problems:
1. Predict the structure of monomeric, Cp3Ga; polymeric Cp3In and CpIn.
Solution:
See the articles Organometallics 1985, 4, 751.
Inorg. Chem. 1972, 11, 2832.
Organometallics 1988, 7, 105.
2. The reaction of [(R3C)4Ga4] ( R = a bulky substituent) (i) with I2 in boiling hexane results in the formation of [(R3C)GaI]2(ii)
and [(R3C)GaI2]2(iii). Draw the structure and state the oxidation state for (i) - (iii).
Solution:

3. The I2 oxidation of [(tBu}4In4] leads to the formation of the InII compound [(tBu}4In4I4] in which each indium atom retains a
tetrahedral environment. Draw the correct structure.
Solution

This page titled 3.3: Organometallic Compounds of Gallium and Indium is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

3.3.3 https://chem.libretexts.org/@go/page/172733
3.4: Zeigler Natta Polymerization Catalysts
Learning Objectives
In this lecture you will learn the following
Olefin polymerization.
Mechanism involved in polymerization process.

Ziegler Natta Polymerization Catalysts

Insertion of aluminum alkyls into olefins was studied by Ziegler. During the systematic investigation of olefin polymerization,
Ziegler realized that the most effective catalyst is the combination of TiCl4/AlEt3which can polymerize ethylene at pressure as low
as 1 bar. The application of Ziegler method to the polymerization of propylene and its establishment and the investigation of bulk
properties was carried out by Natta and hence the methodology is called Ziegler-Natta process.

Important discovery: R3Al + Lewis acids.

3.4.1 https://chem.libretexts.org/@go/page/172734
In the absence of reaction mechanism with solid proof, it is presumed that the reaction is due to the heterogeneous catalysis in
which fibrous TiCl3, alkylated on its surface is considered to be the active catalyst species.

This page titled 3.4: Zeigler Natta Polymerization Catalysts is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

3.4.2 https://chem.libretexts.org/@go/page/172734
3.5: Organosilicon and Organogermanium Compounds
Learning Objectives
In this lecture you will learn the following
Organosilicon and organogermanium compounds.
Compounds with Si=Si and Ge=Ge bonds.

Organosilicon and organogermanium compounds


Organosilicon compounds are extensively studied due to the wide range of commercial applications as water repellents, lubricants,
and sealants. Many oxo-bridged organosilicon compounds can be synthesized. e.g. (CH3)3Si—O—Si(CH3)3 which is resistant to
moisture and air.

The lone pairs on O are partially delocalized into vacant σ*- orbitals of Si, as a result the directionality of the Si-O bond is reduced
making the structure more flexible.

This flexibility permits silicone elastomers to remain rubber-like down to very low temperature.
Delocalization also accounts for low basicity of an O atom attached to silicon as the electrons needed for the O atom to act as a
base are partially removed.

The planarity of N(SiH3)3 is also explained by the delocalization of the lone pair on N which makes it very weakly basic.

nMeCl + Si/Cu ⟶ Me n SiCl


4 −n

SiCl + 4 RLi ⟶ R Si
4 4

SiCl + RLi ⟶ RSiCl


4 3

SiCl + 2 RMgCl ⟶ R SiCl + 2 MgCl


4 2 2 2

\[\ce{Me2SiCl2 + tBuBi -> tBuMe2SiCl + LiCl}|]


Si—C bonds are relatively strong (bond enthalpy is 318 kJ mol-1) and R4Si derivatives possess high thermal stabilities. Et4Si on
chlorination gives (ClCH2CH2)4Si, in contrast to the chlorination of R4Ge or R4Sn which yields RnGeCl4-n or RnSnCl4-n.
Me2SiCl2 on hydrolysis produce silicones.
1
Me SiCl + NaCp ⟶ (η −Cp)SiHMe
3 3

1
(η -C5Me5) 2SiBr2 on treatment with anthracene/potassium gives Cp*2Si
Solid state structure of Cp*2Si consists of two independent molecules which differ in the relative orientations of the Cp rings.
In one molecule, they are parallel and staggered whereas in the other, they are tilted with an angle of 167°at Si.
The reaction between R2SiCl2 and alkali metal or alkali naphthalides give cyclo-(R2Si)n by loss of Cl-and Si—Si bond formation.
Bulky R groups favour small rings [e.g. (2,6-Me2C6H3)6Si3 and tBu6Si3] while smaller R groups encourage the formation of large
rings [Me12Si6, Me14Si7 and Me32Si16]

[P h2SiC l2 + Li(SiP h2)5Li− > cyclo − P h12Si6 + 2LiC l

Bulky substituents stabilize R2Si=SiR2 compounds. The sterically demanding 2,4,6-iPr3C6H2 provided first example of compound
containing conjugated Si=Si bonds.

3.5.1 https://chem.libretexts.org/@go/page/172735
Has s-cis configuration in both solution and the solid state.
Similar germanium compounds are also known

*The spatial arrangement of two conjugated double bonds about the intervening single bond is described as s- cis if synperiplanar
and s-trans if antiperiplanar.

This page titled 3.5: Organosilicon and Organogermanium Compounds is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

3.5.2 https://chem.libretexts.org/@go/page/172735
3.6: Organotin and Organolead Compounds
Learning Objectives
In this section you will learn the following
Organotin and organolead compounds and their preparation.
Bonding in tin compounds with Sn=Sn double bonds.
Uses and environmental issues with tin compounds.
Reactivity of tetraethyl lead.
Structural features of organolead compounds.

Organotin and organolead compounds


Preparation of Sn(IV) derivatives

R 2O

3 SnCl + 4 R Al −−−→ 3 R Sn + 4 AlCl


3 3 4 3

500K

R Sn + SnCl ⟶ R SnCl + RSnCl −−−→ 2 R SnCl


4 4 3 3 2 2

SnCl + 4 RMgBr ⟶ R Sn + 4 MgBrCl


4 4

SnCl + Ph Hg ⟶ Ph SnCl + Hg
2 2 2 2

Tin(II) organometallics of the type R2Sn, containing Sn-C σ-bonds, are stabilized only if R is sterically demanding.

SnCl + 2 Li[ (Me Si) CH] ⟶ [ (Me Si) CH] Sn


2 3 2 3 2 2

(monomeric in solution and dimeric in solid state). But the dimer does not possess a planar Sn2R4framework unlike an analogous
alkene, and Sn—Sn bond distance (267 pm) is shorter than a normal Sn—Sn single bond (276 pm).
Sn2R4 has a trans bent structure with a weak Sn=Sn double bond

Look into the reactions of R3SnCl with various reagents to form useful tin containing starting materials
The first organotin(II) hydride was reported only in 2000.

3.6.1 https://chem.libretexts.org/@go/page/172736
Shows dimeric structure in the solid state containing hydride bridges (Sn-Sn = 312 pm).

Commercial uses and environmental problems


Organotin(II) compounds find wide range of applications due to their catalytic and biocidal properties.
n
Bu3SnOAc is an effective fungicide and bactericide and also a polymerization catalyst.
n
Bu2Sn(OAc)2 is used as a polymerization catalyst and a stabilizer for PVC.
n
Bu3SnOSnnBu3 is algicide, fungicide and wood-preserving agent.
n
Bu3SnCl is a bactericide and fungicide.
Ph3SnOH used as an agricultural fungicide for crops such as potato, sugar beet and peanuts.
The cylic compound (nBu2SnS)3 is used as a stabilizer for PVC.
Tributyltin derivatives have been used as antifouling agents, applied to the underside of ships’ hulls to prevent the build-up of, for
example, barnacles.
Global legislation now bans or greatly restricts the use of organotin-based anti-fouling agents on environmental grounds.
Environmental risks associated with the uses of organotin compounds as pesticides, fungicides and PVC stabilizers are also a cause
for concern.
*A barnacle is a type of arthropod belonging to infraclass Cirripedia in the sub-phylum Crustacea, and is hence related to crabs and
lobsters.

Organolead compounds
Tetraethyllead

4 NaPb + 4 EtCl ⟶ Et Pb + 3 Pb + 4 NaCl[at 373K in an autoclave]


4

Laboratory Scale,
Et2O

2 PbCl + 4 RMgBr −−−−−−→ 2 (R Pb) ⟶ R Pb + Pb


3 2 4
−4MgBrC l

Thermolysis leads to radical reactions.

Et Pb ⟶ Et Pb + Et
4 3

2 Et ⟶ n−C H
4 10

Et Pb + Et ⟶ C H + Et PbH
3 2 4 3

Et Pb + Et Pb ⟶ H + Et Pb + Et PbCH CH
3 4 2 3 3 2 2

Tetraalkyl and tetraaryl lead compounds are inert with respect to attack by air and water at room temperature. WHY ????

3.6.2 https://chem.libretexts.org/@go/page/172736
Me3PbCl consists of linear chain
Solid state structure of Cp2Pb shows polymeric nature, but in the gas phase, discrete Cp2Pb molecules are present which possess
the bent structure similar to silicon analogue.
R2Pb=PbR2 are similar to analogues tin compounds

Problems
1. Find out the structures of (Me3SiCH2)3SnF and Me2SnF2
Solution: use VSEPR theory

This page titled 3.6: Organotin and Organolead Compounds is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

3.6.3 https://chem.libretexts.org/@go/page/172736
CHAPTER OVERVIEW

4: Organoelement Compounds of Group 15


4.1: Organometallic Compounds of As(V) and Sb(V)
4.2: Organometallic Compounds of As(III) and Sb(III)
4.3: Phosphines

This page titled 4: Organoelement Compounds of Group 15 is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

1
4.1: Organometallic Compounds of As(V) and Sb(V)
Learning Objectives
In this section you will learn the following
Organoaresnic and organoantimony compounds.
Preparation and reactivity of pentavlent As and Sb compounds.

Organic chemistry of non-metal phosphorus, metalloids such as arsine and antimony along with metallic element bismuth is termed
as organoelement chemistry. The importance given to organoarsenic compounds earlier due to their medicinal values was waded
out after antibiotics were discovered and also their carcinogenic and toxic properties were revealed. Also, the synthetically
important organometallic compounds of group 13 and 14 masked the growth of group 15 elements. However, the organoelement
compounds of phosphorus, arsenic and antimony find usefulness as ligands in transition metal chemistry due to their σ-donor and
π-acceptor abilities which can be readily tuned by simply changing the substituents. These donor properties are very useful in
tuning them as ligands to make suitable metal complexes for metal mediated homogeneous catalysis. Although organoelement
compounds can be formed in both +3 (trivalent and tricoordinated) and +5(pentavalent and tetra or pentacoordinated) oxidation
states, trivalent compounds are important in coordination chemistry.
For organoelement compounds of group 15, the energy of E—C bond decreases in the order, E = P > As > Sb > Bi, and in the same
sequence E—C bond polarity increases.

Organometallic compounds of As(V) and Sb(V)


Due to the strong oxidizing nature of pentahalides, the direct alkylation or arylation to generate ER5 is not feasible, but can be
prepared in two steps.
A few representative methods of preparation are given below:
C l2 MeLi

Me As −→
− Me AsCl −−−→ Me As
3 3 2 5
Et2O

P hLi

Ph Sb + PhI ⟶ Ph SbI −−−→ Ph Sb


3 4 5
−LI

SO2C l2 P hMgX

Ph Bi −−−−→ Ph BiCl −−−−−→ Ph Bi


3 3 2 5
−SO2 −MgXC l

Structures and properties


Pentaalkyl or pentaaryl derivatives are moderately thermally stable. On heating above 100°C, they form trivalent compounds as
shown below:
T >100°

Me As −−−−→ Me As + CH + CH CH
5 3 4 2 2

T >200°

Ph Sb −−−−→ Ph Sb + Ph−Ph
5 3

Reaction with water,


+
Me As H O ⟶ Me AsOH + MeH
5 2 4

Pentavalent compounds readily form “tetrahedral onium” cations and “octahedral and hexacoordinatged ate” anions.

Ph E + BPh ⟶ [ Ph E][ BPh ](E = As, Sb, Bi)


5 3 4 4

Ph E + LiPh ⟶ Li[ EPh ]


5 6

In solid state, Ph5As adopts trigonal bipyramidal geometry, whereas Ph5Sb prefers square based pyramidal geometry although the
energy difference between the two is marginal.

The salts of the type [R4E]+ adopt tetrahedral geometry, whereas hexacoordinated anions [R6E]-assume octahedral geometry.

4.1.1 https://chem.libretexts.org/@go/page/172738
Mixed organo-halo compounds of the type RnEX5-n adopt often dimeric structures due to the presence of lone pairs of electrons on
X which can readily coordinate to the second molecule. The following structural types can be anticipated.

The thermal stability of RnEX5-n decreases with decreasing ‘n’. Thermal reactions are essentially the reverse reactions of addition
reactions used in the preparation of R5E.
ΔT

R SbX −→
− R SbX + RX
3 2 2

C O2

Ph AsCl −−−→ Ph AsCl + Cl


3 2 2 2
100°C

50°C

Me AsCl −−−→ MeAsCl + MeCl


2 3 2

This page titled 4.1: Organometallic Compounds of As(V) and Sb(V) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

4.1.2 https://chem.libretexts.org/@go/page/172738
4.2: Organometallic Compounds of As(III) and Sb(III)
Learning Objectives
In this lecture you will learn the following
Preparation of trivalent compounds.
Mono and bis derivatives.
Reaction of organo arsenic and antimony compounds.
Structural features of organolead compounds.

Organometallic compounds of As(III) and Sb(III)


Direct synthesis
Mono- derivatives
ΔT

2 As + 3 MeBr −→
− Me AsBr + MeAsBr
2 2
Cu

Me AsBr + PhLi ⟶ Me AsPh


2 2

MeAsBr + 2 PhLi ⟶ MeAsPh


2 2

EX + 3 RMgX ⟶ R E + 3 MgX
3 3 2

EX + 3 RLi ⟶ R E + 3 LiX
3 3

Bis derivatives:

In a similar way, a variety of bisphosphines and arsines can be generated.


Reactions of trialkyl derivatives, R3E

4.2.1 https://chem.libretexts.org/@go/page/172739
The transition metal chemistry of R3E, phosphines, arsines or stibines has been extensively studied because of their distinct donor
and acceptor properties. Among them, the phosphines or tertiary phosphines (R3P) are the most valuable ligands in metal mediated
homogeneous catalysis. Interestingly, the steric and electronic properties can be readily tuned by changing the substituents on
phosphorus atoms. Chapter 16 is fully dedicated to the chemistry of phosphines.
Properties
Trialkyl derivatives are highly air-sensitive liquids with low boiling points and some of them are even pyrophyric. Triphenyl
derivatives are solids at room temperature and are moderately stable and oxidizing agents such as KMnO4, H2O2 or TMNO are
needed for oxidation to form Ph3E=O.

Cyclic and acyclic derivatives containing E—E bonds


E—E single bonds:

The E—E bond energies suggest that they do not have greater stability and the stability decreases down the group.
The simplest molecules include Ph2P—PPh2, Me2As—AsMe2 prepared by coupling reactions:

Me AsH + Me AsCl −−−→ Me As−AsMe


2 2 2 2
−H C l

N a,N H 3

2 Ph BiCl −−−−−→ Ph Bi−BiPh


2 2 2

The weakness of E—E bonds accounts for many interesting reactions and a few of such reactions are listed below:

4.2.2 https://chem.libretexts.org/@go/page/172739
Cyclic and polycyclic derivatives can be prepared by employing any of the following methods:

Problems:
1. Confirm that the octahedral structure of [Ph6Bi]- is consistent with VSEPR theory.
Solution:
Octahedral similar to PF6-
5 (Bi valence electrons) + 6 (each Ph ) + 1 (-ve charge) = 12 electrons
i.e. six pairs, octahedral geometry
2. Comment on the stability of BiMe3 and Al2(iBu)6 with respect to their thermal decomposition and give chemical equations for
their decomposition.

Solution:
Similar to other heavy p-block elements, Bi—C bonds are weak and readily undergo homolytic cleavage. The resulting methyl
radicals will react with other radicals or form ethane.

2 BiMe ⟶ 2 Bi + 3 CH −CH
3 3 3

i
The Al2( Bu)6 dimer readily dissociates. At elevated temperature dissociation is followed by β-hydrogen elimination. This type of
elimination is common for organometallic compounds that have alkyl groups with β-hydrogens, can form stable M—H bonds, and
can provide a coordination site on the central metal.
The decomposition reaction is:
Δ

Al (iBu) −
→ 2 Al (iBu) ⟶ [ΔΔ]Al (iBu) H + (CH ) C=CH
2 6 3 2 3 2 2

3. Using a suitable Grignard reagent, how would you prepare (i) MeC(Et)(OH)Ph; (ii) AsPh3.
Solution
i. Add a Grignard reagent to a C=O bond, then acidify.

4.2.3 https://chem.libretexts.org/@go/page/172739
Several possibilities, e.g.
Me−C(O)−Et + PhMgBr ⟶ Me−C(OMgBr)(Et)(Ph) ⟶ MeC(Et)(OH)Ph or Me−C(O)−Ph + EtMgBr

⟶ etc

ii. AsCl
3
+ 3 PhMgBr ⟶ AsPh
3
+ 3 MgBrCl

This page titled 4.2: Organometallic Compounds of As(III) and Sb(III) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

4.2.4 https://chem.libretexts.org/@go/page/172739
4.3: Phosphines
Learning Objectives
In this lecture you will learn the following
Classification of ligands.
Nature of bonding in phosphines.
Steric and electronic properties of phosphines.
Bonding in phosphines and CO.
Cone angle and its application in catalysis.

Classification of Ligands by donor atoms


Ligand is a molecule or an ion that has at least one electron pair that can be donated. Ligands may also be called Lewis bases; in
terms of organic chemistry, they are ‘nucleophiles’.

Metal ions or molecules such as BF3 (with incomplete valence electron shells (electron deficient) are called Lewis acids or
electrophiles).

Why do molecules like H2O or NH3 give complexes with ions of both main group and transition metals. E.g [Al(OH2)6]3+ or
[Co(NH3)6]3+

Why other molecules such as PF3 or CO give complexes only with transition metals.

Although PF3 or CO give neutral molecules such as Ni(PF3)4 or Ni(CO)4 or Cr(CO)6.

Why do, NH3, amines, oxygen donors, and so on, not give complexes such as Ni(NH3)4.
Classical or simple donor ligands
Act as electron pair donors to acceptor ions or molecules, and form complexes of all types of Lewis acids, metal ions or molecules.

Non-classical ligands, π-bonding or π-acid ligands: Form largely with transition metal atoms.

In this case special interaction occurs between the metals and ligands

These ligands act as both σ-donors and π-acceptors due to the availability of empty orbitals of suitable symmetry, and energies
comparable with those of metal t2g (non-bonding) orbitals.
e.g. Consider PR3 and NH3: Both can act as bases toward H+, but P atom differs from N in that PR3 has σ* orbitals of low energy,
whereas in N the lowest energy d orbitals or σ* orbitals are far too high on energy to use.

Consider CO that do not have measurable basicity to proton, yet readily reacts with metals like Ni that have high heats of
atomization to give compounds like Ni(CO)4.

Ligands may also be classified electronically depending upon how many electrons that they contribute to a central atom. Atoms or
groups that can form a single covalent bond are one electron donors.

EXAMPLES: F, SH, CH3 etc.,


Compounds with an electron pair are two-electron donors
EXAMPLE: NH3, H2O, PR3 etc.,

4.3.1 https://chem.libretexts.org/@go/page/172740
Bonding in Metal –Carbonyl and Metal-Phosphines

4.3.2 https://chem.libretexts.org/@go/page/172740
Steric factors in phosphines (Tolman’s cone angle)
Cone angle is very useful in assessing the steric properties of phosphines and their coordination behavior.

The electronic effect of phosphines can be assessed by IR and NMR spectroscopic data especially when carbonyls are co-ligands.
In a metal complex containing both phosphines and carbonyl, the ν(CO) frequencies would reveal the σ-donor or π–acceptor
abilities of phosphines. If the phosphines employed are strong σ-donors, then more electron density would move from M (t2g
orbitals)- π*(CO) and as a result, a lowering in the ν(CO) is observed. In contrast, if a given phosphine is a poor σ-donor but strong
π -acceptor, then phosphine(σ*-orbitals) also compete with CO for back bonding which results in less lowering in ν(CO) frequency.
Another important aspect is the steric size of PR3 ligands, unlike in the case of carbonyls, which can be readily tuned by changing
R group. This is of great advantage in transition metal chemistry, especially in metal mediated catalysis, where stabilizing the
metals in low coordination states is very important besides low oxidation states. This condition can promote oxidative addition at
the metal centre which is an important step in homogeneous catalysis. The steric effects of phosphines can be quantified with
Tolman’s cone angle.
Cone angle can be defined as a solid angle at metal at a M—P distance of 228 pm which encloses the van der Waal’s surfaces of all
ligand atoms or substituents over all rotational orientations. The cone angles for most commonly used phosphines are listed in the
following table.

Phosphine Cone Angle (°)

PH3 87

PF3 104

P(OMe)3 107

PMe3 118

PMe2Ph 122

PEt3 132

PPh3 145

PCy3 170
t
P(Bu )3 182

P(mesityl)3 212

Phosphines with different cone angles versus coordination number for group 8 metals:

ML4 ML3 ML2

(Me3P)4Ni

(Me3P)4Pd

(Me3P)4Pt (Ph3P)3Pt (tert-Bu3P)2Pt

4.3.3 https://chem.libretexts.org/@go/page/172740
Tolman Angle and Catalysis
Sterically demanding phosphine ligands can be used to create an empty coordination site (16 VE complexes) which is an important
trick to fine tune the catalytic activity of phosphine complexes.

This page titled 4.3: Phosphines is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S. Balakrishna &
Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

4.3.4 https://chem.libretexts.org/@go/page/172740
CHAPTER OVERVIEW

5: Group 12 Elements
5.1: Organometallic Compounds of Zinc and Cadmium
5.2: Organometallic Compounds of Mercury

This page titled 5: Group 12 Elements is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

1
5.1: Organometallic Compounds of Zinc and Cadmium
Learning Objectives
In this section you will learn the following
Organometallic compounds of zinc and cadmium.
Structural features of organozinc compounds

Organometallic compounds of zinc and cadmium


Dialkyl compounds of Zn, Cd and Hg do not associate through alkyl bridges.
Dialkylzinc compounds are only weak Lewis acids, organocadmium compounds are even weaker, and organomercury compounds
do not act as Lewis acids except under special circumstances.
The Group 12 metals form linear molecular compounds, such as ZnMe2, CdMe2 and HgMe2, that are not associated in solid, liquid
or gaseous state or in hydrocarbon solution.
They form 2c, 2e bonds. Unlike Be and Mg analogs, they do not complete their valence shells by association through alkyl bridges.
The bonding in these molecules are similar to d10 metals such as CuI, AgI and AuI with linear geometry ([N≡C-M-C≡N]-, M = Ag
or Au). This tendency is sometimes rationalized by invoking pd hybridization in the M+ ion, which leads to orbitals that favor
linear attachment of ligands (similar to spd hybridization).

The preference for the linear coordination may be due to the similarity in energy of the outer ns, np and (n-1)d orbitals, which
permits the formation of collinear spd hybrids.
The hybridization of s, pz and dz2 with the choice of phases shown here produces a pair of collinear orbitals that can be used to
form strong σ-bonds.

Organozinc and organocadmium compounds


Convenient route is metathesis with alkylaluminium or alkyllithium compounds.
With alkyllithium compounds it is the electronegativity which is decisive, whereas between Al and Zn it is hardness considerations
correctly predict the formation of softer ZnCH3 and harder AlCl pairs.

ZnCl + Al Me ⟶ ZnMe + Al Cl Me
2 2 6 2 2 2 4

Alkylzinc compounds are pyrophoric and readily hydrolyzed, whereas alkylcadmium compounds react more slowly with air. Due
to mild Lewis acidity, dialkylzinc and dialkylcadimum compounds form stable complexes with amines, especially with chelating
amines.
The Zn—C has greater carbanionic character than the Cd—C bond.
For example, addition of alkylzinc compounds across the carbonyl group of a ketone:

5.1.1 https://chem.libretexts.org/@go/page/177781
ZnMe + (CH ) C=O ⟶ (CH ) C−O−ZnCH
2 3 2 3 2 3

This reaction do not proceed with the less polar alkylcadmium or alkylmercury compounds, but organolithium, organomagnesium
and organoaluminium compounds can promote this reaction readily since all of which contain metals with lower electronegativity
than zinc.
Interestingly, the cyclodipentadienyl compounds are structurally unusual. CpZnMe is monomeric in the gas phase with a
pentahapto Cp group.
In the solid state it is associated in a zig-zag chain, each Cp group being pentahapto with respect to two Zn atoms.

Problems:
1. Do you think that the following reaction proceeds? If so, why and how?

ZnCl + Al Me ⟶ ZnMe + Al Cl Me
2 2 6 2 2 2 4

Solution
Al2Me6 being an electron deficient molecule readily exchanges two methyl groups with zinc for two chloride ions. Since chloride
ions have sufficient electron in their valence shell act as four electron donor through bridging coordination mode. Al2Cl2Me4 is no
longer an electron deficient molecule.

This page titled 5.1: Organometallic Compounds of Zinc and Cadmium is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

5.1.2 https://chem.libretexts.org/@go/page/177781
5.2: Organometallic Compounds of Mercury
Learning Objectives
In this section you will learn the following
Dialkymercury preparation.
Mercury toxicity.
Mercury poisoning.

Consider this reaction that proceeds due to both electronegativity and hardness considerations.

2 RMgX + HgX ⟶ HgR + MgX


2 2 2

Dialkylmercury compounds are very versatile starting materials for the synthesis of many organometallic compounds of more
electropositive metals by transmetallation. However, owing to high toxicity of alkylmercury compounds, other synthons are
preferred. In striking contrast to the high sensitivity of dimethylzinc to oxygen, dimethylmercury survives exposure to air.

Mercury Toxicity
The toxicity of mercury arises from the very high affinity of the soft Hg atom for sulfhydryl (—SH) groups in enzymes. Simple
mercury-sulfur compounds have been studied as potential analogs of natural systems. The Hg atoms are most commonly four-
coordinated, as in [Hg2(SMe)6]2-.
Mercury poisoning was a serious concern even from early days. Issac Newton, Alfred Stock worked in the early 20th century.
Later in 60s awareness came following the incidence of brain damage and death it caused among the inhabitants in Minamata,
Japan. Mercury from a plastic company was allowed to escape into a bay where it found its way into fish that were later eaten.
Research has shown that bacteria found in sediments are capable of methylating mercury, and that species such HgMe2 and
[HgCH3]+ enter the food chain because they readily penetrate cell walls. The bacteria appear to produce HgMe2 as a means of
eliminating toxic mercury ions through their cell walls and into the environment.

References
Inorganic Chemistry, Principles of structure and reactivity, 4th edition; 1993, J. E. Huheey, E. A. Keiter, R. L. Keiter, Addison-
Wesley Publishing Co, New York.
Advanced Inorganic Chemistry, 6th edition, 1999, F. A. Cotton, G. Wilkinson, C. A. Murillo, M. Bochmann, John Wiley and
Sons, New York.
Organometallics, A Concise Introduction, 2nd edition (revised), 1992, Ch. Elschenbroich, A. Salzer, Weinheim, Germany.
Inorganic Chemistry, 3rd Edition, 1999, D. F. Shriver, P. W. Atkins, Oxford University Press, Oxford.
Inorganic Chemistry, 2nd Edition, 2005, C.C. Housecroft and A. G. Sharpe, Pearson, Prentice Hall, England.

This page titled 5.2: Organometallic Compounds of Mercury is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

5.2.1 https://chem.libretexts.org/@go/page/177783
CHAPTER OVERVIEW

6: General Properties of Transition Metal Organometallic Complexes


6.1: 18 Valence Electron Rule
6.2: Synthesis and Stability

This page titled 6: General Properties of Transition Metal Organometallic Complexes is shared under a CC BY-NC-SA 4.0 license and was
authored, remixed, and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the
LibreTexts platform; a detailed edit history is available upon request.

1
6.1: 18 Valence Electron Rule
Learning Objectives
In this section you will learn the following
Have an insight about the stability of the transition metal complexes with respect to their total valence electron count.
Be aware of the transition metal complexes that obey or do not obey the 18 Valence Electron Rule.
Have an appreciation of the valence electron count in the transition metal organometallic complexes that arise out of the
metal-ligand orbital interactions.

The transition metal organometallic compounds exhibit diverse structural variations that manifest in different chemical properties.
Many of these transition metal organometallic compounds are primarily of interest from the prospectives of chemical catalysis.
Unlike the main group organometallic compounds, which use mainly ns and np orbitals in chemical bonding, the transition metal
compounds regularly use the (n−1)d, ns and np orbitals for chemical bonding (Figure 6.1.1). Partial filling of these orbitals thus
render these metal centers both electron donor and electron acceptor abilities, thus allowing them to participate in σ-donor/π-
acceptor synergic interactions with donor-acceptor ligands like carbonyls, carbenes, arenes, isonitriles and etc,.

Figure 6.1.1
The 18 Valence Electron (18 VE) Rule or The Inert Gas Rule or The Effective Atomic Number (EAN) Rule: The 18-valence electron
(VE) rule states that thermodynamically stable transition metal compounds contain 18 valence electrons comprising of the metal d
electrons plus the electrons supplied by the metal bound ligands. The counting of the 18 valence electrons in transition metal
complexes may be obtained by following either of the two methods of electron counting, (i). the ionic method and (ii). the neutral
method. Please note that a metal-metal bond contributes one electron to the total electron count of the metal atom. A bridging
ligand donates one electron towards bridging metal atom.

 Example 6.1.1

Ferrocene Fe(C5H5)2

6.1.1 https://chem.libretexts.org/@go/page/172752
 Example 6.1.2

Mn2(CO)10

Transition metal organometallic compounds mainly belong to any of the three categories.
a. Class I complexes for which the number of valence electrons do not obey the 18 VE rule.
b. Class II complexes for which the number of valence electrons do not exceed 18.
c. Class III complexes for which the valence electrons exactly obey the 18 VE rule.
The guiding principle which governs the classification of transition metal organometallic compounds is based on the premise that
the antibonding orbitals should not be occupied; the nonbonding orbitals may be occupied while the bonding orbitals should be
occupied.

Figure 6.1.2 : A simplified molecular orbital diagram for an octahedral transition metal complex showing σ−interactions only.

6.1.2 https://chem.libretexts.org/@go/page/172752
Figure 6.1.3 : A simplified molecular orbital diagram for an octahedral transition metal complex showing σ−and π−interactions
only.

Class I
In class I complexes, the Δo splitting is small and often applies to 3d metals and σ ligands at lower end of the spectrochemical
series. In this case the t2g orbital is nonbonding in nature and may be occupied by 0−6 electrons (Figure 6.1.2). The eg* orbital is
weakly antibonding and may be occupied by 0−4 electrons. As a consequence, 12−22 valence electron count may be obtained for
this class of compounds. Owing to small Δtetr splitting energy, the tetrahedral transition metal complexes also belongs to this class.

Class II
In class II complexes, the Δo splitting is relatively large and is applicable to 4d and 5d transition metals having high oxidation state
and for σ ligands in the intermediate and upper range of the spectrochemical series. In this case, the t2g orbital is essentially
nonbonding in nature and can be filled by 0−6 electrons (Figure 3). The eg* orbital is strongly antibonding and is not occupied at
all. Consequently, the valence shell electron count of these type of complexes would thus be 18 electrons or less.

Class III
In class III complexes, the Δo splitting is the largest and is applicable to good σ donor and π acceptor ligands like CO, PF3, olefins
and arenes located at the upper end of the spectrochemical series.
The t2gorbital becomes bonding owing to interactions with ligand orbitals and should be occupied by 6 electrons. The eg* orbital is
strongly antibonding and therefore remains unoccupied.

Problems
State the oxidation state of the metal and the total valence electron count of the following species.
1. V(C2O4)33−
Ans: +3 and 14
2. Mn(acac)3
Ans: +3 and 16
3. W(CN)83−
Ans: +5 and 17
4. CpMn(CO)3
Ans: 0 and 18

6.1.3 https://chem.libretexts.org/@go/page/172752
5. Fe2(CO)9
Ans: 0 and 18

Self Assessment test


State the oxidation state of the metal and the total valence electron count of the following species.
1. TiF62-
Ans: +4 and 12
2. Ni(en)32+
Ans: +2 and 20
3. Cu(NH3)62+
Ans: +2 and 21
4. W(CN)84-
Ans: +4 and 18
5. CH3Co(CO)4
Ans: 0 and 18

Summary
The transition metal complexes may be classified into the following three types.
i. The ones that do not obey the 18 valence electron rule are of class I type
ii. The ones that do not exceed the 18 valence electron rule are of class II and
iii. The ones that strictly follow the 18 valence electron rule.
Depending upon the interaction of the metal orbitals with the ligand orbitals and also upon the nature of the ligand position in
spectrochemical series, the transition metal organometallic compounds can form into any of the three categories.

This page titled 6.1: 18 Valence Electron Rule is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

6.1.4 https://chem.libretexts.org/@go/page/172752
6.2: Synthesis and Stability
Learning Objectives
In this lecture you will learn the following
Understand the role lead by ligands in stabilizing organometallic transition metal complexes.
Know about various synthetic methods available for preparing the organometallic transition metal complexes.
Understand the various factors like β-elimination and other bimolecular decomposition pathways that contribute to the
observed instability of the organometallic transition metal complexes.
Obtain insight about making stable organometallic transition metal complexes by suppression of the destabilizing factors
mentioned.

Ligands play a vital role in stabilizing transition metal complexes. The stability as well as the reactivity of a metal in its complex
form thus depend upon the number and the type of ligands it is bound to. In this regard, the organometallic carbon based ligands
come in diverse varieties displaying a wide range of binding modes to a metal. In general, the binding modes of the carbon-derived
ligands depend upon the hybridization state of the metal bound carbon atom. These ligands can thus bind to a metal in many
different ways as depicted below. Lastly, these ligands can either be of (a) purely σ−donor type, or depending upon the capability
of the ligand to form the multiple bonds may also be of (b) a σ−donor/π−acceptor type, in which the σ−interaction is supplemented
by a varying degree of π−interaction.

Figure 6.2.1

6.2.1 https://chem.libretexts.org/@go/page/172753
Figure 6.2.2

Figure 6.2.3

Preparation of transition metal-alkyl and transition metal-aryl complexes


The transition metal−alkyl and transition−metal aryl complexes are usually prepared by the following routes discussed below,
a. Metathesis
This involves the reactions of metal halides with organolithium, organomagnesium, organoaluminium, organotin and organozinc
reagents.

6.2.2 https://chem.libretexts.org/@go/page/172753
Of the different organoalkyl compounds listed above, the organolithium and organomagnesium compounds are strongly
carbanionic while the remaining main group organometallics like the organoalkyl, organozinc and organotin reagents are relatively
less carbanionic in nature. Thus, the main group organometallic reagents have attenuated alkylating power, that can be productively
used in partial exchange of halide ligands.
Al2Me6

TiCl −−−−→ MeTiCl


4 3

ZnMe2

NbCl −−−−→ Me NbCl


5 2 3

b. Alkene insertion or Hydrometallation


As the name implies, this category of reaction involves an insertion reaction between metal hydride and alkene as shown below.
These type reactions are relevant to certain homogeneous catalytic processes in which insertion of an olefin to M−H bond is often
observed.

c. Carbene insertion
This category represents the reaction of metal hydrides with carbenes.

d. Metallate alkylation reaction


This category represents the reaction of carbonylate anions with alkyl halides as shown below.

e. Metallate acylation reaction


This category involves the reaction of carbonylate anions with acyl halides

6.2.3 https://chem.libretexts.org/@go/page/172753
f. Oxidative addition reaction
Many unsaturated 16 VE transition metal complexes having d8 or d10 configuration undergo oxidative addition reactions with alkyl
halides. The oxidative addition reactions proceed with the oxidation state as well as coordination number of the metal increasing by
+2.

g. Addition reaction
This category involves the reaction of an activated metal bound olefin complex with a nucleophile as shown below.

Thermodynamic Stability and Kinetic Lability


The transition metal organometallic compounds are often difficult to synthesize under ordinary laboratory conditions and require
stringent experimental protocols involving the exclusion of air and moisture for doing so. As a consequence, many homoleptic
binary transition metal−alkyl and transition metal−aryl compounds like, Et2Fe or Me2Ni cannot be made under normal laboratory
conditions. More interestingly, most of the examples of transition metal−aryl and transition metal−alkyl compounds, known in the
literature, invariably contain additional ligands like η5-C5H5, CO, PR3 or halides.
For example,

Transition metal−carbon (TM−C) bond energy values are important for understanding the instability of transition metal
organometallic compounds. In general, the TM−C bonds are weaker than the transition metal−main group element (TM−MGE)
bonds (MGE = F, O, Cl, and N) and more interestingly so, unlike the TM−MGE bond energies, the TM−C bond energy values
increase with increasing atomic number. The steric effects of the ligands also play a crucial role in influencing the TM−C bond
energies and thus have to be given due consideration.
Contrary to the popular belief, the difficulty in obtaining transition metal−aryl and transition metal−alkyl complexes does primarily
arise from the thermodynamic reasons but rather the kinetic ones. β−elimination is by far the most general decomposition

6.2.4 https://chem.libretexts.org/@go/page/172753
mechanism that contribute to the instability of transition metal organometallic compounds. β−elimination results in the formation
of metal hydrides and olefin as shown below.

β−elimination can also be reversible as shown below.

The instability of transition metal organometallic compounds can arise out of kinetic lability like in the case of the β−elimination
reactions that trigger decomposition of these complexes. Thus, the suppression of the decomposition reactions provides a viable
option for the stabilization of the transition metal organometallic complexes. The β−elimination reactions in transition metal
organometallic complexes may be suppressed under any of the following three conditions.
a. Formation of the leaving olefin becomes sterically or energetically unfavorable
In the course of β−elimination, this situation arises when the olefinic bond is formed at a bridgehead carbon atom or when a double
bond is formed with the elements of higher periods. For instance, the norbornyl group is less prone to decomposition by
β−elimination because that would require the formation of olefinic double bond at a bridgehead carbon atom in the subsequent
olefin, i.e. norbornene, and which is energetically unfavorable.

b. Absence of β−hydrogen atom in organic ligands


Transition metal bound ligands that do not possess β−hydrogen cannot decompose by β−elimination pathway and hence such
complexes are generally more stable than the ones containing β−hydrogen atoms. For example, the neopentyl complex,
Ti[CH2C(CH3)3]4 (m.p 90 °C), and the benzyl complex, Zr(CH2Ph)4 (m.p. 132 °C), exhibit higher thermal stability as both of the
neopentyl and benzyl ligands lack β−hydrogens.
c. Central metal atom is coordinatively saturated
Transition metal organometallic complexes in which the central metal atom is coordinatively saturated tend to be more stable due
to the lack of coordination space available around the metal center to facilitate β−elimination reaction or other decomposition
reactions. Thus, the absence of free coordination sites at the metal is crucial towards enhancing the stability of the transition metal
organometallic complexes. For example, Ti(Me)4, which is coordinatively unsaturated can undergo a bimolecular decomposition
reaction via a binuclear intermediate (A), is unstable and exhibits a decomposition temperature of –40 °C. On the contrary,
Pb(Me)4, that cannot undergo decomposition by such bimolecular pathway, is more stable and distills at 110 °C at 1 bar
atmospheric pressure.

6.2.5 https://chem.libretexts.org/@go/page/172753
The Ti(Me)4 decomposes by dimerization involving the formation of Ti−C (3c−2e) bonds. For Pb(Me)4, such bimolecular
decomposition pathway is not feasible, as being a main group element it has higher outer d orbital for extending the coordination
number. If the free coordination site of Ti(Me)4 is blocked by another ligand, as in [(bipy)Ti(Me)4], then the thermal stability of the
complex, [(bipy)Ti(Me)4], increased significantly. Other bidentate chelating ligands like bis(dimethylphosphano)ethane (dmpe)
also serve the same purpose.

Coordinative saturation thus brings in kinetic stabilization in complexes. For example, Ti(Me)4is extremely reactive as it is
coordinatively unsaturated, while W(Me)4 is relatively inert for reasons of being sterically shielded and hence, coordinatively
saturated. Thus, if all of the above discussed criteria for the suppression of β-elimination are taken care of, then extremely stable
organometallic complexes can be obtained like the one shown below.

Problems
1. Arrange the following compounds in the order of their stability.
a. Ti(Et)4
b. Ti(Me)4 and
c. Ti(6-norbornyl)4
Ans: Ti(Et)4 < Ti(Me)4 < Ti(6-norbornyl)4
2. Predict the product of the reaction given below.
(BuP)CuCH CD C H ⟶
2 2 2 5

Ans: Equi molar amounts of (Bu3P)CuD and CH2=CDC2H5


3. Will the compound β-eliminate,
\Beta−elimination

PtH(C ≡ CH)L −−−−−−−−−−−→


2

(a). readily, (b). slowly and (c). not at all.


Explain your answer with proper reasoning.
Ans: Not at all as the ß-hydrogens are pointing away from the metal and cannot participate in ß-elimination recation.

6.2.6 https://chem.libretexts.org/@go/page/172753
Self Assessment test
1. Write the product(s) of the reactions.

Ans:

Summary
Ligands assume a pivotal role in the stabilization of the organometallic transition metal complexes. There are several methods
available for the preparation of the organometallic transition metal complexes. The observed instability of the organometallic
transition metal complexes can be attributed to two main phenomena namely β-elimination and bimolecular decomposition reaction
that severely undermine the instability of these complexes. The suppression of these decomposition pathway thus pave way for
obtaining highly stable organometallic transition metal complexes.

6.2.7 https://chem.libretexts.org/@go/page/172753
This page titled 6.2: Synthesis and Stability is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

6.2.8 https://chem.libretexts.org/@go/page/172753
CHAPTER OVERVIEW

7: Metal Alkyls and Metal Hydrides


7.1: Transition Metal Alkyl Complexes
7.2: Metal Hydrides

This page titled 7: Metal Alkyls and Metal Hydrides is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

1
7.1: Transition Metal Alkyl Complexes
Learning Objectives
In this lecture you will learn the following
Get a general prospective on the historical background of transition metal organometallic compounds with particular
emphasis on metal alkyls.
Know more about stable metal alkyls.
Get introduced to transition metal agostic alkyls.
Develop an understanding of reactions of relevance to metal alkyls like the reductive eliminations, oxidative additions and
the halide eliminations.

Transition metal σ−bonded organometallic compounds like the metal alkyls, aryls and the hydrides derivatives are by for the most
common organometallic species encountered in the world of chemistry. Yet, these compounds remained elusive till as late as the
1960s and the 1970s.

Historical background
Metal alkyls of the main group elements namely, Li, Mg, Zn, As and Al, have been known for a long time and which over the years
have conveniently found applications in organic synthesis whereas development on similar scale and scope in case of the transition
metal counterparts were missing till only recently. The origin of the organometallic compounds traces back to 1757, when Cadet
prepared a foul smelling compound called cacodyl oxide from As2O3 and CH3COOK, while working in a military pharmacy in
Paris. Years later in 1840, R. W. Bunsen gave the formulation of cacodyl oxide as Me2As−O−AsMe2. The next known transition
metal organometallic compound happens to be Et2Zn, which was prepared serendipitously in 1848 from the reaction of ethyl iodide
(EtI) and Zn with the objective of generating free ethyl radical. Frankland further synthesized alkyl mercury halides like, CH3HgI,
from the reaction of methyl iodide (CH3I) and Hg in sunlight. It is important to note that the dialkyl mercury, R2Hg, and the dialkyl
zinc, R2Zn, have found applications as alkyl transfer reagents in the synthesis of numerous main group organometallic compounds.
Another notable development of the time was of the preparation of Et4Pb from ethyl iodide (EtI) and Na/Pb alloy by C. J. Lowig
and M. E. Schweizer in 1852. They subsequently extended the same method for the preparation of the Et3Sb and Et3Bi compounds.
In 1859, aluminumalkyliodides, R2 AlI, were prepared by W. Hallwachs and A Schafarik from alkyl iodide (RI) and Al. The year
1863 saw the preparation of organochlorosilanes, RmSiCl4−m, by C. Friedel and J. M. Crafts while the year 1866 saw the synthesis
of halide-free alkyl magnesium compound, Et2Mg, by J. A. Wanklyn from the reaction of Et2Hg and Mg. In 1868, M. P.
Schutzenberger reported the first metal−carbonyl complex in the form of [Pt(CO)Cl2]2. In 1890, the first binary metal−carbonyl
compound, Ni(CO)4 was reported by L. Mond, who later founded the well−known chemical company called ICI (Imperial
Chemical Industries). In 1909, W. J. Pope reported the first σ−organotransition metal compound in the form of (CH3)3PtI. In 1917,
the alkyllithium, RLi, compounds were prepared by W. Schlenk by transalkylation reactions. In 1922, T. Midgley and T. A. Boyd
reported the utility of Et4Pb as an antiknock agent in gasoline. A. Job and A. Cassal prepared Cr(CO)6 in 1927. In 1930, K. Ziegler
showed the utility of organolithium compounds as alkylating agent while in the following year in 1931, W. Heiber prepared
Fe(CO)4H2 as the first transition metal−hydride complex. O. Roelen discovered the much renowned hydroformylation reaction in
1938, that went on to become a very successful industrial process worldwide.
The large scale production and the use of silicones were triggered by E. G. Rochow, when he reported the ‘direct synthesis’ from
methyl chloride (CH3Cl) and Si using Cu catalyst at 300 °C in 1943. The landmark compound, ferrocene (C5H5)2Fe, known as the
first sandwich complex was obtained by P. Pauson and S. A. Miller in 1951. H. Gilman introduced the important utility of
organocuprates when he prepared LiCu(CH3)2, in 1952. In the subsequent year 1953, G. Wittig found a new method of
synthesizing olefins from phosphonium ylides and carbonyl compounds that fetched him a Nobel prize in 1979. The year 1955
turned out to be a year of path breaking discoveries with E. O. Fischer reporting the rational synthesis of bis(benzene)chromium,
(C6H6)2Cr while K. Ziegler and G. Natta announcing the ground breaking polyolefin polymerization process that subsequently
gave them the Nobel prizes, E. O. Fischer sharing with G. Wilkinson in 1973 while K. Ziegler and G. Natta shared the same in
1963. In 1956, H. C. Brown reported hydroboration for which he too received the Nobel prize in 1979. In 1963, L. Vaska reported
the famous Vaska’s complex, trans−(PPh3)2Ir(CO)Cl, that reversibly binds to molecular oxygen. In 1964, E. O. Fischer reported
the first carbene complex, (CO)5WC(OMe)Me. In 1965, G. Wilkinson and R. S. Coffey reported the Wilkinson catalyst,
(PPh3)3RhCl, for the hydrogenation of alkenes. In 1973, E. O. Fischer synthesized the first carbyne complex, I(CO)4Cr(CR).

7.1.1 https://chem.libretexts.org/@go/page/172759
After the early 1970s, there were tremendous outburst in activity, in the area of transition metal organometallic chemistry leading to
phenomenal developments having far-reaching consequences in various branches of the main stream and interfacial chemistry.
Several Nobel prizes that have been awarded to the area in recent times fully recognized the significance of these efforts with Y.
Chauvin, R. R. Schrock, and R. H. Grubbs winning it in 2005 for olefin metathesis and Akira Suzuki, Richard F Heck and E.
Negishi receiving the same for the Pd catalyzed C−C cross-coupling reactions in organic synthesis in 2010.

Metal alkyls
In day to day organic synthesis, particularly from the application point of view, the metal alkyls are often perceived as a source of
stabilized carbanions for reactions with various electrophiles. The extent of stabilization of alkyl carbanions in metal alkyl
complexes depend upon the nature of the metal cations. For example, the alkyls of electropositive metals like that of Group 1 and
2, Al and Zn are regarded as polar organometallics as the alkyl carbanions remain weakly stabilized while retaining strong
nucleophilic and basic character of a free anion. These polar alkyls are extremely air and moisture sensitive as in their presence
they often get hydrolyzed and oxidized readily. Similar high reactivity was also observed in case of the early transition metal
organometallic compounds particularly of Ti and Zr. On the contrary the late transition metal organometallic compounds are much
less reactive and stable. For example, the Hg−C bond of (Me−Hg)+ cation is indefinitely stable in aqueous H2SO4 solution in air.
Thus, on moving from extremely ionic Na alkyls to highly polar covalent Li and Mg alkyls and to essentially covalent
late−transition metal alkyls, a steady decrease in reactivity is observed. This trend can be correlated to the stability of alkyl
carbanions that also depended on the nature of hybridization of the carbon center, with sp3 hybridized carbanions being the least
stable and hence most reactive, followed by the sp2 carbanions being moderately stable while the sp carbanions being the least
reactive and most stable. The trend also correlates well with the respective pKa values observed for CH4 (pKa = ~50), C6H6 (pKa =
~43) and RC≡CH (pKa = ~25).

Stable alkyls
As has been mentioned earlier, that the β−elimination is a crucial destabilizing influence on the transition metal organometallic
complexes. Hence, inhibition of this decomposition pathway leads to increased stabilities of organometallic compounds. Thus,
many stable alkyl transition metal complexes do not possesses β−hydrogens like, W(Me)6 and Ti(CH2Ph)4. In some cases despite
the presence of the β−hydrogens the organometallic complexes are stable as the β−hydrogens are deposed away from the metal
center like in, Cr(CHMe2)4, and Cr(CMe3)4. In this category of stable transition metal organometallic compounds also falls the
ones that contain β−hydrogens but cannot β−eliminate owing to the formation of a olefinic bond at a bridgehead, which is
unfavorable, like in Ti(6−norbornyl)4 and Cr(1−adamantyl)4. Lastly, some 18 VE metal complexes are stable, again despite having
β−hydrogens, for reasons of being electronically as well as coordinatively saturated at the metal center owing to attaining the stable
18 electron configuration.

Agostic alkyls
Agostic alkyls are extremely rare but very interesting species that represents a frozen point in a β−elimination pathway that have
fallen short of the completion of the decomposition reaction. Thus, these agostic alkyl complexes can be viewed as snap shots of a
β−elimination trajectory thereby providing valuable mechanistic understanding of the decomposition reaction. The agostic
interaction has characteristic signatures in various spectroscopic techniques as observed from the decreasing JC−Hcoupling constant
values in the 1H NMR and the 13C NMR spectra and the lowering of the νC−Hstretching frequencies in the IR spectroscopy. The
agostic alkyl complexes can be definitively proven by X−ray diffraction or neutron diffraction studies. The agostic alkyls thus have
activated C−H bonds which are of interest for their utility in chemical catalysis. Quite interestingly, many d0 Ti agostic alkyl
complexes do not β−eliminate primarily for the metal center being too electron deficient to donate electron to the σ* C−H orbital as
required for the subsequent β−elimination process.

Reductive elimination
Reductive elimination represents a major decomposition pathway of the metal alkyls. Opposite of oxidative addition, the reductive
elimination is accompanied by the decrease in the oxidation state and the valence electron count of the metal by two units. The
metal alkyl complexes may thus reductively eliminate with an adjacent hydrogen atom to yield an alkane, (R−H) or undergo the
same with an adjacent alkyl group to give an even larger alkane (R−R) as shown below.

Ln MRH ⟶ R−H + Ln M

Ln MR ⟶ R−R + Ln M
2

7.1.2 https://chem.libretexts.org/@go/page/172759
The reductive elimination is often facilitated by an electron deficient metal center and by sterically demanding ligand systems.
Often d8 metals like Ni(II), Pd(II), and Au(III) and d6 metals in high oxidation state like, Pt(IV), Pd(IV), Ir(III), and Rh(III) exhibit
reductive elimination.

Oxidative addition
Unlike the reductive elimination that represents a decomposition pathway of metal alkyls, the oxidative addition reaction represents
a useful method for the formation of the metal alkyl complexes. The oxidative addition thus leads to increase in valence electron
count and the oxidation state of the metal center by two units. The oxidative addition reactions are often facilitated by low valent
electron rich metal centers and by less sterically demanding ligands.

Halide elimination
β−halide elimination is observed for the early transition metals and the f−block elements resulting in the formation of stable alkyl
halides. The phenomenon is mostly seen in case of the metal fluorides and arise owing to the very high alkyl−fluoride bond
strengths that favor the halide elimination.

Problems
1. Who elucidated the structure of cacodyl oxide?
Ans: R. W. Bunsen in 1840
2. Give the example of the first olefin bound transition metal complex?
Ans: Zeise's salt, Na[PtCl3(C2H4)]
3. Who discovered olefin polymerization?
Ans: K. Ziegler and G. Natta
4. What kind of metal center promotes oxidative addition reactions?
Ans: Electron rich
5. The 18 VE complex would favor/disfavor oxidative addition reactions?
Ans: Disfavor

Self Assessment test


1. O. Roelen discovered which famous reaction?
Ans: Hydroformylation
2. What is the first binary metal−carbonyl complex?
Ans: Ni(CO)4
3. Who discovered the hydroboration reaction?
Ans: H. C. Brown
4. Reductive elimination reaction is favored by what kind of ligands?
Ans: Sterically demanding
5. β−halide elimination is mainly observed for what type of metal halide complexes?
Ans: Metal fluorides

7.1.3 https://chem.libretexts.org/@go/page/172759
Summary
A broader outlook on metal alkyls is obtained from the study of its historical background thus dispelling many myths about these
compounds like them being inherently unstable. It also establishes newer founding principles like these compounds indeed being
thermodynamically stable under certain experimental conditions and thus facilitating further attempts to take up the synthesis of
these compounds. Another important class of transition metal organometallic compounds are the agostic alkyls, which can be
viewed as the ones that have proceeded along but have fallen short of the final sequence of the β−elimination step. While oxidative
addition reaction remains a key method for synthesizing metal alkyls, the complementary reaction, i.e, the reductive elimination,
represents a decomposition reaction of these compounds. β−halide elimination reactions are observed for early transition metal
elements and f−block elements.

This page titled 7.1: Transition Metal Alkyl Complexes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

7.1.4 https://chem.libretexts.org/@go/page/172759
7.2: Metal Hydrides
Learning Objectives
In this lecture you will learn the following
Know about metal hydrides, their synthesis, characterization and their reactivity.
Know about σ−complexes and their properties.
Know about transition metal dihydrogen complexes.

Metal hydrides occupy an important place in transition metal organometallic chemistry as the M−H bonds can undergo insertion
reactions with a variety of unsaturated organic substrates yielding numerous organometallic compounds with M−C bonds. Not only
the metal hydrides are needed as synthetic reagents for preparing the transition metal organometallic compounds but they also are
required for important hydride insertion steps in many catalytic processes. The first transition metal hydride compound was
reported by W. Heiber in 1931 when he synthesized Fe(CO)4H2. Though he claimed that the Fe(CO)4H2 contained Fe−H bond, it
was not accepted until 1950s, when the concept of normal covalent M−H bond was widely recognized.
The metal hydride moieties are easily detectable in 1H NMR as they appear high field of TMS in the region between 0 to 60 ppm,
where no other resonances appear. The hydride moieties usually couple with metal centers possessing nuclear spins. Similarly, the
hydride moieties also couple with the adjacent metal bound phosphine ligands, if at all present in the complex, exhibiting
characteristic cis (J = 15 − 30 Hz) and trans (J = 90 − 150 Hz) coupling constants. In the IR spectroscopy, the M−H frequencies
appear between (1500 − 2200) cm−1 but their intensities are mostly weak. Crystallographic detection of metal hydride moiety is
difficult as hydrogen atoms in general are poor scatterer of X−rays. Located adjacent to a metal atom in a M−H bond, the detection
of hydrogen atom thus becomes challenging and as a consequence the X−ray crystallographic method systematically
underestimates the M−H internuclear distance by ~ 0.1 Å. However, better data could be obtained by performing the X−ray
diffraction studies at a low temperature in which the thermal motion of the atoms are significantly reduced. In light of these facts,
the neutron diffraction becomes a powerful method for detection of the metal hydride moieties as hydrogen scatters neutrons more
effectively and hence the M−H bond distances can be measured more accurately. A limitation of neutron diffraction method is that
large sized crystals are required for the study.

Synthesis
Following reactions are employed for synthesizing metal hydrides.
i. Protonation reactions
For this reaction to occur the metal center has to be basic and electron rich.
H+ H+
−2 −
[Fe(CO) ] −→
− [HFe(CO) ] −→
− H Fe(CO)
4 4 2 4

ii. From hydride donors


Generally for this method, a main group hydride is reacted with metal halide.

WCl + LiBEt H + PR ⟶ WH (PR )


6 3 3 6 3 3

iii. Using dihydrogen (H2) addition


This method involves oxidative addition of H2 and thus requires metal centers that are capable of undergoing the oxidative addition
step.
H2

WMe + PMe Ph −→ WH (PMe Ph)


6 2 6 2 3

iv. From a ligand


This method takes into account the β−elimination that occur in a variety of metal bound ligand moieties, thereby yielding a M−H
bond.

RuCl (PPh ) + KOCHMe + PPH ⟶ RuH (PPh ) + Me CO + KCl


2 3 3 2 3 2 3 4 2

7.2.1 https://chem.libretexts.org/@go/page/172760
Reactions of metal hydrides
Metal hydrides are reactive species kinetically and thus participate in a variety of transformations like the ones discussed below.
i. Deprotonation reactions
The deprotonation reaction can be achieved by a hydride moiety resulting in the formation of H2gas as shown below.

WH (PMe ) + NaH ⟶ Na[ WH (PMe ) ] + H


6 3 3 5 3 3 2

ii. Hydride transfer and insertion


In this reaction a hydride transfer from a metal center to formaldehyde resulting in the formation of a metal bound methoxy moiety
is observed as shown below.

Cp ⋅ ZrH + CH O ⟶ Cp ⋅ Zr(OMe)
2 2 2 2 2

iii. Hydrogen atom transfer reaction


An example of hydrogen atom transfer reaction is given below.
3 − 3 −
[Co (CN) H] + PhCH=CHCOOH ⟶ [Co (CN) ] + PHCH−CH COOH
5 5 2

It is interesting to note that the nature of hydrogen atom in a M−H bond can vary from being protic in nature, when bound to
electron deficient metal centers as in metal carbonyl compounds, to that of being hydridic in nature, when bound to more
electropositive early transition metals. In the latter case, the hydride moieties tend to be basic and exhibit hydride transfer reactions
with electrophiles like aldehydes or ketones. Furthermore, the protonation of these basic metal hydrides leads to the elimination of
dihydrogen (H2) gas along with the generation of a vacant coordination site at the metal center.

Bridging hydrides
The metal hydrides usually show two modes of binding, namely terminal and bridging. In case of the bridging hydrides, the
hydrogen atom can bridge between two or even more metal centers and thus, the bridging hydrides often display bent geometries.
σ−complexes
σ−complexes are rare compounds, in which the σ bonding electrons of a X−H bond further participate in bonding with a metal
center (X = H, Si, Sn, B, and P). The σ complexes thus exhibit an askewed binding to a metal center with the hydrogen atom,
containing no lone pair, being more close to the metal center and thereby resulting in a side−on structure. Many times if the metal
center is electron rich, then further back donation to the σ* orbital of the metal bound X−H moiety may occur resulting in a
complete cleavage of the X−H bond.

Metal dihydrogen complexes


The simplest variant of a σ−complex contains a dihydrogen ligand. The first dihydrogen complex was isolated by Kubas, after
which many new ones were reported.

7.2.2 https://chem.libretexts.org/@go/page/172760
Quite expectedly, the dihydrogen moiety bound to a metal in a σ−complex is found to be more acidic (pKa = 0 − 20) when
compared to the free dihydrogen molecule (pKa = 35). It is interesting to note that the pKa change associated with the binding of
dihydrogen to a metal in a σ−complex relative to that of the free H2 molecule is significantly larger than the change associated with
binding of H2O to metal. Owing to this inherent acidity, the deprotonation of the metal bound dihydrogen moiety by a base can thus
be appropriately employed for heterolytic activation of the dihydrogen moiety as illustrated below.

The dihydrogen complexes of metals are often referred to as nonclassical hydrides. The electron rich π basic metals are anticipated
to split the metal bound dihydrogen moieties resulting in classical dihydride complexes. Along the same line of thinking, the
electron deficient and less π basic metal would tend to stabilize a dihydrogen complex. The dihydrogen complexes can also be
characterized by the X ray diffraction as well as neutron diffraction methods. In IR spectrum, the metal bound H−H stretch appear
in the range (2300 − 2900) cm−1 while in the 1H NMR spectrum the same appear between 0 to −10 ppm as a broad peak. The
dihydrogen complexes are often characterized by isotopic labeling studies of metal bound H−D moiety that shows a coupling
constant of 20 – 34 Hz as supposed to 43 Hz observed in case of the free H−D molecule.

Problems
1. Predict the product of the reaction.

Ans:

7.2.3 https://chem.libretexts.org/@go/page/172760
2. Give the oxidation state and total valence electron count of the metal center.

Ans: Oxidation state 0 and 18 VE


3. What kind of metal centers would stabilize metal dihydrogen complexes?
Ans: Electron deficient and less π basic ligands
4. Specify whether the nature of hydrogen moiety in the complex, HCo(CO)4 is acidic or basic?
Ans: Acidic
5. Where do the M−H stretching bands appear in the IR spectrum of metal hydride complexes?
Ans: 1500 to 2200 cm-1

Self Assessment test


1. Predict the product of the reaction.
H2

IrCl(Co)(Pph ) −→
3 2

Ans:
H2

IrCl(CO)(PPh ) −
−→ IrH Cl(CO)(PPh )
3 2 2 3 2

2. Give the oxidation state and total valence electron count of the metal center.

Ans: Oxidation state +2 and 18 VE


3. What kind of metal centers would stabilize classical dihydride complexes?
Ans: Electron rich and more p basic ligands
4. Specify whether the nature of hydrogen moiety in the complex, IrH5(PCy3)2 is acidic or basic?
Ans: Basic
5. Between X−ray diffraction and neutron diffraction, which is a better method for the characterization of the M−H moiety?
Ans: Neutron diffraction

Summary
Metal hydrides are important compounds in the overall scheme of organometallic chemistry as they are involved in many crucial
steps of numerous catalytic reactions. Apart from metal hydrides another important class of compounds are transition metal
σ−complexes whose simplest variant are the metal dihydrogen complexes. These σ−complexes and the metal dihydrogen
complexes are important for the heterolytic activations of the respective metal bound H−heteroatom and the H−H bonds.

7.2.4 https://chem.libretexts.org/@go/page/172760
This page titled 7.2: Metal Hydrides is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S. Balakrishna
& Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

7.2.5 https://chem.libretexts.org/@go/page/172760
CHAPTER OVERVIEW

8: Carbonyls and Phosphine Complexes


8.1: Metal Carbonyls
8.2: Metal Phosphines

This page titled 8: Carbonyls and Phosphine Complexes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

1
8.1: Metal Carbonyls
Learning Objectives
In this lecture you will learn the following
The historical background of metal carbonyl complexes.
The CO ligand and its binding ability to metal.
Synergism between the ligand to metal forward σ–donation and the metal to ligand backward π–donation observed in a
metal-CO interaction.
The synthesis, characterization and their reactivity of the metal carbonyl compounds.

Metal carbonyls are important class of organometallic compounds that have been studied for a long time. Way back in 1884,
Ludwig Mond, upon observing that the nickel valves were being eating away by CO gas in a nickel refining industry, heated nickel
powder in a stream of CO gas to synthesize the first known metal carbonyl compound in the form Ni(CO)4. The famous Mond
refining process was thus born, grounded on the premise that the volatile Ni(CO)4 compound can be decomposed to pure metal at
elevated temperature. Mond subsequently founded the Mond Nickel Company Limited for purifying nickel from its ore using this
method.
The carbonyl ligand (CO) distinguishes itself from other ligands in many respects. For example, unlike the alkyl ligands, the
carbonyl (CO) ligand is unsaturated thus allowing not only the ligand to σ−donate but also to accept electrons in its π* orbital from
dπ metal orbitals and thereby making the CO ligand π−acidic. The other difference lies in the fact that CO is a soft ligand compared
to the other common σ−and π−basic ligands like H2O or the alkoxides (RO−), which are considered as hard ligands.
Being π−acidic in nature, CO is a strong field ligand that achieves greater d−orbital splitting through the metal to ligand π−back
donation. A metal−CO bonding interaction thus comprises of a CO to metal σ−donation and a metal to CO π−back donation
(Figure 8.1.1). Interestingly enough, both the spectroscopic measurements and the theoretical studies suggest that the extent of the
metal to CO π−back donation is almost equal to or even greater than the extent of the CO to metal σ−donation in metal carbonyl
complexes. This observation is in agreement with the fact that low valent−transition metal centers tend to form metal carbonyl
complexes.

Figure 8.1.1 : Orbital diagram showing ligand to metal forward σ-donation and the metal to ligand backward π-donation in metal-
CO interaction.
In the metal carbonyl complexes, the direct bearing of the π−back donation is observed on the M−C bond distance that becomes
shorter as compared to that of a normal M−C single bond distance. For example, the CpMo(CO)3CH3 complex, exhibits two kind
of M−C bond distances that comprise of a longer Mo−CH3 distance (2.38 Å) and a much shorter Mo−CO distance (1.99 Å) arising
out of a metal to ligand π−back donation. It becomes thus apparent that the metal−CO interaction can be easily characterized using
X−ray crystallography. The infrared spectroscopy can also be equally successfully employed in studying the metal−CO interaction.
Since the metal to CO π−back bonding involves a π−donation from the metal dπ orbital to a π* orbital of a C−O bond, significant
shift of the ν(CO) stretching frequency towards the lower energy is observed in metal carbonyl complexes with respect to that of
free CO (2143 cm−1).

8.1.1 https://chem.libretexts.org/@go/page/172766
Preparation of metal carbonyl complexes
The common methods of the preparation of the metal carbonyl compounds are,
i. Directly using CO
C O, 200atm, 200°C

Fe −−−−−−−−−−−→ Fe(CO)
5

The main requirement of this method is that the metal center must be in a reduced low oxidation state in order to facilitate CO
binding to the metal center through metal to ligand π−back donation.
ii. Using CO and a reducing agent
2 −
NiSO + CO + S O4 ⟶ Ni (CO)
4 2 4

This method is commonly called reductive carbonylation and is mainly used for the compounds having higher oxidation state metal
centers. The reducing agent first reduces the metal center to a lower oxidation state prior to the binding of CO to form the metal
carbonyl compounds.
iii. From carbonyl compounds
This method involves abstraction of CO from organic compounds like the alcohols, aldehydes and CO2.

Reactivities of metal carbonyls


i. Nucleophilic attack on carbon

The reaction usually gives rise to carbene moiety.


ii. Electrophilic attack at oxygen

Cl (PR ) Re−CO + AlMe ⟶ Cl (PR ) Re−CO ⟶ AlMe


3 4 3 3 4 3

iii. Migratory insertion reaction

MeMn(CO) + PMe ⟶ (MeCO)Mn(CO) (PMe )


5 3 4 3

The metal carbonyl displays two kinds of bindings in the form of the terminal and the bridging modes. The infrared spectroscopy
can easily distinguish between these two binding modes of the metal carbonyl moiety as the terminal ones show ν(CO) stretching
band at ca. 2100-2000 cm−1 while the bridging ones appear in the range 1720−1850 cm−1. The carbonyl moiety can bridge between
more than two metal centers (Figure 8.1.2).

8.1.2 https://chem.libretexts.org/@go/page/172766
Figure 8.1.2 : Different bridging modes of the carbonyl binding to a metal is shown.

Problems
1. How many lone pairs are there in the CO molecule?
Ans: Three (one from carbon and two from oxygen).
2. Despite O being more electronegative than C, the dipole moment of CO is almost zero. Explain.
Ans: Because of the electron donation from oxygen to carbon.
3. What type of metal centers form metal carbonyl complexes?
Ans: Low−valent metal centers.
4. What are the two main modes of binding exhibited by CO ligand?
Ans: Terminal and bridging modes of binding.

Self Assessment test


1. Predict the product of the reaction?

Ni + CO ⟶
Excess

Ans: Three (one from carbon and two from oxygen).


2. Upon binding to a metal center the C−O stretching frequency increases/decreases with regard to that of the free CO?
Ans: Decreases.
3. Explain why do low−valent metal centers stabilize CO binding in metal carbonyl complexes?
Ans: Because metal to ligand π−back donation.
4. Give an example of a good σ−donor and π−donor ligand?
Ans: Alkoxides (RO-).

Summary
CO is a hallmark ligand of organometallic chemistry. The metal carbonyl complexes have been studied for a long time. The CO
ligands bind tightly to metal center using a synergistic mechanism that involves σ−donation of the ligand lone pair to metal and
followed by the π−back donation from a filled metal d orbital to a vacant σ* orbital of C−O bond of the CO ligand. The metal
carbonyl complexes are prepared by several methods. The metal carbonyl complexes are usually stabilized by metal centers in low
oxidation states.

This page titled 8.1: Metal Carbonyls is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

8.1.3 https://chem.libretexts.org/@go/page/172766
8.2: Metal Phosphines
Learning Objectives
In this lecture you will learn the following
Know about metal phosphine complexes.
Have an understanding of the steric and electronic properties of the phosphine ligands.
Obtain a deeper insight about the metal phosphine interactions.
Be introduced to other π−basic ligands.

Phosphines are one of the few ligands that have been extensively studied over the last few decades to an extent that the systematic
fine tuning of the sterics and electronics can now be achieved with certain degree of predictability. Phosphines are better spectator
ligands than actor ligands. Tolman carried out pioneering infrared spectroscopy experiments on the PR3Ni(CO)3 complexes looking
at the ν(CO)stretching frequencies for obtaining an insight on the donor properties of the PR3 ligands. Thus, a stronger σ−donor
phosphine ligand would increase the electron density at the metal center leading to an enhanced metal to ligand π−back bonding
and thereby lowering of the ν(CO) stretching frequencies in these complexes. Another important aspect of the phosphine ligand is its
size that has significant steric impact on its metal complexes. Thus, unlike CO ligand, which is small and hence many may
simultaneously be able to bind to a metal center, the same is not true for the phosphine ligands as only a few can bind to a metal
center. The number of phosphine ligands that can bind to a metal center also depends on the size of its R substituents. For example,
up to two can bind to a metal center in case of the PCy3 or P(i−Pr)3 ligands, three or four for PPh3, four for Me2PH, and five or six
for PMe3. The steric effect of phosphine was quantified by Tolmann and is given by a parameter called Cone Angle that measures
the angle at the metal formed by the PR3 ligand binding to a metal (Figure 8.2.1).

Figure 8.2.1 : Cone Angle in metal−phosphine complexes.


The Cone Angle criteria has been successfully invoked in rationalizing the properties of a wide range of metal phosphine
complexes. One unique feature of the phosphine ligand is that it allows convenient change of electronic effect without undergoing
much change in its steric effects. For example, PBu3 and P(OiPr)3 have similar steric effects but vary in their electronic effects. The
converse is also true as the steric effect can be easily changed without undergoing much change in the electronic effect. For
example, PMe3 and P(o−tolyl)3 have similar electronic effect but differ in their steric effects. Thus, the ability to conveniently
modulate the steric and the electronic effects make the phosphine ligands a versatile system for carrying out many organometallic
catalysis.

Structure and Bonding


Phosphines are two electron donors that engage a lone pair for binding to metals. These are thus considered as good σ−donors and
poor π−acceptors and they belong to the same class with the aryl, dialkylamino and alkoxo ligands. In fact they are more π−acidic
than pure σ−donor ligands like NH3and, more interestingly so, their π−acidity can be varied significantly by systematic

8.2.1 https://chem.libretexts.org/@go/page/172767
incorporation of substituents on the P atom. For example, PF3 is more π−acidic than CO. Analogous to what is observed in case of
the benchmark π−acidic CO ligand, in which the metal dπ orbital donates electron to a π* orbital of a C−O bond, in the case of the
phosphines ligands, such π−back donation occurs from the metal dπ orbital occurs on to a σ* orbital of a P−R bond (Figure 8.2.2).
In phosphine ligands, with the increase of the electronegativity of R both of the σ and the σ* orbitals of the P−R bond gets
stabilized. Consequently, the contribution of the atomic orbital of the P atom to the σ*−orbital of the P−R bond increases, which
eventually increases the size of the σ* orbital of the P−R bond. This in turn facilitates better overlap of the σ* orbital of the P−R
bond with the metal dπ orbital during the metal to ligand π−back donation in these metal phosphine complexes.

Figure 8.2.2 : Back donation from the metal dπ orbital to a σ* orbital of a P−R bond.
Starting from CO, which is a strong π−acceptor ligand, to moving to the phosphines, which are good σ−donors and poor
π−acceptor ligands, to even going further to other extreme to the ligands, which are both good σ−donors as well as π−donors, a rich
variety of phosphine ligands thus are available for stabilizing different types of organometallic complexes. In this context the
following ligands are discussed below.

π-basic ligands
Alkoxides (RO−) and halides like F−, Cl− and Br− belong to a category of π−basic ligands as they engage a second lone pair for
π−donation to the metal over and above the first lone pair partaking σ−donation to the metal. Opposite to what is observed in the
case of π−acidic ligands, in which the π* ligand orbital stabilizes the dπ metal orbital and thereby affecting a larger ligand field
splitting, as consistent with the strong field nature of these ligands (Figure 3), in the case of the π−basic ligands, the second lone
pair destabilizes the dπ metal orbitals leading to a smaller ligand field splitting, which is in agreement with the weak field nature of
these ligands. The orbitals containing the lone pair of the ligands are usually located on the more electronegative heteroatoms and
so they are invariably lower in energy than the metal dπ orbitals. Hence, the destabilization of the metal dπ orbitals occurs due to the
repulsion of the filled ligand lone pair orbital with the filled metal dπ orbitals. In case of the situations in which the metal dπ orbitals
are vacant, like in d0 systems of Ti4+ ions, the possibility of the destabilization of the metal dπ orbitals do not arise but instead
stabilization occurs through the donation of the filled ligand lone pair orbital electrons to the empty metal dπ orbitals as seen in the
case of TiF6 and W(OMe)6. Thus, this scenario in π−basic ligands is opposite to that observed in case of the π−acidic ligands, for
which the empty π* ligand orbitals are higher in energy than the filled metal dπ orbitals.

8.2.2 https://chem.libretexts.org/@go/page/172767
Figure 8.2.3 : Orbital interactions in the presence of the π−acceptor, (pure) σ−donor and π−basic ligands are shown.

This page titled 8.2: Metal Phosphines is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

8.2.3 https://chem.libretexts.org/@go/page/172767
CHAPTER OVERVIEW

9: Complexes of π−bound Ligands


9.1: Metal Alkene Complexes
9.2: Metal Allyl and Diene Complexes
9.3: Metal Cyclopentadienyl Complexes

This page titled 9: Complexes of π−bound Ligands is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M.
S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

1
9.1: Metal Alkene Complexes
Learning Objectives
In this lecture you will learn the following
The metal alkene complexes.
The metal−olefin bonding interactions.
The synthesis and reactivities of the metal−olefin complexes.
The umpolung reactivities of olefins in the metal alkene complexes.

Though the first metal olefin complex dates back a long time to the beginning of 19th century, its formulation was established only
a century later in the 1950s. While reacting K PtCl with EtOH in 1827, the Danish chemist Zeise synthesized the famous Zeise’s
2 4

salt K[PtCl (C H )] ∙ H O containing a Pt bound ethylene moiety and which incidentally represented the first metal−olefin
3 2 4 2

complex (Figure 9.1.1).

Figure 9.1.1 : Zeise’s salt


The metal−olefin bonding interaction is best explained by the Dewar−Chatt model, that takes into account two mutually opposing
electron donation involving σ−donation of the olefinic C=C π−electrons to an empty dπ metal orbital followed by π−back donation
from a filled metal dπ orbital into the unoccupied C=C π* orbital. Quite understandably so, for the d0 systems, the formations of
metal−olefin complexes are not observed. The extent of the C=C forward π-donation to the metal and the subsequent π−back
donation from the filled dπ orbital to the olefinic C=C π* orbital have a direct bearing on the C=C bond of the metal bound olefinic
moiety in form of bringing about a change in hybridization as well as in the C−C bond distance (Figure 9.1.2).

Figure 9.1.2 : Metal olefin bonding interactions.


If the metal to ligand π−back donation component is smaller than the ligand to metal σ−donation, then the lengthening of the C−C
bond in the metal bound olefin moiety is observed. This happens primarily because of the fact that the alkene to metal σ−donation
removes the C=C π−electrons away from the C−C bond of the olefin moiety and towards the metal center, thus, decreasing its bond
order and increasing the C−C bond length. Additionally, as the metal to ligand π−back donation increases, the electron donation of
the filled metal dπ orbital on to the π* orbital of the metal bound olefin moiety is enhanced. This results in an increase in the C−C
bond length. The lengthening of the C−C bond in metal bound olefin complex can be correlated to the π−basicity of the metal. For

9.1.1 https://chem.libretexts.org/@go/page/172773
example, for a weak π−basic metal, the C−C bond lengthening is anticipated to be small while for a strong π−basic metal, the C−C
lengthening would be significant.
Another implication of ligand−metal π−back donation is in the observed change of hybridization at the olefinic C atoms from pure
sp2, in complexes with no metal to ligand π−back donation, to sp3, in complexes with significant metal to ligand π−back donation,
is observed. The change in hybridization from sp2 to sp3 centers of the olefinic carbon is accompanied by the substituents being
slightly bent away from the metal center in the final metalacyclopropane form (Figure 9.1.3). This change in hybridization can be
conveniently detected by 1H and 13C NMR spectroscopy. For example, in case of the metalacyclopropane systems, which have
strong metal to ligand π−back donation, the vinyl protons appear 5 ppm (in the 1H NMR) and 100 ppm (in the 13C NMR) high field
with respect to the respective position of the free ligands.
An interesting fallout of the metal to ligand π−back bonding is the tighter binding of the strained olefins to the metal center as
observed in the case of cyclopropene and norbornene. The strong binding of these cyclopropene and norbornene moieties to the
metal center arise out of the relief of ring strain upon binding to the metal. Lastly, in the metal−olefin complexes having very little
π−back bonding component, the chemical reactivities of the metal bound olefin appear opposite to that of a free olefin. For
example, a free olefin is considered electron rich by virtue of the presence of π−electrons in its outermost valence orbital and hence
it undergoes an electrophilic attack. However, the metal bound olefin complexes having predominantly σ−donation of the olefinic
π−electrons and negligible metal to ligand π−back donation, the olefinic C becomes positively charged and hence undergoes a
nuclophilic attack. This nature of reversal of olefin reactivity is called umpolung character.

Figure 9.1.3 : Metalacyclopropane system and Dewar−Chatt model

Synthesis
Metal alkene complexes are synthesized by the following methods.
i. Substitution in low valent metals

AgOSO CF +C H ⟶ (C H )AgOSO CF
2 3 2 4 2 4 2 3

ii. Reduction of high valent metal in presence of an alkene


− −
(cod)PtCl +C H ⟶ [ PtCl (C H )] + Cl
2 2 4 3 2 4

iii. From alkyls and related species

Cp TaCl + n−BuMgX ⟶ Cp TaBu


2 3 2 3

\Beta−elimination

Cp TaBu −−−−−−−−−−−−→ Cp TaH(1 −Butene) + Butene + Butane


2 3 2
reductive elimination

Reaction of alkenes
The metal alkene complexes show the following reactivities.

i. Insertion reaction
These reactions are commonly displayed by alkenes as they insert into metal−X bonds yielding metal alkyls. The reaction occurs
readily at room temperature for X = H, whereas for other elements (X = other atoms), such insertions become rare. Also, the
strained alkenes and alkynes undergo such insertion readily.

PtHCl (PEt ) +C H −
↽⇀
− PtElCl (PEt )
3 2 2 4 3 2

9.1.2 https://chem.libretexts.org/@go/page/172773
ii. Umpolung reactions
Umpolung reactions are observed only for those metal−alkene complexes for which the metal center is a poor π−base and as a
result of which the olefin undergoes a nuclophilic attack.

iii. Oxidative addition


Alkenes containing allylic hydrogens undergo oxidative addition to give a allyl hydride complex.

 Exercise 9.1.1

Predict the product of the reaction.

AuMe(PPh ) + CF =CF ⟶ A ⟶ B
3 2 2

Answer
A = {(CF2=CF2)AuMe(PPh3)} and
B = Au(CF2-CF2Me)(PPh3)

 Exercise 9.1.2

Specify whether the lengthening/shortening of the C−C bond distance in the metal bound olefin moiety is observed as a result
of metal to ligand π−back donation?

Answer
Lengthening

 Exercise 9.1.3

Draw the structure of Zeise’s salt.

Answer

 Exercise 9.1.4

The change in hybridization at the olefinic C from sp2 to sp3 primarily arise due to?

Answer
Metal-ligand π-back donation.

Self Assessment test


1. Predict the product of the reaction.

PtCl +C H ⟶
2 2 4

9.1.3 https://chem.libretexts.org/@go/page/172773
Ans: [PtCl3(C2H4)]- and Cl-
2. Specify whether the lengthening/shortening of the C−C bond distance in the metal bound olefin moiety is observed as a result of
ligand to metal σ− donation?
Ans: Lengthening.
3. Metalacyclopropane intermediate in a metal bound olefin complex is primarily formed due to which kind of interaction?
Ans: Metal−ligand π−back donation
4. The oxidation state of Pt in Zeise’s salt is?
Ans: PtII

Summary
Alkenes are an important class of unsaturated ligands that bind to a metal by σ−donating its C=C π−electrons and also accepts
electrons from the metal in its π* orbital of C=C bond. These symbiotic σ−donation and π−back donation in metal bound olefin
complexes have a significant impact on their structure and reactivity properties. Quite importantly, the structural manifestations
arising out of these forward σ−donation and π−back donation can be characterized by using 1H, 13C NMR and IR spectroscopic
methods.

This page titled 9.1: Metal Alkene Complexes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

9.1.4 https://chem.libretexts.org/@go/page/172773
9.2: Metal Allyl and Diene Complexes
Learning Objectives
In this lecture you will learn the following
The metal−allyl complexes.
The metal−diene complexes.
The metal−cyclobutadiene complexes.
The respective metal−ligand interactions.

The allyl ligand is often referred to as an “actor” ligand rather than a “spectator” ligand. It binds to metals in two ways i.e. in a η1
(monohapto) form and a η3 (trihapto) form (Figure 9.2.1). (i). In its monohapto (η1) form, it behaves as an anionic 1e−donor X
type of a ligand analogous to that of a methyl moiety while (ii) in a trihapto (η3) form, it acts as an anionic 3e−donor LX type of a
ligand.

Figure 9.2.1 : Metal−allyl interaction.

Metal−allyl interaction
Of particular interest are the molecular orbitals namely Ψ1, Ψ2 and Ψ3 of the allyl ligand that interact with the metal in a metal allyl
complex. The energy of these molecular orbitals increase with the increase in the number of nodes. Of the three, the Ψ1 and Ψ2
orbitals usually engage in ligand to metal σ−donation, with Ψ1 involving in a dative L−type bonding and Ψ2 participating in a
covalent X−type bonding with the metal d orbitals (Figure 9.2.2).

9.2.1 https://chem.libretexts.org/@go/page/172774
Figure 9.2.2 : Metal−allyl interactions.

Synthesis of the metal allyl complexes


The metal allyl complexes are synthesized by the following methods.
i. From an alkene complex as shown below.

Mo (dpe) (η2 −propene) −


↽⇀
− Mo (dpe) (η3 −allyl)H
2 2

ii. By a nucleophilic attack of an allyl compound as shown below.

iii. By an electrophilic attack of an allyl compound as shown below.

9.2.2 https://chem.libretexts.org/@go/page/172774
iv. From a diene complex as shown below.

Reactions of metal allyl complexes


The reactivities of the metal allyl complexes toward various species are illustrated below.
i. Reaction with nucleophiles

ii. Reaction with electrophiles

iii. Insertion reaction

9.2.3 https://chem.libretexts.org/@go/page/172774
iv. Reductive elimination

Diene complexes
1,3−Butadiene is a 4e−donor ligand that binds to a metal in a cisoid conformation. The Dewar−Chatt model, when applied to
1,3−butadiene, predicts that the ligand may bind to metal either as a L2 (π2) donor type, similar to that of an alkene, or as an LX2
(σ2π) donor type, similar to that of a metalacyclopropane form. The L2 binding of 1,3−butadiene is rare, e.g. asin
(butadiene)Fe(CO)3, while the LX2 type binding is more common, e.g. as in Hf(PMe3)2Cl2. An implication of the LX2 type binding
is in the observed shortening of the C2−C3 (1.40 Å) distance alongside the lengthening of the C1−C2(1.46 Å) and C3−C4 (1.46 Å)
distances (Figure 9.2.3).

Figure 9.2.3 : Metal−diene interaction in cisoid binding.


The molecular orbitals of the 1,3−butadiene ligand comprises of two filled Ψ1 (HOMO−1) and Ψ2(HOMO) orbitals and two empty
Ψ3 (LUMO) and Ψ4 (LUMO+1) orbitals. In a metal−butadiene interaction the ligand to metal σ−donation occurs from the filled Ψ2

9.2.4 https://chem.libretexts.org/@go/page/172774
orbital of the 1,3−butadiene ligand while the metal to ligand π−back donation occurs on to the empty Ψ3 orbital of the
1,3−butadiene ligand (Figure 9.2.4).

Figure 9.2.4 : Metal−diene interaction.


Though cisoid binding is often observed in metal butadiene complexes, a few instances of transoidbinding is seen in dinuclear, e.g.
as in Os3(CO)10(C4H6), and in mononuclear complexes e.g. as in Cp2Zr(C4H6) (Figure 9.2.5).

Figure 9.2.5 : Metal−diene interaction in transoid binding.

Synthesis of metal butadiene complex


Metal butadiene complexes are usually prepared by the same methods used for synthesizing metal alkene complexes. Two
noteworthy synthetic routes are shown below.

9.2.5 https://chem.libretexts.org/@go/page/172774
Metal cyclobutadiene complexes
Cyclobutadiene is an interesting ligand because of the fact that its neutral form, being anti−aromatic (4π−electrons), is unstable as a
free molecule (Figure 9.2.6), but its dianionic form is stable because of being aromatic (6π−electrons). Consequently, the
cyclobutadiene ligand is stabilized by significant metal to ligand π−back donation to the vacant ligand orbitals.

Figure 9.2.6 : Electronic structure of cyclobutadiene ligand.


A synthetic route to metal cyclobutadiene complex is shown below.

9.2.6 https://chem.libretexts.org/@go/page/172774
Problems
1. The hapticities displayed by an allyl moiety in binding to metals are?
Ans: 1 and 3.
2. Identify which molecular orbitals of an allyl moiety engage in σ−interaction with a suitable d orbital of a metal in a η3−metal
allyl complex?
Ans: Ψ1 and Ψ2.
3. Predict the product of the reaction.

Ans:

4. Identify which molecular orbitals of a butadiene moiety engage in σ−interaction with a suitable dorbital of a metal in a η4−metal
butadiene complex?
Ans: Ψ2.

Self Assessment test


1. Predict the product of the reaction.

< (−−Ni−−) > +CO ⟶


2

Ans:

2. Identify which molecular orbitals of a butadiene moiety engage in π−interaction with a suitable dorbital of a metal in a η4−metal
allyl complex?
Ans: Ψ3.
3. Mention the type of orientations displayed by butadiene ligands for binding to metal.

9.2.7 https://chem.libretexts.org/@go/page/172774
Ans: Cisoid (common) and transoid (rare).
4. Comment on the number of π−electrons present in the cyclobutadiene moiety of a metal cyclobutadiene complex.
Ans: 6 π−electrons.

Summary
Allyl, 1,3−butadiene and cyclobutadiene together constitute an important class of σ−donor/π−acceptor ligands that occupy a special
place in organometallic chemistry. The complexes of these ligands with metals are important intermediates in many catalytic cycles
and hence an understanding of their interaction with metal is of significant importance. In this context, the synthesis,
characterization and the reactivities of the organometallic complexes of these ligands are described alongside the respective
metal−ligand interactions.

This page titled 9.2: Metal Allyl and Diene Complexes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

9.2.8 https://chem.libretexts.org/@go/page/172774
9.3: Metal Cyclopentadienyl Complexes
Learning Objectives
In this lecture you will learn the following
The cyclopentadienyl ligands.
The synthesis and reactivity of metal−cyclopentadienyl complexes.
The metal−cyclopentadienyl interaction.

Cyclopentadienyl moiety acts as an important “spectator” ligand and is quite ubiquitous in organometallic chemistry. It remains
inert to most nucleophiles and electrophiles and solely engages in stabilizing organometallic complexes. The cyclopentadienyl
ligands form a wide array of organometallic compounds exhibiting different formulations that begin with the so-called “piano
stool” CpMLn (n = 2,3 or 4) type ones and extends to the most commonly observed “metallocene” Cp2M type ones to even go
beyond further to the “bent metallocene” Cp2MXn (n = 1,2 or 3) type ones. In the “piano stool” CpMLn structure, the
cyclopentadienyl (Cp) ligand is regarded as the “seat” of the piano stool while the remaining L ligands are referred to as the “legs”
of the piano stool. Though the cyclopentadienyl ligand often binds to metal in a η5 (pentahapto) fashion, e. g. as in ferrocene, the
other form of binding to metal at lower hapticities, like that of the η3 (trihapto) binding e. g. as in (η5−Cp)(η3−Cp)W(CO)2 and that
of the η1 (monohapto) binding e. g. as in (η5−Cp)(η1−Cp)Fe(CO)2, are also seen on certain rare occasions.
The binding modes of the cyclopentadienyl ligand in metal complexes can be ascertained to a certain degree by 1H NMR in the
diamagnetic metal complexes, in which the Cp−protons appear as a singlet between 5.5−3.5 ppm while the β and γ hydrogens come
at 7−5 ppm.

Cyclopentadienyl−metal interaction
The frontier molecular orbital of the cyclopentadienyl ligand contains 5 orbitals (Ψ1−Ψ5) residing in three energy levels (Figure
9.3.1 ). The lowest energy orbital Ψ1 does not contain any node and is represented by an a1 state, followed by a doubly degenerate
e1 states that comprise of the Ψ2 and Ψ3orbitals, which precede another doubly degenerate e2 states consisting of Ψ4 and Ψ5
orbitals.

Figure 9.3.1 : Molecular orbital diagram of cyclopentadienyl ligand.


The above frontier molecular orbital diagram becomes more intriguing on moving over to the metallocenes that contain two such
cyclopentadienyl ligands. Specifically, in the Cp2M system, (e. g. ferrocene) each of these above five molecular orbital of the two
cyclopentadienyl ligands combines to give ten ligand molecular orbitals in three energy levels (Figure 9.3.2 ). Of these, the orbitals

9.3.1 https://chem.libretexts.org/@go/page/172775
that subsequently interact with the metal orbitals to generate the overall molecular orbital correlation diagram for the Cp2M type of
complexes are shown below (Figure 9.3.3 ).

Figure 9.3.2 : Metal−cyclopentadienyl Bonding interactions.

9.3.2 https://chem.libretexts.org/@go/page/172775
Figure 9.3.3 : MO diagram of ferrocene.
Generic metallocene Cp2M type complexes are formed for many from across the 1st row transition metal series along Sc to Zn. The
number of unpaired electrons thus correlates with the number unpaired electrons present in the valence orbital of the metal (Figure
4). Of the complexes of the 1st row transition metal series, the manganocene exists in two distinct forms, one in a high-spin form
with five unpaired electrons, e.g. as in Cp2Mn and the other in a low-spin form with one unpaired electron, e.g. as in Cp*2Mn
owing to the higher ligand field strength of the Cp* ligand. Cobaltocene, Cp2Co, has 19 valence electrons (VE) and thus gets easily
oxidized to the diamagnetic 18 VE valence electron species, Cp2Co+. Of these metallocenes, the much-renowned ferrocene, Cp2Fe
is a diamagnetic 18 VE complex, whose molecular orbital diagram is shown above (Figure 9.3.3 ).

9.3.3 https://chem.libretexts.org/@go/page/172775
Figure 9.3.4 : Metallocene

Bent metallocenes
Bent metallocenes are Cp2MXn type complexes formed of group 4 and the heavier elements of groups 5−7. In these complexes the
frontier doubly degenerate e2g orbitals of Cp2M fragment interacts with the filled lone pair orbitals of the ligand (Figure 9.3.5 ).

Figure 9.3.5 : Bent metallocene.

9.3.4 https://chem.libretexts.org/@go/page/172775
Synthesis of cyclopentadienyl-metal complexes
The metal−cyclopentadienyl complexes are synthesized by the following methods.
i. from Cp−

ii. from Cp+

iii. from hydrocarbon

Reactivity of cyclopentadienyl-metal complexes


The reactivity of cyclopentadienyl−metal complexes of the type Cp2M is shown for a representative nickellocene complex.
i. reaction with NO

9.3.5 https://chem.libretexts.org/@go/page/172775
ii. reaction with PR3

iii. reaction with CO

iv. reaction with H+

Problems
1. Comment on the p−acceptor property of the cyclopentadienyl ligand.
Ans: The ligand being anionic shows very little π-acceptor properties.
2. Give the total valence electron count at the metal in a nickellocene complex.
Ans: 20 electrons.
3. Explain why the metal center in cobalticene gets easily oxidized.
Ans: 19 electrons cobalticene gets easily oxidized to 18 electron Cp2Co+.
4. Specify the number unpaired electrons present in chromocene.
Ans: 2

9.3.6 https://chem.libretexts.org/@go/page/172775
Self Assessment test
1. Specify the number of unpaired electron present in vanadocene.
Ans: The ligand being anionic shows very little π-acceptor properties.
2. What different hapticities are exhibited by cyclopentadienyl ligand?
Ans: 1, 3, and 5.
3. Specify the hapticities of the cyclopentadienyl ligands in Cp2W(CO)2.
Ans: 5 and 3.
4. Specify the hapticity of the cyclopentadienyl ligands in CpRh(CO)2(PMe3).
Ans: 3.

Summary
Cyclopentadienyl moiety is almost synonymous with the transition metal organometallic complexes as the ligand played a pivotal
role at the early developmental stages of the field of organometallic chemistry in the 1960s and 1970s. An important quality of the
cyclopentadienyl ligand is that it behaves as an extremely good “spectator” ligand being inert to nucleophiles and electrophiles and
displays uncanny ability towards stabilizing metal complexes of elements from across the different parts of the periodic table.
Cyclopentadienyl moiety thus forms several types of complexes of different formulations like that of the “piano stool” CpMLn (n =
2,3 or 4) types, the metallocene Cp2M types and the bent metallocene Cp2MXn (n = 1,2 or 3) types. Cyclopentadienyl metal
complexes make valuable catalysts for many chemical transformations of interest to academia and industries alike. The
cyclopentadienyl moiety participates in a complex interaction with the metal involving ligand frontier molecular orbitals and the
metal valence orbitals. Cyclopentadienyl metal complexes can be accessed by many methods.

This page titled 9.3: Metal Cyclopentadienyl Complexes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

9.3.7 https://chem.libretexts.org/@go/page/172775
CHAPTER OVERVIEW

10: Reaction Mechanisms


10.1: Oxidative Addition and Reductive Elimination
10.2: Insertion and Elimination Reactions
10.3: Nucleophilic and Electrophilic Addition and Abstraction

This page titled 10: Reaction Mechanisms is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

1
10.1: Oxidative Addition and Reductive Elimination
Learning Objectives
In this lecture you will learn the following
The oxidative addition reactions.
The reductive elimination reactions.
Various mechanistic pathways prevalent for these reactions.

Oxidative addition (OA) is a process that adds two anionic ligands e. g. A and B, that originally are a part of a A-B molecule, like
in H2 or Me−I, on to a metal center and is of significant importance from the perspective of both synthesis and catalysis. The exact
reverse of the same process, in which the two ligands, A and B, are eliminated from the metal center forming back the A−B
molecule, is called the reductive elimination (RA). As A and B are anionic X type ligands, the oxidative addition is accompanied
by an increase in the coordination number, valence electron count as well as in the formal oxidation state of the metal center by two
units. The oxidative addition step may proceed by a variety of pathways. It requires the metal center to be both coordinatively
unsaturated and electron deficient.

Oxidative addition transfers a single mononuclear metal center having 16 VE to a 18 VE species upon oxidative addition. Another
frequently observed pathway is that a 18 VE complex looses a ligand to become a 16 VE species which then undergoes an
oxidative addition. Apart from above two types, another possible pathway for oxidative addition proceeds as a binuclear oxidative
addition in which each of the two metal centers undergo change in oxidation state, electron count and coordination number by one
unit instead of two. This type of a binuclear oxidative addition is observed for a 17 VE metal complex or for a binuclear 18 VE
metal complex having a metal−metal bond and, for which the metal has a stable oxidation state at a higher positive oxidation state
by one unit.

It is interesting to note that in the oxidative addition the breakage of A−B σ−bond occurs as a result of a net transfer of electrons
from the metal center to a σ*−orbital of the A−B bond, thus resulting in the formation of the two new M−A and M−B bonds. The
oxidative addition is facilitated by electron rich metal centers having low oxidation state whereas the reductive elimination is
facilitated by metal centers in higher oxidation state.
Table 10.1.1 . Common types of oxidative addition reactions.

10.1.1 https://chem.libretexts.org/@go/page/172780
Abbreviations: Lin. = linear, Tet. = tetrahedral, Oct. = octahedral, Sq. Pl. = square planar, TBP = trigonal bipyramidal, Sq. Pyr. =
square pyramidal: 7-c, 8-c = 7- and 8-coordinate.
In principle, the oxidative addition is the reverse of reductive elimination, but in practice one may dominate over the other. Thus,
the favorability of one over the other is depends on the position of equilibrium, which is further dependent on the stability of the
two oxidation states of the metal and on the difference of bond strengths of A−B versus that of the M−A and M−B bonds. For
example, metal hydride complexes frequently undergo reductive elimination to give alkanes but rarely an alkane undergoes
oxidative addition to give an alkyl hydride complex. Along the same line, alkyl halides frequently undergo oxidative addition to a
metal giving metal−alkyl halide complexes but these complexes rarely reductively eliminate to give back alkyl halides. Usually the
oxidative addition is more common for 3rd row transition metals because they tend to possess stronger metal ligand bond strengths.
The oxidative addition is also favored by strong donor ligands, as they stabilize the higher oxidation state of the metal. The
oxidative addition reaction can expand beyond transition metals as observed in the case of the Grignard reagents as well as for
some main group elements.
Oxidative addition may proceed by several pathways as discussed below.

Concerted oxidative addition pathway


Oxidative addition may proceed by a concerted 3−centered associative mechanism involving the incoming ligand with the metal
center. Specifically, the addition proceeds by the formation of a σ−complex upon binding of an incoming ligand say, H2, followed
by the cleavage of the H−H bond as a result of the back donation of electrons from the metal to the σ*−orbital of the H−H bond.
Such type of addition is common for the H−H, C−H and Si−H bonds. As expected these proceed by two steps (i) the formation of a

10.1.2 https://chem.libretexts.org/@go/page/172780
σ−complex and (ii) the oxidation step. For example, the oxidative addition of H2 to Vaska’s complex (PMe3)2Ir(CO)Cl proceeds by
this pathways.

SN2 pathway
This pathway of oxidative addition is operational for the polarized AB type of ligand substrates like the alkyl, acyl, allyl and benzyl
halides. In this mechanism, the LnM fragment directly donates electrons to the σ*−orbital of the A−B bond by attacking the least
electronegative atom, say A, of the AB molecule and concurrently initiating the elimination of the most electronegative atom of the
AB molecule in its anionic form, B−. These reactions proceed via a polar transition state that is accompanied by an inversion of the
stereochemistry at the atom of attack by the metal center and are usually accelerated in polar solvents.

Radical pathway
This type of oxidative addition proceeds via a by radical pathway that generally are vulnerable to the presence of impurities. The
radical processes can be of non−chain and chain types. In a non−chain type of mechanism, the metal (M) transfer one electron to
the σ*−orbital of the RX bond resulting in the formation of a radical cation M+• and a radical anion RX−•. The generation of the
two radical fragments occurs by the way of the elimination of the anion X− from the radical anion RX−• leaving behind the radical
R• while the subsequent reaction of X− anion with the radical cation M+• generates the other radical MX• in the course of the
reaction. Such type of non−chain type of oxidative addition is observed for the addition of the alkyl halide to Pt(PPh3)3 complexes.
f ast

PtL −−→ PtL


3 2

slow
∙ ∙− ∙ ∙
PtL + RX ⟶ PtL + RX −−→ PtXL + R
2 2 2

f ast
∙ ∙
PtXL + R−−→ RPtXL
2 2

The other type in this category is the chain radical type reaction that is usually observed for the oxidative addition of EtBr and
PhCH2Br to the (PMe3)2Ir(CO)Cl complex. For this process a radical initiator is required and the reaction proceeds along a series
of known steps common to a radical process.
∙ I II∙
R + IR Cl(CO)L ⟶ RIr Cl(CO)L
2 2

II∙ III ∙
RIr Cl(CO)L + RX ⟶ RXIr Cl(CO)L +R
2 2


2R ⟶ R
2

Ionic pathway
This is kind of pathway for the oxidative addition reaction is common to the addition of hydrogen halides (HX) in its dissociated
H+ and X− forms. The ionic pathways are usually of the following two types (i) the ones in which the starting metal complex adds
to H+ prior to the addition of the halide X−and (ii) the other type, in which the halide anion X− adds to the starting metal complex
first, and then the addition of proton H+ occurs on the metal complex.

10.1.3 https://chem.libretexts.org/@go/page/172780
Reductive Elimination
The reductive eliminations are reverse of the oxidative addition reactions and are accompanied by the reduction of the formal
oxidation state of the metal and the coordination numbers by two units. The reductive eliminations are commonly observed for d8
systems, like the Ni(II), Pd(II) and Au(III) ions and the d6 systems, like the Pt(IV), Pd(IV), Ir(III) and Rh(III) ions. The reaction
may proceed by the elimination of several groups.

Ln MRH ⟶ Ln M + R−H

Ln MR ⟶ Ln M + R−R
2

Ln MH(COR) ⟶ Ln M + RCHO

Ln MR(COR) ⟶ Ln M + R CO
2

Ln MR(SiR ) ⟶ Ln M + R−SiR
3 3

Binuclear Reductive Elimination


Similar to what has been observed in the case of binuclear oxidative addition, the binuclear reductive elimination is also observed
in some instances. As expected, the oxidation state and the coordination number decrease by one unit in the binuclear reductive
elimination pathway.
heat

2 MeCH=CHCu(PBu ) −−→ MeCH=CHCH=CHMe


3

ArCOMn(CO) + HMn(CO) ⟶ ArCHO + Mn (CO)


5 5 2 10

Problems
1. What kind of metal centers favor oxidative addition?
Ans: Electron rich low valent metal centers.
2. Complete the sentence correctly.
(a) Reductive elimination is frequently observed in coordinatively saturated/unsaturated metal complexes.
(b) Reductive elimination is accompanied by increase/decrease in the oxidation state of the metal.
(c) Oxidative addition is accompanied by increase/decrease in the coordination number of the metal
Ans:
(a) Saturated.
(b) Decrease in the oxidation state by two units.
(c) Increase in the coordination number by two units
3. State the various mechanistic pathways involved in oxidative addition reactions.
Ans: Concerted oxidative addition, SN2 mechanism, radical and ionic mechanism.
4. Complete the reaction.

10.1.4 https://chem.libretexts.org/@go/page/172780
Ans:

Self Assessment test


1. What kind of metal centers favor reductive elimination?
Ans: Electron deficient high valent metal centers
2. Complete the sentence correctly.
(a) Oxidative addition is frequently observed in coordinatively saturated/unsaturated metal complexes.
(b) Oxidative addition is accompanied by increase/decrease in the oxidation state of the metal.
(c) Reductive elimination is accompanied by increase/decrease in the coordination number of the metal.
Ans:
(a) Unsaturated.
(b) Increase in the oxidation state by two units.
(c) Decrease in the coordination numbers by two units.
3. How does the geometry of the square planar complexes change upon oxidative addition reactions?
Ans: Square planar to octahedral.
4. Complete the reaction.

Ans:

Summary
The oxidative addition and the reductive elimination reactions are like the observe and reverse of the same coin. The oxidative
addition is generally observed for metal centers with low oxidation state and is usually accompanied by the increase in the
oxidation state, the valence electron count and the coordination number of the metal by two units. Being opposite, the reductive
elimination is seen in the case of the metal centers with higher oxidation state and is accompanied by the decrease in the oxidation
state, the valence electron count and the coordination number of the metal by two units. The oxidative addition may proceed by a
variety of pathways that involve concerted, ionic and the radical based mechanisms. More interestingly, the oxidative addition and

10.1.5 https://chem.libretexts.org/@go/page/172780
reductive elimination reactions are not solely restricted to the mononuclear metal complexes but can also be observed for the
binuclear complexes.

This page titled 10.1: Oxidative Addition and Reductive Elimination is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

10.1.6 https://chem.libretexts.org/@go/page/172780
10.2: Insertion and Elimination Reactions
Learning Objectives
In this lecture you will learn the following
The insertion and elimination reactions.
The various mechanistic pathways by which these reactions proceed.
Their relevance in some important catalytic cycles.

Unlike what we have learned about the oxidative addition and the reductive elimination reactions, that facilitate the addition or
removal of 1−electron and 2−electron ligands on to a metal center, the insertion and the elimination reactions perform the
subsequent transformation of these ligands from within the same coordination sphere of a metal. Thus, in an insertion reaction a
metal bound 2 electron A=B type of a ligand can insert on to a M−X bond resulting in a new metal bound 1−electron ligand like,
M−A−B−X, which is formed as a result of the formations of the M−A and B−X bonds. The insertion reaction thus leads to the
generation of one vacant site created at the initial metal bound A=B site. Thus a primary requirement for reverse elimination
reaction to occur is the presence of a cis vacant site.
Insertions are of two types, 1,1−insertion and 1,2−insertion. In 1,1−insertion both the metal M and the ligand X of the M−X bond
end up on the same atom like in the M−A(X)−B moiety formed after the insertion of the A=B molecule in the M−X bond, whereas
in the 1,2−insertion, these end up on the adjacent atoms like in the M−A−B−X moiety formed after the insertion of the A=B
molecule in the M−X bond. The type of insertion depends on the type of the ligand undergoing the insertion like η1−ligand
showing 1,1−insertion and η2−ligands showing 1,2−insertion. For example, the CO ligand exclusively undergoes 1,1−insertion
while the C2H4 ligand undergoes 1,2−insertion. The SO2 ligand remains the only exception as it can bind by both η1−(by S−donor
site) and η2−(by S− and O−donor sites) modes and thus shows both type of insertions.

Though the insertion and the elimination reactions are mutually reversible, owing to the thermodynamical reasons one is favored
over the other. For example, SO2 is known to insert into the M−R bond with no report of its elimination is known of it, while for N2
ligand, no report of its insertion is known but it’s elimination from a M−N=N−R bond is known.

M−R + SO ⟶ M−SO R
2 2

M−N=N−Ar ⟶ M−Ar=N
2

The CO ligand inserts readily into a metal−alkyl bond. Sterically demanding substituents (R) is found to accelerate the reaction as a
bulky R group in an acyl moiety in the final M−CO−R bond is far removed from the metal center than that in the starting M−R
bond.

10.2.1 https://chem.libretexts.org/@go/page/172781
Isonitriles readily insert into M−R and M−H bonds giving η2-bound iminoacyls.

Olefins usually inserts across a M−H bond and such insertions are of relevance to the commercially important olefin
polymerization process. In certain cases the 1,2−insertions of olefins give species exhibiting agostic insertions.

ß-Elimination
β−elimination is just a reverse of 1,2−insertion and is a major cause of decomposition of metal alkyl bond having a b−hydrogen
atom. A pre-requisite for the β−elimination reaction to occur is the presence of an adjacent vacant site next to the metal alkyl bond
undergoing the β−elimination. The β−elimination step results in the formation of a metal hydride species that also contain a metal
bound olefin moiety.

α-elimination
In absence of a β−hydrogen, a metal bound alkyl moiety may undergo the cleavage of a C−H bond at the α, γ and δ positions. For
example, a methyl moiety may α−eliminate to give a metal bound methylene hydride moiety.

10.2.2 https://chem.libretexts.org/@go/page/172781
Problems
1. Give an example of a ligand that undergo 1,1−insertion.
Ans: CO
2. Give an example of a ligand that undergo 1,2−insertion.
Ans: C2H4
3. Give an example of a ligand that undergo both the 1,1−insertion and the 1,2−insertion reactions.
Ans: C2H4

Self Assessment test


1. On what type of bonds does CO insertion usually occur?
Ans: Metal−alkyl (M−R) bonds.
2. On what type of bonds does isonitrile insertion usually occur?
Ans: Metal-alkyl (M-R) and the metal-hydride (M-H) bonds.
3. On what type of bonds does C2H4 insertion usually occur?
Ans: Metal-hydride (M-H) bonds.
4. State an important requirement for the occurrence of an elimination reaction.
Ans: The presence of an adjacent vacant site.

Summary
Insertion and elimination reactions are important sequences that carryout transformation of the metal bound ligands to the
corresponding product from within the same coordination sphere of the metal center and thus together represent key steps of an
overall catalytic cycle. The insertions are highly ligand dependent and may proceed by 1,1−insertion and 1,2−insertion
mechanisms. Elimination reactions are just reverse of the insertion reactions and they too may proceed by several pathways. The
two pathways that are commonly observed are the β−elimination and the α−eliminations, even though other type of elimination
pathways exists.

This page titled 10.2: Insertion and Elimination Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

10.2.3 https://chem.libretexts.org/@go/page/172781
10.3: Nucleophilic and Electrophilic Addition and Abstraction
Learning Objectives
In this lecture you will learn the following
Ligand activation by metal that leads to a direct external attack at the ligand.
Nucleophilic addition and nucleophilic abstraction reactions.
Electrophilic addition and electrophilic abstraction reactions.

The nucleophilic and electrophilic substitution and abstraction reactions can be viewed as ways of activation of substrates to allow
an external reagent to directly attack the metal activated ligand without requiring prior binding of the external reagent to the metal.
The attacking reagent may be a nucleophile or an electrophile. The nucleophilic attack of the external reagent is favored if the LnM
fragment is a poor π−base and a good σ−acid i.e., when the complex is cationic and/or when the other metal bound ligands are
electron withdrawing such that the ligand getting activated gets depleted of electron density and can undergo an external attack by a
nucleophile Nu−, like LiMe or OH−. The attack of the nucleophiles may result in the formation of a bond between the nucleophiles
and the activated unsaturated substrate, in which case it is called nucleophilic addition, or may result in an abstraction of a part or
the whole of the activated ligand, in which case it is called the nucleophilic abstraction. The nucleophilic addition and the
abstraction reactions are discussed below.

Nucleophilic addition
An example of a nucleophilic addition reaction is shown below.

Carbon monoxide (CO) as a ligand can undergo nucleophilic attack when bound to a metal center of poor π−basicity, as the carbon
center of the CO ligand is electron deficient owing to the ligand to metal σ−donation not being fully compensated by the metal to
ligand π−back donation. Thus, activated CO ligand undergoes nucleophilic attack by the lithium reagent to give an anionic acyl
ligand, which upon alkylation generates the famous Fischer carbene complex.

Nucleophilic abstraction
An example of a nucleophilic abstraction reaction is shown below.

Electrophilic addition
Similar to the nucleophilic addition and abstraction reactions, the electrophilic counterparts of these reactions also exist. An
electrophilic attack is favored if the LnM fragment is a good π−base and a poor σ−acid i.e., when the complex is anionic with the
metal center at low−oxidation state and/or when the other metal bound ligands are electron donating such that the ligand getting
activated becomes electron rich from the π−back donation of the metal center and thus can undergo an external attack by an
electrophile E+ like H+ and CH3I. The attack of the electrophiles may result in the formation of a bond between the electrophile and

10.3.1 https://chem.libretexts.org/@go/page/172782
the activated unsaturated substrate, in which case it is called electrophilic addition, or may result in an abstraction of a part or the
whole of the activated ligand, in which case it is called the electrophilic abstraction.

Electrophilic abstraction
An example of an electrophilic abstraction reaction is shown below.

Alkyl abstractions are often achieved by Hg2+ that can proceed in two ways, (i) by an attack at the α−carbon of a metal alkyl bond
leading to an inversion of configuration at the alkyl carbon and (ii) by an attack at the metal center leading to retention of
configuration at the alkyl carbon. The inversion of configuration proceeds by the following pathway.

The retention of configuration proceeds by the following pathway.

10.3.2 https://chem.libretexts.org/@go/page/172782
This page titled 10.3: Nucleophilic and Electrophilic Addition and Abstraction is shared under a CC BY-NC-SA 4.0 license and was authored,
remixed, and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts
platform; a detailed edit history is available upon request.

10.3.3 https://chem.libretexts.org/@go/page/172782
CHAPTER OVERVIEW

11: Applications
11.1: Homogeneous Catalysis - I
11.2: Homogeneous Catalysis - II

This page titled 11: Applications is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S. Balakrishna &
Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1
11.1: Homogeneous Catalysis - I
Learning Objectives
In this lecture you will learn the following
The application of organometallic complexes in homogeneous catalysis.
Alkene isomerization.
Alkene and the arene hydrogenations.
Transfer hydrogenation.

One of the most important exploits of the organometallic chemistry is its application in the area of homogeneous catalysis. The
field has now expanded its territory to accommodate in equal measures many large-scale industrial processes as well as numerous
small scale reactions of the day-to-day organic synthesis. A few representative examples of organometallic catalysis are outlined
below.

Alkene Isomerization
Alkene isomerization is a transformation that involve a shift of a double bond to an adjacent position followed by 1,3−migration of
a H atom. The isomerization reaction is transition metal catalyzed.

The alkene isomerization reaction may proceed by two pathways, (i) one through a η1−alkyl intermediate and (ii) the other through
η3−allyl intermediate. In the η1−alkyl pathway, an alkene first binds to a metal at a vacant site next to M−H bond and then
subsequently undergoing an insertion into the M−H bond thus creating back the vacant site. The resultant species then undergoes a
H atom transfer from the alkyl moiety to give the isomerized olefin along with the regeneration of the M−H species.

The η3−allyl mechanism requires the presence of two vacant sites. This mechanism goes through a η3−allyl intermediate formed by
a C−H activation at the allylic position of the olefin formed after binding to the metal and alongside leads to the formation of a
M−H bond. Subsequent H transfer from the metal back to the η3−allyl moiety leads to the alkene isomerized product.

11.1.1 https://chem.libretexts.org/@go/page/172787
Alkene Hydrogenation
The transition metal catalyzed alkene hydrogenation reactions are of significant industrial and academic interest. These reactions
involve the H2 addition on a C=C bond of olefins to give alkenes. The alkene hydrogenation may proceed by three different
pathways namely the (i) oxidative addition (ii) heterolytic activation and (iii) the homolytic activation of the H2 molecule.
The oxidative addition pathway is commonly observed for the Wilkinson’s catalyst (PPh3)3RhCl and is the most studied among all
of the three pathways that exist. The catalytic cycle initiates with the oxidative addition of H2 followed by alkene coordination. The
resultant species subsequently get converted to the hydrogenated product.

The second pathway proceeds by the heterolytic activation of the H2 molecule and requires the presence of a base like NEt3, which
facilitates the heterolytic cleavage by abstracting a proton from the H2 molecule and leaving behind a hydride H− ion that

11.1.2 https://chem.libretexts.org/@go/page/172787
participates in the hydrogenation reaction. This type of mechanism is usually followed by the (PPh3)3RuCl2 type of complexes.

X−Y + M−Z ⟶ M−X + Y−Z

Homolytic cleavage of H2 is the third pathway for the alkene hydrogenation. It is the rarest of all the three methods and proceeds
mainly in a binuclear pathway. Paramagnetic cobalt based Co(CN)53− type catalysts carries out alkene hydrogenation by this
pathway via the formation of the HCo(CN)53−species.

Arene hydrogenations
Examples of homogeneous catalysts for arene hydrogenation are rare though it is routinely achieved using catalysts like Rh/C
under the heterogeneous conditions. A representative example of a homogeneous catalyst of this class is (η3−allyl)Co[P(OMe)3]3
that carry out the deuteration of benzene to give the all-cis-C6H6D6 compound.

Transfer hydrogenation
This is a new kind of a hydrogenation reaction in which the source of the hydrogen is not the H2molecule but an easily oxidizable
substrate like isopropyl alcohol. The method is particularly useful for the reduction of ketones and imines but not very effective for
the olefins.

Me CHOH + RCH=CH ⟶ Me C=O + RCH CH


2 2 2 2 3

Summary
The applications of organometallic compounds in homogeneous catalysis have transcend the boundaries of industry to meet the
day-to-day synthesis in laboratory scale reactions. The alkene isomerization is one such application of homogeneous catalysis by
the transition metal organometallic complexes. The hydrogenation reactions of alkene, arene, ketone and imine substrates are

11.1.3 https://chem.libretexts.org/@go/page/172787
achieved by several types of the transition metal organometallic catalysts. They also proceed by different mechanisms involving
oxidative addition, heterolytic and homolytic cleavages of the H−H bond. The transfer hydrogenation reaction uses easily
oxidizable substrates like i−PrOH instead of H2 as the hydrogenation source.

This page titled 11.1: Homogeneous Catalysis - I is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

11.1.4 https://chem.libretexts.org/@go/page/172787
11.2: Homogeneous Catalysis - II
Learning Objectives
In this lecture you will learn the following
The hydroformylation reaction and its mechanism.
The C−C cross-coupling reactions and their mechanisms.

It is truly an exciting time for the field of organometallic chemistry as its potentials in homogeneous catalysis are being realized in
an unprecedented manner. The growth in the field organometallic chemistry has been rightly acknowledged by the award of three
Nobel prizes in over a decade in the areas of asymmetric hydrogenation (Nyori and Knowles in 2001), olefin metathesis (Grubbs,
Schrock and Chauvin in 2006) and palladium mediated C−C cross coupling reactions (Suzuki, Negishi and Heck, 2010). A few
representative examples of such landmark discoveries of homogeneous catalysis by organometallic compounds are discussed
below.

Hydroformylation reaction
Hydroformylation, popularly known as the "oxo" process, is a Co or Rh catalyzed reaction of olefins with CO and H2 to produce
the value-added aldehydes.

The reaction, discovered by Otto Roelen in 1938, soon assumed an enormous proportion both in terms of the scope and scale of its
application in the global production of aldehydes. The metal hydride complexes namely, the rhodium based HRh(CO)(PPh3)3 and
the cobalt based HCo(CO)4 complexes, catalyzed the hydroformylation reaction as shown below.

11.2.1 https://chem.libretexts.org/@go/page/172788
C−C cross-coupling reactions
The palladium catalyzed cross-coupling reactions are a class of highly successful reactions with applications in the organic
synthesis to have emerged recently. The reactions carry out a coupling of the aryl, vinyl or alkyl halide substrates with different
organometallic nucleophiles and as such encompasses a family of C−C cross-coupling reactions that are dependent on the nature of
nucleophiles like that of the B based ones in the Suzuki-Miyuara coupling, the Sn based ones in the Stille coupling, the Si based
ones in the Hiyama coupling, the Zn based ones in the Negishi coupling and the Mg based ones in the Kumada coupling reactions
(Figure 11.2.1).

11.2.2 https://chem.libretexts.org/@go/page/172788
Figure 11.2.1 : Various types of the palladium mediated C−C cross-coupling reactions.
An unique feature of these reactions is the exclusive formation of the cross-coupled product without the accompaniment of any
homo-coupled product. Another interesting feature of these coupling reactions is that they proceed via a common mechanism
involving three steps that include the oxidative addition, the transmetallation and the reductive elimination reactions (Figures
11.2.2 and 11.2.3).

Figure 11.2.2 : A general catalytic cycle for the palladium mediated C−C cross-coupling reactions.

11.2.3 https://chem.libretexts.org/@go/page/172788
Figure 11.2.3 : A catalytic cycle for the palladium mediated Heck coupling reaction.

Summary
Organometallic complexes play a pivotal role in several successful homogeneous catalysis reactions like that of the
hydroformylation and the C−C cross-coupling reactions. These reactions are important because of the fact that both of the
hydroformylation and the C−C cross-coupling reactions give more value added products compared to the starting reactants. The
palladium catalyzed C−C cross-coupling reactions are a class of highly successful reactions that have permanently impacted the
area of organic synthesis in a profound way to an extent that the 2010 Nobel prize has been conferred on one of these reactions
thereby recognizing the importance of the C−C cross-coupling recations.

This page titled 11.2: Homogeneous Catalysis - II is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M.
S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

11.2.4 https://chem.libretexts.org/@go/page/172788
CHAPTER OVERVIEW

12: Physical Methods in Organometallic Chemistry


12.1: Characterization of Organometallic Complexes

This page titled 12: Physical Methods in Organometallic Chemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

12.1 https://chem.libretexts.org/@go/page/172800
12.1: Characterization of Organometallic Complexes
Learning Objectives
In this lecture you will learn the following
The characterization techniques of organometallic compounds.
The NMR analysis of these compounds.
The IR analysis of these compounds.
The X-ray single crystal diffraction studies of these compounds.

The characterization of an organometallic complex involves obtaining a complete understanding of the same right from its
identification to the assessment of its purity content, to even elucidation of its stereochemical features. Detailed structural
understanding of the organometallic compounds is critical for obtaining an insight on its properties and which is achieved based on
the structure-property paradigm.

Synthesis and isolation


Synthesis and isolation are two very important experimental protocols in the overall scheme of things of organometallic chemistry
and thus these needs to be performed carefully. The isolation of the organometallic compounds is essential for their characterization
and reactivity studies. Fortunately, many of the methods of organic chemistry can be used in organometallic chemistry as the
organometallic compounds are mostly nonvolatile crystalline solids at room temperature and atmospheric pressure though a few
examples of these compounds are known to exist in the liquid [(CH3C5H4Mn(CO)3] and even in the vapor [Ni(CO)4] states. The
organometallic compounds are comparatively more sensitive to aerial oxygen and moisture, and because of which the manipulation
of these compounds requires stringent experimental skills to constantly provide them with anaerobic environment for their
protection. All of these necessities led to the development of the so-called special Schlenk techniques, requiring special glasswares
and which in conjunction with a high vacuum line and a dry box allow the lab bench-top manipulation of these compounds.
Successful isolation of organometallic compounds naturally points to the need for various spectroscopic techniques for their
characterizations and some of the important ones are discussed below.

1H NMR spectroscopy
The 1H NMR spectroscopy is among the extensively used techniques for the characterization of organometallic compounds. Of
particular interest is the application of 1H NMR spectroscopy in the characterization of the metal hydride complexes, for which the
metal hydride moiety appear at a distinct chemical shift range between 0 ppm to −40 ppm to the high field of tetramethyl silane
(TMS). This upfield shift of the metal hydride moiety is attributed to a shielding by metal d−electrons and the extent of the upfield
shift increases with higher the dn configuration. Chemical shifts, peak intensities as well as coupling constants from the through-
bond couplings between adjacent nuclei like that of the observation of JP-H, if a phosphorous nucleus is present within the coupling
range of a proton nucleus, are often used for the analysis of these compounds. The 1H NMR spectroscopy is often successfully
employed in studying more complex issues like fluxionality and diastereotopy in organometallic molecules (Figure 12.1.1).

12.1.1 https://chem.libretexts.org/@go/page/172794
Figure 12.1.1 : Different phosphorous-proton coupling patterns in various iridium hydride complexes.
The paramagnetic organometallic complexes show a large range of chemical shifts, for example, (η6−C6H6)2V exhibits proton
resonances that extend even up to 290 ppm.

13C NMR spectroscopy


Although the natural abundance of NMR active 13C (I = ½) nuclei is only 1 %, it is possible to obtain a proton decoupled 13C{1H}
NMR spectra for most of the organometallic complexes. In addition, the off−resonance 1H decoupled 13C experiments yield 1JC-H
coupling constants, which contain vital structural information, and hence are very critical to the 13C NMR spectral analysis. For
example, the 1JC-H coupling constants directly correlate with the hybridization of the C−H bonds with sp center exhibiting a 1JC-H
coupling constant of ~250 Hz, a sp2 center of 160 Hz and a sp3 center of 125 Hz. Similar to what is seen in 1H NMR, a
phosphorous−carbon coupling is also observed in a 13C NMR spectrum with the trans coupling (~100 Hz) being larger than the cis
coupling (~10 Hz).
31
P NMR spectroscopy
The 31P NMR spectroscopy, which in conjunction with 1H and 13C NMR spectroscopies, is a useful technique in studying the
phosphine containing organometallic complexes. The 31P NMR experiments are routinely run under 1H decoupled conditions for
simplification of the spectral features that allow convenience in spectral analysis. Thus, for this very reason, many mechanistic
studies on catalytic cycle are conveniently undertaken by 31P NMR spectroscopy whenever applicable.

IR spectroscopy
Qualitative to semi-quantitative analysis of organometallic compounds using IR spectroscopy are performed whenever possible. In
general the signature stretching vibrations for chemical bonds are more conveniently looked at in these studies. The frequency (ν )
of a stretching vibration of a covalent bond is directly proportional to the strength of the bond, usually given by the force constant
(k) and inversely proportional to the reduced mass of the system, which relates to the masses of the individual atoms.
−−−
1 k
ν = √
2πC mr

m1 m2
mr =
m1 + m2

12.1.2 https://chem.libretexts.org/@go/page/172794
The organometallic compounds containing carbonyl groups are regularly studied using IR spectroscopy, and in which the CO

peaks appear in the range between 2100−1700 cm-1 as distinctly intense peaks.

Crystallography
The solid state structure elucidation using single crystal diffraction studies are extremely useful techniques for the characterization
of the organometallic compounds and for which the X-ray diffraction and neutron diffraction studies are often undertaken. As these
methods give a three dimensional structural rendition at a molecular level, they are of significant importance among the various
available characterization methods. The X-ray diffraction technique is founded on Bragg’s law that explains the diffraction pattern
arising out of a repetitive arrangement of the atoms located at the crystal lattices.

2d sin θ = nλ

A major limitation of the X-ray diffraction is that the technique is not sensitive enough to detect the hydrogen atoms, which appear
as weak peaks as opposed to intense peaks arising out of the more electron rich metal atoms, and hence are not very useful for
metal hydride compounds. Neutron diffraction studies can detect hydrogens more accurately and thus are good for the analysis of
the metal hydride complexes.

Summary
Along with the synthesis, the isolation and the characterization protocols are also integral part of the experimental organometallic
chemistry. Because of their air and moisture sensitivities, specialized experimental techniques that succeed in performing the
synthesis, isolation and storage of these compounds in an air and moisture-free environment are often used. The organometallic
compounds are characterized by various spectroscopic techniques including the 1H NMR, 13C NMR and IR spectroscopies and the
X-ray and the neutron diffraction studies.

This page titled 12.1: Characterization of Organometallic Complexes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

12.1.3 https://chem.libretexts.org/@go/page/172794
CHAPTER OVERVIEW

13: Multiply-Bonded Ligands


13.1: Metal-Carbenes
13.2: Metal-Carbynes

This page titled 13: Multiply-Bonded Ligands is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

1
13.1: Metal-Carbenes
Learning Objectives
In this lecture you will learn the following
The metal−ligand multiple bonding and their relevance in.
The Fischer type carbene complexes.
The Schrock type carbene complexes.

The organometallic compounds containing metal−ligand multiple bonds of the types, M=X and M≡X (X = C, N, O) are of current
interest as they are valuable intermediates in many important catalytic cycles. In this regard, considerable attention has been paid
towards developing an understanding of the metal−ligand multiply bonded systems like that of the metal carbene LnM=CR2 type
complexes and of the metal carbyne LnM≡CR type complexes. A detailed account of the metal−carbene complexes is presented in
this chapter.

Metal carbene complexes


Carbenes are highly reactive hexavalent species that exist in two spin states, i.e. (i) in a singlet form ( ), in which two electrons
are paired up and (ii) in a triplet form ( ), in which the two electrons remain unpaired. Of the two, the singlet form is the more
reactive one. The instability of carbene accounts for its unique reactivity like that of the insertion reaction, which has aroused
significant interest in recent years. The singlet carbene and the triplet carbene bind differently to metals, with the singlet one
yielding Fischer type carbene complexes while the triplet one yielding Schrock type carbene complexes (Figure 13.1.1).

Figure 13.1.1 : Metal−ligand multiple bonding in the Fischer and Schrock carbene system.

13.1.1 https://chem.libretexts.org/@go/page/172801
The LnM=CR2 type Fischer carbene complexes comprise of two dative covalent interactions that include (i) a LnM←CR2 type
ligand to metal σ−donation and (ii) a LnM→CR2 type metal to ligand π−back donation. The Fischer type carbene complexes are
usually formed with metal centers at a low oxidation state. These are also commonly observed for the more electron rich
late−transition metals that participate in the LnM→CR2 type metal to ligand π−back donation. Another characteristic of the Fischer
type carbene complex is the presence of the heteroatom substituents like R = OMe or NMe2 on the carbene CR2 moiety which
makes the carbene carbon significantly cationic (δ+) to facilitate the LnM→CR2 type metal to ligand π−back donation.

Figure 13.1.2 : Fischer and Schrock type complexes.


Similarly, the LnM=CR2 type Schrock carbene complexes comprise of two covalent interactions that involve one electron donation
towards the σ−bond from each of the metal LnM and the carbene CR2fragments. Schrock carbene complexes are thus formed with
the metal centers having high oxidation state and are usually observed for electron deficient early−transition metals (Figure
13.1.2).

Carbene complexes can be prepared by the following methods.


i. by the reaction with electrophiles

ii. by H−/H+ abstraction reactions as shown below

iii. from low−valent metal complexes

13.1.2 https://chem.libretexts.org/@go/page/172801
Ln M + CH N ⟶ Ln MCH +N
2 2 2 2

Because of the electronically different metal−ligand interaction that exist between the LnM and the carbene CR2 moiety, the
reactivity of Fischer and Schrock carbene complexes are completely different. For example, the Fischer type carbene complexes
undergo attack by nucleophiles at its carbene−C center.

The Schrock type carbene complexes on the other hand undergo attack by electrophiles at its carbene−C center.

Summary
The metal−ligand multiple bonding is of significant interest as many of the compounds containing such bonds are important
intermediates in various catalytic cycles. The metal−ligand doubly bonded carbene systems can exist in two varieties like the
Fischer type and the Schrock type carbene complexes. Due to their different electronic structures, the reactivities of these Fischer
type and the Schrock type carbene complexes differ significantly, with the former undergoing nucleophilic attack while the later
undergo electrophilic attack at their respective carbene−C centers.

This page titled 13.1: Metal-Carbenes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

13.1.3 https://chem.libretexts.org/@go/page/172801
13.2: Metal-Carbynes
Learning Objectives
In this lecture you will learn the following
The metal−ligand multiple bonding and their relevance in.
The Fischer type carbyne complexes.
The Schrock type carbyne complexes.

The metal−ligand multiply bonded systems even extended beyond the doubly bonded Fischer and the Schrock carbenes to the
triply bonded LnM≡CR type Fischer carbyne and the Schrock carbyne complexes. Similar to carbene that exists in a singlet and a
triplet spin state, the carbyne also exists in two other spin states i.e in a doublet and a quartet form.
Upon binding to the metal in its doublet spin state as in the Fischer carbene system, the carbyne moiety donates two electrons via
its sp hybridized lone pair containing orbital to an empty metal d orbital to yield a LnM←CR type ligand to metal dative bond. It
also makes a covalent π−bond through one of its singly occupied pz orbital with one of the metal d orbitals. The carbyne−metal
interaction consist of two ligand to metal interactions namely a dative one and a covalent one that together makes the carbyne
moiety a LX type of a ligand. In addition to these two types of ligand to metal bonding interactions, there remains an empty py
orbital on the carbyne−C atom that can accommodate electron donation from a filled metal d orbital to give a metal to ligand
π−back bonding interaction (Figure 13.2.1).

Figure 13.2.1 : Metal−ligand multiple bonding in the Fischer and Schrock carbyne system
Analogously, in the quartet carbyne spin state in the Schrock carbyne systems three covalent bonds occur between the singly
occupied sp, pyand pz orbitals of carbyne−C moiety with the respective singly occupied metal d orbitals (Figure 13.2.1).
Similar to what has been observed earlier in the case of the Fischer carbenes and Schrock carbenes, the Fischer carbyne complexes
are formed with metal centers in lower oxidation states for e.g. as in Br(CO)4W≡CMe, while the Schrock carbyne complexes are
formed with metals in higher oxidation state, e.g. as in (t−BuO)3W≡Ct−Bu.
Carbyne complexes can be prepared by the following methods.

13.2.1 https://chem.libretexts.org/@go/page/172802
i. The Fischer carbyne complexes can be prepared by the electrophilic abstraction of a methoxy group from a methoxy methyl
substituted Fischer carbene complex.
+ −
L(CO) M=C(OMe)Me + 2 BX ⟶ [L(CO) M ≡ CMe] BX + BX (OMe) ⟶ X(CO) M ≡ CMe
4 3 4 4 2 4

ii. Schrock carbynes can be prepared by the deprotonation of a α−CH bond of a metal−carbene complex.
(i)P Me3

CPCl Ta=CHR −−−−−−−−−→ Cp(PMe )ClTa ≡ CR


2 3
(ii)P h3P =C H 2

iii. by an α−elimination reaction on a metal−carbene complex


(i)dmpe

Cp ⋅ Br Ta=CHt−Bu −−−−−−→ Cp ⋅ (dmpe)HTa ≡ Ct−Bu


2
(ii)N a/H g

iv. by metathesis reaction

(t−BuO) W ≡ W (Ot−Bu) + t−BuC ≡ Ct−Bu ⟶ 2 (t−BuO) W ≡ Ct−Bu


3 3 3

The reactivities of Fischer and the Schrock carbynes mirror that of the Fischer and Schrock carbenes. For example, the Fischer
carbyne undergo nucleophilic attack at the carbyne−C atom while the Schrock carbyne undergo electrophilic attack at the
carbyne−C atom.

Summary
The theme of metal−ligand multiple bonding extends beyond the doubly bonded Fischer and the Schrock carbene systems to even
triply bonded Fischer and the Schrock carbyne systems. The carbyne moieties in these Fischer and the Schrock carbyne systems
respectively exist in a doublet and a quartet spin state. The carbyne complexes are generally prepared from the respective carbene
analogues by the abstraction of alkoxy (OR), proton (H+), hydride (H−) moieties, the α−elimination reactions and the metathesis
reactions. The reactivity of the Fischer and the Schrock carbyne complexes parallel the corresponding Fischer and the Schrock
carbene counterparts with regard to their reactivities toward electrophiles and nucleophiles.

This page titled 13.2: Metal-Carbynes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S.
Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

13.2.2 https://chem.libretexts.org/@go/page/172802
CHAPTER OVERVIEW

14: Metathesis
14.1: Catalytic Applications of Organometallic Compounds- Alkene Metathesis
14.2: Credits

This page titled 14: Metathesis is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S. Balakrishna &
Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1
14.1: Catalytic Applications of Organometallic Compounds- Alkene Metathesis
Learning Objectives
In this lecture you will learn the following
The alkene metathesis reactions and their different variants.
The application of metal carbenes in alkene metathesis reactions.
The functional group tolerance, air and moisture sensitivity and high efficiency as important catalyst attributes for the
alkene metathesis reactions.

The application of organometallic chemistry in homogenous catalysis is progressively increasing with the fast pace of discovery of
new catalysts in the area. The benefits of organometallic catalysis have now percolated to all facets of the chemical world that span
from the confines of the industry to the day-to-day small scale use in organic synthesis in academic laboratories. Quite a few of
these applications of organometallic complexes in homogeneous catalysis have made a permanent imprint on the ever going
developmental process that is constantly transforming our day-to-day life. An example of such a success story is of alkene
metathesis, which is described in this chapter.

Alkene metathesis
Alkene metathesis reactions are gaining wide popularity in synthesizing unsaturated olefinic compounds as well as the unsaturated
polymeric counterparts. Central to this catalysis is a metal carbene intermediate that reacts with olefins to give different olefinic
compounds or even the unsaturated olefinic polymers depending upon the reaction conditions of the metathesis reaction.
Metathesis is an unusual transformation in which a C=C is broken and also formed during catalysis to generate new unsaturated
olefins.
′ ′ ′
RCH=CHR + R CH=CHR ↽−⇀
− 2 RCH=CHR

Though a large variety of metal−carbene catalysts have been developed for the metathesis reaction, only a few have been found to
be functional group tolerant. Thus a critical step in broadening the utility of metathesis reaction has been in developing catalysts
that are functional group tolerant. In this regard, the early-transition metal based carbene catalysts like that of the Ti based ones are
highly oxophilic and hence are intolerant to the functional groups. On the other hand, the more electron-rich Mo and W based
catalysts are of intermediate character. Finally, the late-transition metal based Ru catalysts are found to be exceptionally tolerant
toward functional groups but all the while exhibiting high reactivity toward olefinic bonds. In this context notable are the Grubb’s
Ru catalyst, which is easy to handle, and the Schrock’s Mo catalyst, which display high activity.

Figure 14.1.1 : The Grubb’s and the Schrock’s catalyst.


The metathesis reaction as such stands for a family of related reactions all of which involve a “cutting and stitching” of olefinic
bonds leading to different unsaturated products. When two different olefin substrates are used, the reaction is called the “cross
metathesis” owing to the fact that the olefinic ends are exchanged.
′ ′ ′
RCH=CHR + R CH=CHR −⇀
↽− 2 RCH=CHR

The metathesis reactions can even extend further to the conjugated dienes that can undergo Ring Closing Metathesis (RCM) in
systems where the ring strain is not too high in the final product. The reverse of Ring Closing Metathesis (RCM) is called the Ring
Opening Metathesis (ROM), and which is usually favored in the presence of large excess of C2H4.

14.1.1 https://chem.libretexts.org/@go/page/172808
The variants of metathesis often used in producing polymers are, (i) the Acyclic Diene Metathesis (ADMET) and (ii) the Ring
Opening Metathesis Polymerization (ROMP), in which the relief of ring-strain of cycloalkenes drives the polymerization reaction
forward. Both of these reactions, produce long chain polymers in a living fashion and as a result of which these reactions are useful
for producing block copolymers −(AAABBBB) n−.

Though several possibilities have been debated for the mechanism of the metathesis reaction, the one proceeding via a
metalacyclobutane intermediate has gained credence.

Several important industrial applications have emerged out of the metathesis reaction like that of the commercial synthesis of the
housefly pheromone.
Me(CH ) CH=CH + Me(CH ) CH=CH ⟶ Me(CH ) CH=C (CH ) Me[housefly pheromone] + C H +
2 7 2 2 12 2 2 7 2 12 2 4

other products

Similarly, the polycyclopentadiene polymer, which is formed from the Ring Opening Metathesis Polymerization (ROMP) of
dicyclopentadiene substrate, is used for bullet proof related applications because of its exceptional strength owing to its cross-
linked nature.

14.1.2 https://chem.libretexts.org/@go/page/172808
Summary
Alkene metathesis represents a distinct class of related chemical reactions that involve the “cutting and stitching” of olefinic bonds
to give unsaturated organic products. Depending upon the nature of the product formed, different type of alkene metathesis
reactions exist like the alkene metathesis, cross-metathesis, Ring Closing Metathesis (RCM), Ring Opening Metathesis (ROM),
Acyclic Diene Metathesis (ADMET), and the Ring Opening Metathesis Polymerization (ROMP). A commonality that runs through
all of these different varieties of the metathesis reaction is its mechanism that involves a catalytically active metal−carbene species.
The mechanism is said to be proceed via a 4−membered metalacyclobutane intermediate. The alkene metathesis has found
important applications in organic synthesis as well as in the chemical industry.

This page titled 14.1: Catalytic Applications of Organometallic Compounds- Alkene Metathesis is shared under a CC BY-NC-SA 4.0 license and
was authored, remixed, and/or curated by M. S. Balakrishna & Prasenjit Ghosh via source content that was edited to the style and standards of the
LibreTexts platform; a detailed edit history is available upon request.

14.1.3 https://chem.libretexts.org/@go/page/172808
14.2: Credits

This page titled 14.2: Credits is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by M. S. Balakrishna &
Prasenjit Ghosh via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

14.2.1 https://chem.libretexts.org/@go/page/172809
Index
A H R
Agostic alkyls hydroformylation reaction Reductive elimination
7.1: Transition Metal Alkyl Complexes 11.2: Homogeneous Catalysis - II 10.1: Oxidative Addition and Reductive Elimination
Hydrometallation
B 2.1: General Methods of Preparation S
Bridging hydrides Schrock carbynes
7.2: Metal Hydrides M 13.2: Metal-Carbynes
Mercury poisoning Schrock’s catalyst
C 5.2: Organometallic Compounds of Mercury 14.1: Catalytic Applications of Organometallic
Metal Carbonyls Compounds- Alkene Metathesis
Carbene insertion
6.2: Synthesis and Stability 8.1: Metal Carbonyls
concerted 3−centered associative metal hydrides T
mechanism 7.2: Metal Hydrides Transmetallation
2.1: General Methods of Preparation
10.1: Oxidative Addition and Reductive Elimination
O
D olefin polymerization U
Dewar−Chatt model 3.4: Zeigler Natta Polymerization Catalysts umpolung reactions
9.1: Metal Alkene Complexes
organoarsenic compounds 9.1: Metal Alkene Complexes
4.1: Organometallic Compounds of As(V) and Sb(V)
Dialkylmercury
5.2: Organometallic Compounds of Mercury
5.1: Organometallic Compounds of Zinc and Z
Cadmium
Zeigler Natta Polymerization
Organogermanium Compounds
F 3.5: Organosilicon and Organogermanium
3.4: Zeigler Natta Polymerization Catalysts

Fischer carbyne Compounds Zeise’s salt


Organosilicon Compounds 9.1: Metal Alkene Complexes
13.2: Metal-Carbynes
3.5: Organosilicon and Organogermanium
G Compounds
Oxidative addition
Grubb's catalyst
10.1: Oxidative Addition and Reductive Elimination
14.1: Catalytic Applications of Organometallic
Compounds- Alkene Metathesis

1 https://chem.libretexts.org/@go/page/212284
Index
A H R
Agostic alkyls hydroformylation reaction Reductive elimination
7.1: Transition Metal Alkyl Complexes 11.2: Homogeneous Catalysis - II 10.1: Oxidative Addition and Reductive Elimination
Hydrometallation
B 2.1: General Methods of Preparation S
Bridging hydrides Schrock carbynes
7.2: Metal Hydrides M 13.2: Metal-Carbynes
Mercury poisoning Schrock’s catalyst
C 5.2: Organometallic Compounds of Mercury 14.1: Catalytic Applications of Organometallic
Metal Carbonyls Compounds- Alkene Metathesis
Carbene insertion
6.2: Synthesis and Stability 8.1: Metal Carbonyls
concerted 3−centered associative metal hydrides T
mechanism 7.2: Metal Hydrides Transmetallation
2.1: General Methods of Preparation
10.1: Oxidative Addition and Reductive Elimination
O
D olefin polymerization U
Dewar−Chatt model 3.4: Zeigler Natta Polymerization Catalysts umpolung reactions
9.1: Metal Alkene Complexes
organoarsenic compounds 9.1: Metal Alkene Complexes
4.1: Organometallic Compounds of As(V) and Sb(V)
Dialkylmercury
5.2: Organometallic Compounds of Mercury
5.1: Organometallic Compounds of Zinc and Z
Cadmium
Zeigler Natta Polymerization
Organogermanium Compounds
F 3.5: Organosilicon and Organogermanium
3.4: Zeigler Natta Polymerization Catalysts

Fischer carbyne Compounds Zeise’s salt


Organosilicon Compounds 9.1: Metal Alkene Complexes
13.2: Metal-Carbynes
3.5: Organosilicon and Organogermanium
G Compounds
Oxidative addition
Grubb's catalyst
10.1: Oxidative Addition and Reductive Elimination
14.1: Catalytic Applications of Organometallic
Compounds- Alkene Metathesis
Glossary
Sample Word 1 | Sample Definition 1

1 https://chem.libretexts.org/@go/page/279560
Detailed Licensing
Overview
Title: Introduction to Organometallic Chemistry (Ghosh and Balakrishna)
Webpages: 61
Applicable Restrictions: Noncommercial
All licenses found:
CC BY-NC-SA 4.0: 93.4% (57 pages)
Undeclared: 6.6% (4 pages)

By Page
Introduction to Organometallic Chemistry (Ghosh and 4: Organoelement Compounds of Group 15 - CC BY-NC-
Balakrishna) - CC BY-NC-SA 4.0 SA 4.0
Front Matter - CC BY-NC-SA 4.0 4.1: Organometallic Compounds of As(V) and Sb(V)
TitlePage - CC BY-NC-SA 4.0 - CC BY-NC-SA 4.0
InfoPage - CC BY-NC-SA 4.0 4.2: Organometallic Compounds of As(III) and
Table of Contents - Undeclared Sb(III) - CC BY-NC-SA 4.0
Licensing - Undeclared 4.3: Phosphines - CC BY-NC-SA 4.0
1: Introduction - CC BY-NC-SA 4.0 5: Group 12 Elements - CC BY-NC-SA 4.0
1.1: Bonding Concepts in Main Group Chemistry - 5.1: Organometallic Compounds of Zinc and
CC BY-NC-SA 4.0 Cadmium - CC BY-NC-SA 4.0
1.2: VSEPR Theory and its Utility - CC BY-NC-SA 5.2: Organometallic Compounds of Mercury - CC BY-
4.0 NC-SA 4.0
2: Organometallic Chemistry of s- and p-block Elements 6: General Properties of Transition Metal Organometallic
- CC BY-NC-SA 4.0 Complexes - CC BY-NC-SA 4.0
2.1: General Methods of Preparation - CC BY-NC-SA 6.1: 18 Valence Electron Rule - CC BY-NC-SA 4.0
4.0 6.2: Synthesis and Stability - CC BY-NC-SA 4.0
2.2: Organometallic Compounds of Alkali Metals 7: Metal Alkyls and Metal Hydrides - CC BY-NC-SA 4.0
(Sodium and Lithium) - CC BY-NC-SA 4.0 7.1: Transition Metal Alkyl Complexes - CC BY-NC-
2.3: Organometallic Compounds of Alkaline Earth SA 4.0
Metals (Beryllium and Magnesium) - CC BY-NC-SA 7.2: Metal Hydrides - CC BY-NC-SA 4.0
4.0 8: Carbonyls and Phosphine Complexes - CC BY-NC-SA
2.4: Structure and Bonding - CC BY-NC-SA 4.0 4.0
3: Organometallic Chemistry of p-block Elements - CC
8.1: Metal Carbonyls - CC BY-NC-SA 4.0
BY-NC-SA 4.0
8.2: Metal Phosphines - CC BY-NC-SA 4.0
3.1: Reactions of Organometallic Compounds - CC
9: Complexes of π−bound Ligands - CC BY-NC-SA 4.0
BY-NC-SA 4.0
9.1: Metal Alkene Complexes - CC BY-NC-SA 4.0
3.2: Organometallic Compounds of Boron and
9.2: Metal Allyl and Diene Complexes - CC BY-NC-
Aluminium - CC BY-NC-SA 4.0
SA 4.0
3.3: Organometallic Compounds of Gallium and
9.3: Metal Cyclopentadienyl Complexes - CC BY-NC-
Indium - CC BY-NC-SA 4.0
SA 4.0
3.4: Zeigler Natta Polymerization Catalysts - CC BY-
NC-SA 4.0 10: Reaction Mechanisms - CC BY-NC-SA 4.0
3.5: Organosilicon and Organogermanium 10.1: Oxidative Addition and Reductive Elimination -
Compounds - CC BY-NC-SA 4.0 CC BY-NC-SA 4.0
3.6: Organotin and Organolead Compounds - CC BY- 10.2: Insertion and Elimination Reactions - CC BY-
NC-SA 4.0 NC-SA 4.0

1 https://chem.libretexts.org/@go/page/417156
10.3: Nucleophilic and Electrophilic Addition and 13.2: Metal-Carbynes - CC BY-NC-SA 4.0
Abstraction - CC BY-NC-SA 4.0 14: Metathesis - CC BY-NC-SA 4.0
11: Applications - CC BY-NC-SA 4.0 14.1: Catalytic Applications of Organometallic
11.1: Homogeneous Catalysis - I - CC BY-NC-SA 4.0 Compounds- Alkene Metathesis - CC BY-NC-SA 4.0
11.2: Homogeneous Catalysis - II - CC BY-NC-SA 4.0 14.2: Credits - CC BY-NC-SA 4.0
12: Physical Methods in Organometallic Chemistry - CC Back Matter - CC BY-NC-SA 4.0
BY-NC-SA 4.0 Index - CC BY-NC-SA 4.0
12.1: Characterization of Organometallic Complexes Index - Undeclared
- CC BY-NC-SA 4.0 Glossary - CC BY-NC-SA 4.0
13: Multiply-Bonded Ligands - CC BY-NC-SA 4.0 Detailed Licensing - Undeclared
13.1: Metal-Carbenes - CC BY-NC-SA 4.0

2 https://chem.libretexts.org/@go/page/417156

You might also like