Sunteți pe pagina 1din 181

Prof. Dr.

Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 1

Table of Contents

1. INTRODUCTION 4

Catalysis - Term and basic principles ................................................................................................... 4

Why catalysis? .................................................................................................................................... 7

Digression: The development of organotransition metal catalysis ................................................................... 9

Principles of catalysis ........................................................................................................................ 17


1.3.1. Energy considerations ........................................................................................................................ 17
1.3.2. Reaction profiles of catalytic reactions .............................................................................................. 18
1.3.3. Strategies to stabilize the transition state ......................................................................................... 20
1.3.4. Terminology ....................................................................................................................................... 21

2. STRUCTURE AND BONDING IN ORGANOMETALLIC COMPOUNDS ........ FEHLER!


TEXTMARKE NICHT DEFINIERT.

Metal ion .......................................................................................................................................... 26


1.4.1. Oxidation number .............................................................................................................................. 26
1.4.2. d-electron configuration .................................................................................................................... 26
1.4.3. Periodic trends ................................................................................................................................... 27
2.1.1.1. Ionization energies ..................................................................... Fehler! Textmarke nicht definiert.
2.1.1.2. Ionic radius ................................................................................. Fehler! Textmarke nicht definiert.
2.1.1.3. Binding energy............................................................................................................................... 29

Ligand characteristics ....................................................................................................................... 30


1.5.1. Ligand classifications: Dative/covalent, neutral/anionic, even/uneven number of electrons, L type/X
type 30
1.5.2. Ligand classifications: Number of electrons ...................................................................................... 30

Metal-ligand bond ............................................................................................................................ 31


1.6.1. Electron accounting ........................................................................................................................... 31
1.6.2. Structures ........................................................................................................................................... 33
1.6.3. Molecule orbital theory ..................................................................................................................... 34
1.6.4. The  bond in metal complexes ......................................................................................................... 35
2.1.1.4. CO and CO analogs ........................................................................................................................ 35
2.1.1.5. Carbene and carbyne complexes .................................................................................................. 37
2.1.1.6.  bond in alkene complexes .......................................................................................................... 39
1.6.5.  donor ligands .................................................................................................................................. 41

3. ELEMENTARY STEPS IN ORGANOTRANSITION METAL CATALYSIS................. 41

Dissociation and coordination of ligands .......................................................................................... 41

Oxidative addition and reductive elimination ................................................................................... 43

Oxidative coupling and reductive cleavage ....................................................................................... 45


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 2

Migratory insertion and -elimination.............................................................................................. 45

-hydrogen elimination and carbene insertion reactions ................................................................. 46

Addition of nucleophiles and heterolytic fragmentations ................................................................. 49

One-electron reduction and oxidation .............................................................................................. 50

4. HOMOGENEOUS HYDROGENATION ............................................................................ 51

Introduction ..................................................................................................................................... 51

Selected examples of hydrogenation catalysts ................................................................................. 54


4.2.1. Rhodium ............................................................................................................................................ 54
4.2.2. Iridium ............................................................................................................................................... 56
4.2.3. Ruthenium ......................................................................................................................................... 58

Mechanisms ..................................................................................................................................... 58
4.3.1. Typical mechanisms of alkene hydrogenation ................................................................................. 59
4.3.2. Oxidative addition as a key step in hydrogenation catalysis ........................................................... 64
4.3.2.1. Oxidative addition in Wilkinson's catalyst ................................................................................... 64
4.3.2.2. Vaska's complex............................................................................................................................ 65
4.3.2.3. Oxidative addition of alkyl halides RX to [Pt(PPh3)2] .................................................................. 66
4.3.2.4. Oxidative addition of CH bonds ................................................................................................. 67
4.3.2.5. Digression: σ complexes and agostic interactions ........................................................................ 71
4.3.3. Mechanism of the asymmetric alkene hydrogenation .................................................................... 84

5. CATALYTIC CARBONYLATIONS .................................................................................... 89

Introduction ..................................................................................................................................... 89

Rhodium-catalyzed carbonylation of methanol: Monsanto acetic acid process ................................ 89

Hydroformylation of alkenes ............................................................................................................ 91


5.3.1. Introduction ....................................................................................................................................... 91
5.3.2. Cobalt-catalyzed hydroformylation .................................................................................................. 92
5.3.3. Rhodium-catalyzed hydroformylation .............................................................................................. 97
5.3.4. Enantioselective hydroformylation ................................................................................................ 108

6. OXIDATIONS .....................................................................................................................114

The Wacker process ........................................................................................................................ 114


6.1.1. Mechanism of the Wacker oxidation .............................................................................................. 114

Epoxidations ................................................................................................................................... 117

Iron-catalyzed oxidations ............................................................................................................... 120


6.3.1. Nature as a model ........................................................................................................................... 120
6.3.2. Gif chemistry: Radical reactions ..................................................................................................... 121
6.3.3. Bioinspired oxidation catalysis in the lab ....................................................................................... 123
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 3

7. METATHESIS .....................................................................................................................131

Alkene metathesis .......................................................................................................................... 131


7.1.1. Introduction ..................................................................................................................................... 131
7.1.2. Historical development ................................................................................................................... 134
7.1.3. Catalysts........................................................................................................................................... 136
7.1.4. Mechanism ...................................................................................................................................... 141

8. OLEFIN OLIGOMERIZATION ........................................................................................149

The Ziegler reaction ........................................................................................................................ 150

Olefin Dimerization ........................................................................................................................ 151

Ethene trimerization ....................................................................................................................... 154

Ethene tetramerization .................................................................................................................. 155

Shell Higher Olefin Process (SHOP) ................................................................................................. 156

Acetylene Tetramerization ............................................................................................................. 159

Butadiene Di- and Trimerization ..................................................................................................... 159

9. POLYMERIZATION ..........................................................................................................163

Olefin Polymerization ..................................................................................................................... 164


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 4

Prof. Dr. Robert Wolf Universität Regensburg

Organotransition Metal Catalysis


Literature:
Dirk Steinborn, Grundlagen der metallorganischen Komplexkatalyse, Vieweg+Teubner,
Wiesbaden, 2010.
John F. Hartwig, Organotransition Metal Chemistry. From Bonding to Catalysis, University
Science Books, Mill Valley, California, 2010.
Arno Behr, Angewandte homogene Katalyse, Wiley-VCH, Weinheim, 2008.
Piet W.N.M. van Leeuwen, Homogeneous Catalysis - Understanding the Art, Kluwer
Academic Publisher, Dordrecht, Niederlande, 2004.

1. Introduction

Catalysis - Term and basic principles

Probably the oldest catalysts are metal-containing enzymes, which have existed in nature for
millions of years ( see lecture “Bioinorganic Chemistry”). The ancient Germanic tribes
already used catalysis to produce alcoholic beverages with the help of yeast cells, which they
then consumed with delight. Homogeneous biocatalysis thus was in the cradle of catalysis.
“Catalysis” as a scientific phenomenon was discovered in the 19th century. Already in the 2nd
half of the 18th century, an increasing number of reactions were described that marked the
beginnings of catalytic research.
The establishment of stoichiometry (J. B. Richter 1792/93) and the formulation of the laws
of constant and multiple proportions (J. L. Proust and J. Dalton, 1799 and 1803) provided an
important basis. It was thus possible to realize that substoichiometric quantities of a substance
can induce chemical reactions without the substance being consumed in the reaction.

The word “catalysis” (gr.  = dissolution) was coined by Jöns Jakob Berzelius
(1779-1848). In a report to the Royal Swedish Academy of Sciences, he summarized some of
the observations mentioned above:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 5

In accordance with the belief at that time, a reaction between two “bodies” required their
“chemical affinity”. Catalysts were therefore seen as “bodies” by Berzelius, which could
"awaken dormant affinities" by their "catalytic force" without themselves having "affinities"
to the reacting substances.
Berzelius uses the term catalysis in a purely descriptive way and brings together phenomena
that could not be explained by the doctrine that reactions are caused by chemical affinity.
Liebig attacked Berzelius several times for his definition of catalysis, with his criticism focusing
on the "creation of a new force through a new word that does not explain the phenomenon
either".
A prerequisite for understanding catalysis in more detail was the development of reaction
kinetics, which allowed to discover and analyze the laws of chemical reactions. This led to the
“kinetic definition of catalysis” by Wilhelm Ostwald (1853-1932; 1932 Nobel Prize in
chemistry). In his experiments on the acid-catalyzed oxidation of hydrogen iodide by bromic
acid, Ostwald detected that the nature of catalysis was not the inducement of a reaction, but
its acceleration.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 6

Catalysis is often not just the "reduction of the energy mountain" but allows for an alternative
reaction path that has a lower activation energy for the rate-determining step.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 7

Wilhelm Ostwald summarized his understanding of catalysis in his lecture on the occasion of
the awarding of the Nobel Prize in 1909. He stated: "…that until a way has been found whereby
a rate of chemical reaction can generally be calculated in advance ..., the catalysis problem
cannot satisfactorily be answered." Not until after Ostwald's death, Henry Eyring (1901-1981)
created an important theoretical basis with his transition-state theory (1935).

Why catalysis?

Economic significance: (Behr, 2008)


 Catalysis creates added value!
 80% of all chemical processes use a catalyst
 Value of the goods that are produced by means of a catalyst is > 400 billion € per year
 Current catalyst market: > 10 billion €
The possibility to create a desired product as selectively as possible by means of catalysts
makes catalysis a cornerstone of green chemistry.
Principles of green chemistry:
 Avoid waste  maximize the incorporation of all educts used in the process into the
final product ("atom economy")
 Energy use as low as possible
 Avoid unnecessary reaction steps (derivatizations, protective groups)
 Use renewable resources
The homogeneous transition metal catalysis in particular frequently satisfies these
requirements.
Example: Synthesis paths of ibuprofen in comparison
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 8

(Behr, 2008)
1st step: Acylation of isobutylbenzene, same for both processes
2nd step: Heterogeneous catalytic hydrogenation of the ketone group
3rd step: Homogeneous catalytic carbonylation of the hydroxy group with carbon monoxide
 After only three steps, ibuprofen is synthesized and all reagents used can be found in the
end product!

This allows, for example, for a direct C-C coupling of various molecules without the need for
any interim steps.
Important processes:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 9

Challenges in homogeneous catalysis:


 Cheap and available metals!
 Renewable resources
 "Dream reactions":
E.g.
 Photocatalytic water splitting
 Material use of CO2 (hydrogenation, incorporation into polymers)
 C-H functionalization, most suitably of methane
 Iron-catalyzed Fischer-Tropsch synthesis
 Hydrogenation of N2 to ammonia
 Catalytic functionalization of white phosphorus
 In all of these areas, organometallic catalysis can provide valuable contributions

Digression: The development of organotransition metal catalysis

After Döbereiner, Berzelius, Ostwald and others had detected the basic principles of catalysis,
catalysis developed rapidly as a scientific sub-discipline.
Catalysis is a key driver for investigating organometallic compounds because they are of great
importance as catalytically active species in industrial processes. The driving force for many
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 10

discoveries in catalysis research was "coincidence", fueled by the curiosity of chemists who
were pursuing other objectives.
Examples:
 Karl Ziegler set out to produce short-chained olefin oligomers and discovered low-
pressure polymerization of ethene to PE.
 Otto Roelen: Discovered aldehydes ("oxo products") in the Fischer-Tropsch reactor in
1938
Overall, four Nobel Prizes have been awarded so far to ten researches altogether for their
outstanding merit in homogeneous transition metal catalysis.
The development of catalysis can be roughly divided into four phases.
Phase 1 (1898-1918): Inorganic basic chemicals
 Sulfuric acid contact process (BASF, 1898)
 Ammonia synthesis (Fritz Haber, 1903, and Carl Bosch, 1908)
 Preparation of nitric acid by catalytic ammonia oxidation (Wilhelm Ostwald, post 1900)
Phase 2 (1919-1945): Synthesis gas and acetylene chemistry
Catalytic refinery processes and works on catalytic synthesis gas and acetylene chemistry
 Acetaldehyde as acetylene with HgSO4 as catalyst
 Acid-catalyzed synthesis of isopropanol from propene and water
 Production of methanol from synthesis gas (CO/H2, Zn/Cr cat., 1923, BASF)
 Steam reforming of methane to synthesis gas
 Catalytic cracking of hydrocarbons with solid acid catalysts
 Fischer-Tropsch synthesis: Reaction of synthesis gas, forming alkanes and alkenes with
heterogeneous iron and cobalt catalysts
 Oxosynthesis (= hydroformylation) with cobalt catalysts
 Reppe chemistry: Acetylene chemistry, nickel-catalyzed oligomerizations, reaction of
acetylene with CO and H2O or alcohols
Hydroformylation, which was discovered by Otto Roelen, and the C-C coupling reactions by
Reppe are the oldest homogeneous catalytic reactions involving transition metal complexes.
Phase 3 (1945-1970): Major petrochemical products
The great age of petrochemistry where homogeneous transition metal catalysis is
increasingly used - numerous developments in catalytic conversion of alkenes into
chemical base products
The homogeneous transition metal catalysis of 1938-1965:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 11

(Behr, 2008)
Phase 4 (since 1971): Fine chemicals and special products
The homogeneous transition metal catalysis of 1966-1985:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 12

Homogeneous transition metal catalysis since 1986:


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 13

(Behr, 2008)

Historical milestones of organotransition metal catalysis:


 1916, Wacker-Chemie (Burghausen):
Technical equipment for mercury-catalyzed hydration of acetylenes

 1917, Walter Reppe: Acetylene chemistry


“Reppe chemistry”
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 14

 1938, Otto Roelen (Ruhrchemie): Hydroformylation of olefins, “oxo synthesis”


Reaction of ethene with synthesis gas, forming propionaldehyde in the presence of a
heterogeneous cobalt-thorium catalyst; it was demonstrated at the end of the 1940s that
there is a homogeneous catalyst

 1953, Karl Ziegler (MPI Mülheim), and 1954/55, Gulio Natta (Institute of Technology,
Milan): Alkene polymerization
Karl Ziegler discovered low-pressure polymerization of ethene with organometallic mixed
catalysts at the MPI for Coal Research in Mülheim/Ruhr: Starting point for the development
of the organometallic chemistry of complex catalysis
Natta demonstrated that propene and other  olefins are stereoselectively polymerized with
Ziegler's catalysts

 1955, Günther Wilke (MPI Mülheim): Nickel-catalyzed cyclo-oligomerization, linear


oligomerization and telomerization reactions of butadiene

 1956, Phillips Petroleum: First technical synthesis of high cis 1.4 polybutadiene

 1956-1959, Jürgen Smidt, Wacker process: Catalytic process to produce acetaldehyde by


oxidation of ethene - of great importance since the development took place during the
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 15

time of the transition from acetylene chemistry (coal chemistry) to petrochemistry (ethene
chemistry)

 1965, Geoffrey Wilkinson (Imperial College): Homogeneous hydrogenation of olefins


using rhodium catalysts (previously only heterogeneously catalyzed processes, e.g. with
Ni, Paul Sabatier, Toulouse, 1912 Nobel Prize):

 1968, William S. Knowles (Monsanto Co., St. Louis, 2001 Nobel Prize) and L. Horner
(Universität Mainz): Enantioselective hydrogenation of prochiral olefins with Wilkinson-
type complexes with chiral phosphines, use of L-DOPA on an industrial scale

 1966, Nissim Calderon (Goodyear, Ohio): Homogeneously catalyzed olefin metathesis,


previously with heterogeneous catalysts, mechanism in 1971 by Y. Chauvin (2005 Nobel
Prize) and J.-L. Hérisson

 1966, Hitosi Nozaki and Ryori Noyori (Kyoto University, Japan): Asymmetric
cyclopropanation of styrene; first example of an asymmetric catalysis using a structurally
well-defined chiral transition metal complex

 1968, Monsanto process: Production of acetic acid by carbonylation of methanol; since


1970, the majority of acetic acid has been produced using this process
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 16

 1972, Heck reaction (R. F. Heck, University of Delaware): Palladium-catalyzed coupling of


aryl and vinyl halides with olefins

 1972-1979, metal-catalyzed cross-couplings: Couplings of organyl halides and


organometallic bonds

M. Kumada, 1972: M = Mg
E.-i. Negishi, 1976/77: M = Zn
A. Suzuki, 1979 (2001 Nobel Prize): M = B
J. K. Stille, 1979: M = Sn
 1977, Shell Higher Olefin Process (SHOP), Wilhelm Keim (TU Aachen): Nickel-catalyzed
ethene oligomerization in a liquid-liquid biphasic system

 1980, K. Barry Sharpless (Scripps Research Institute, La Jolla, USA, 2001 Nobel Prize):
Asymmetric epoxidation of alcohols

 Since 1980, Hansjörg Sinn, Walter Kaminski (Universität Hamburg), Hans-Herbert


Brintzinger (Universität Konstanz): Ethene polymerization with metallocene catalysts and
methylaluminoxane (MAO)

Use of ansa-metallocenes for stereoselective propene polymerization


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 17

 Since 1988, Richard R. Schrock (MIT Cambridge, 2005 Nobel Prize), Robert H. Grubbs
(Caltech, Pasadena, 2005 Nobel Prize): Alkylidene complexes of Mo, W and Ru with a
broad range of applications for olefin metathesis

 1997, Jean-Marie Basset (CNRS Lyon): Metathesis of alkanes with a tantalum hybrid
immobilized on a silica gel surface

Principles of catalysis

1.3.1. Energy considerations

As a reminder:
Basic principle of catalysis: Catalysts reduce the free enthalpy of the transition state with
the highest energy (i.e. which is rate-determining).
Catalysts do not lower the energy of reactants nor products! They therefore do not affect
thermodynamics (equilibrium constants, molar ratios of products and educts). However,
catalysts allow reactions to happen faster under mild conditions, which makes a higher
product yield feasible.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 18

Here is a simple example: Haber process

 exothermic reaction. The yield is limited by thermodynamics. It can be increased by


developing catalysts that allow reactions to be performed at a lower temperature.

1.3.2. Reaction profiles of catalytic reactions

A catalyst reduces the activation energy of a reaction as compared to an uncatalyzed reaction


by stabilizing the transition state. The catalyst does not only influence the transition state,
which represents an unstable, volatile species, but also changes the reaction profile of the
reaction by binding to one or more reactants and staying bound during the transition state.
The dissociation of the product either generates the catalyst or generates a species that will
be converted to the starting catalyst.
Reaction coordinate diagrams illustrate the interactions between catalyst and reactants -
three scenarios are possible:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 19

Scenario I: Uncatalyzed reaction substrate (S)  product (P)


Scenario II: Catalyst A stabilizes the transition state (TS) more than it stabilizes the ground
state by forming the complex SA  catalysis
Scenario III: Additive B stabilizes the ground state more than it stabilizes the TS by forming
the complex SB. This increases the activation energy as compared to scenario I
 no catalysis!
Scenario IV: Additive C forms the complexes SC or PC with both educt S and product P.
Product P may be released in an additional reaction (not shown), e.g. by adding
water, or by reacting with acid or oxidizing agents. Although the reaction is
quicker than in the absence of C, C works in this case as a stoichiometric additive
and not as a catalyst.
Example: Hydroamination of alkenes

Amine adds to the Pt(II)-alkene starting complex - formation of a stable


aminoalkyl intermediate, amine can be generated by reduction with NaBH4 
not catalytic
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 20

1.3.3. Strategies to stabilize the transition state

 Stabilization of a structure similar to the transition state of the uncatalyzed reaction by


interaction between substrate and catalyst (frequently observed in enzyme catalysis)
Example: Copper-catalyzed Diels-Alder reaction

or (very frequently in organometallic chemistry)


 completely new reaction path with a low barrier: Typically, the reaction takes place in
several reaction steps, all of which, however, have a lower energy barrier than the
activation energy of the uncatalyzed reaction
Example: Hydroboration catalyzed by electron-rich transition metals
 In the absence of the catalyst, addition of the B-H bond takes place in the concerted
transition state  increased temperatures are necessary, reaction with alkene as
solvent
 When using metal catalysts, the reaction takes place via oxidative addition of the B-H
bond to the metal ion, followed by coordination and migratory insertion of the alkene
into the M-H bond. This is concluded by reductive elimination and B-C bond formation
and regeneration of the starting catalyst.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 21

1.3.4. Terminology

Catalytic cycle: Series of steps during which the reactants are used up and the products are
formed; starting and finishing points of the reaction are the same.
Example of a catalytic cycle:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 22

Formation of the active catalyst LnM usually takes place from a precatalyst, in this case LnMX2.
The active catalyst complex binds the substrate while forming the catalyst-substrate adduct
LnMS. The bound substrate is converted to the intermediate LnMI and then to the product
LnMP. The dissociation of the product P leads to the regeneration of the catalyst complex,
coming full circle. The catalyst complex can leave the cycle reversibly by adding a ligand L or
by dimerizing the catalyst-substrate complex LnMS. The species that formed in the process,
Ln+1M and [LnMS]2, can be considered as resting state.
Catalyst complex: All species in the catalytic cycle
Precatalyst: Compounds from which active catalyst complexes can be generated; they are
generally so stable that they can be isolated in substance.
Examples:

 Formation of active hydrogenation catalysts by hydrogenation of COD (cycloocta-1.5-


diene) or COT (cycloocta-1,3,5,7-tetraene), e.g. [Rh(cod)L2]+ and [RuL2H(cot)]+
 Formation of active palladium(0) complexes by reducing [L2Pd(OAc)2] with nucleophiles
Cocatalyst: Component necessary for catalyst generation, which is not catalytically active
on its own
Promoter (activator): Additions to the catalyst that increase its effectiveness without being
catalytically active themselves
Initiator: Is not regenerated in the catalytic cycle, but is consumed irreversibly at the start
of the reaction
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 23

Example: Radical polymerization of styrene

Deactivation: Irreversible decomposition of the catalyst by a step outside of the catalytic


cycle (see slide above), e.g. P-C bond cleavage or cyclometalation of phosphine ligands,
irreversible dissociation of ligands, formation of metal particles, ...
Catalyst deactivation is an important phenomenon: It is the topic of entire conferences, books,
etc.
Resting state: Complex in considerably higher concentration than all other complexes in the
reaction mixture, which is connected to the active catalyst complex via a balance
Example: Rhodium catalyzed hydroformylation of alkenes
O R
PPh3
O CO PPh3
OC Rh OC Rh
R PPh3
PPh3 CO
active catalyst resting state
complex (reservoir for active catalyst)

Reversible reactions outside of the catalytic cycle may slow down catalytic reactions by
reducing the catalyst concentration
Rate-determining step: Reaction step with the highest activation energy; if there is complex
reaction kinetics, there is not necessarily a rate-determining step, although there is in most
cases.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 24

Solution:
−∆𝐺‡
𝑘𝑏 ∙𝑇
Eyring equation: 𝑘 = 𝑒 𝑅𝑇

Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 25


−∆𝐺1
𝑘1 𝑒 𝑅𝑇 𝑘 −∆𝐺1‡ −∆𝐺2‡ 𝑘
= ‡ , logarithmizing: 𝑙𝑛
1
= − , transforming 𝑅𝑇𝑙𝑛 𝑘1 = ∆𝐺2‡ − ∆𝐺1‡ = ∆∆𝐺 ‡
𝑘2 −∆𝐺 2 𝑘 2 𝑅𝑇 𝑅𝑇 2
𝑒 𝑅𝑇

𝑅 = 8.314 𝐽 ∙ 𝐾 ∙ 𝑚𝑜𝑙 −1 , 𝑇 = 298 𝐾


𝑘1
= 10 → ∆∆𝐺 ‡ = 5.70 𝑘𝐽 ∙ 𝑚𝑜𝑙 −1
𝑘2

𝑘1
= 100 → ∆∆𝐺 ‡ = 11.41 𝑘𝐽 ∙ 𝑚𝑜𝑙 −1
𝑘2

𝑘1
= 1000 → ∆∆𝐺 ‡ = 17.11 𝑘𝐽 ∙ 𝑚𝑜𝑙 −1
𝑘2

Conclusion: Even a comparably small decrease of the activation barrier (compare rotation
barrier around C-C bonds of ethane: 11-13 kJ/mol) causes a high reaction acceleration!
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 26

2. Structure and bonding in organometallic compounds

Metal ion

2.1.1. Oxidation number

 Formal charge
 Results from the formal charge of the ligands and the overall charge of the complex
Examples:
 Neutral complex with two mono-anionic ligands  oxidation number +II for the metal
atom
 Monocationic complex with two mono-anionic ligands  oxidation number +III for the
metal ion
Important: The oxidation number is just a formal charge that has no direct reference to
physical reality! Due to the covalent bond components of the metal-ligand bond (in particular
with organometallic compounds), the actual charge of the metal ion is between 1 (= L.
Pauling's principle of electroneutrality).
Why are oxidation numbers useful nevertheless?
 Useful to determine the valence electron number of the metal ion (i.e. the number of
electrons in orbitals with a high metal content)
 In combination with the valence electron number, the oxidation number allows to make
predictions about the structure and reactivity of complexes

2.1.2. d-electron configuration

Simple rule:
 (n)d shell is filled first, (n+1)s shell remains empty
Cause: Different shielding of s and d orbitals. The 4s orbital is less shielded in the gas phase
than a 3d orbital; therefore, the 4s orbital has a lower energy in the gas phase than the
3d orbital. In a metal complex, in which the metal ion has a partially positive charge, the
difference of the shielding is lower (the orbitals are contracted more in the positively
charged ion).
Number of d-electrons: Number of d-electrons not involved in metal-ligand bonds. For neutral
complexes, the number of d-electrons corresponds to the group number
Examples - jointly compile on the board:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 27

Oxidation d-electron configuration


number
Fe(CO)5 0 d8
[Fe(CO)4]2- -II d10
Cp2Fe: +II d6
[ReH9]2- +VII d0

d-electron configuration allows to predict whether the complex will be diamagnetic or


paramagnetic:
Even number of electrons  usually diamagnetic
Uneven number of electrons  paramagnetic

2.1.3. Periodic trends

Change in the stability, basicity and d orbital energies within the transition metals  influence
on the reactivity!
 Often opposing trends to the main group elements
 No continuous trends, but areas with similar characteristics within the transition metals
(e.g. valence electron poor vs. valence electron rich transition metals, first-row vs. second-
and third-row metal, iron group vs. platinum group)

2.1.3.1. Ionization energies

 Closely connected to the energies of the d orbitals!


Trends of the ionization energy:
 Increase from left to right
Question: Reasons for this trend?
Answer: The more electronegative an element, the higher the effective nuclear charge and
the lower the ionization energy
 For two metal atoms with the same oxidation number and similar ligands, the metal that
is poorer in valence electrons is usually easier to oxidize, is more basic and richer in electrons
Ex.: Comparison between Zr(II) and Pd(II)
Zr(II): Easier to oxidize, more basic, richer in electrons
 First and second ionization energies of a group - no general trend: For some groups, the
oxidation of the second-row metal requires more energy than the oxidation of the first-
row metal, and the oxidation of the third-row metal more energy than the oxidation of
the second-row metal (Ex. 1. IE: groups 4 and 7-9); for some groups, the trend is reversed
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 28

 Stability of higher oxidation states increases within a group,


e.g.:
 Pt(IV) compounds are often very stable, Ni(IV) compounds are very reactive (stable
only as fluoride and oxide),
 Os(VIII) compounds well known, Fe(VIII) unknown,
 Ir(V) normal oxidation state, Co(V) unknown
 Basicity generally increases within a group, i.e. 5d metals are more basic than 2nd row
metals, which are more basic than 3d metals
Question: How can basicity be determined?
Answer: pKS values of the transition metal hybrid!
E.g. pKS values: Os(CO)4(H)2 < Ru(CO)4(H)2 < Fe(CO)4H2
CpW(CO)3H < CpMo(CO)3H < CpCr(CO)3H

2.1.3.2. Ionic radius

Size trends:
 3d metals < 4d metals  third-row metals
Question: Why does this trend exist?
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 29

Answer: Lanthanide contraction!, effective nuclear charge of the 5d metals is considerably


higher because 14 lanthanides are incorporated beforehand
 For low-spin complexes of the groups 4-7 with the same charge and oxidation number:
Size decreases from left to right; reason: bigger effective nuclear charge of the elements
on the right side

2.1.3.3. Binding energy

Trend of the binding energies:


 3d metals < 4d metals < 5d metals
 Main group elements: Opposing trend (e.g. CF > CCl > CBr > CI)
The reasons are complex due to the large degree of variation in the characteristics of the
transition metals, both covalent and ionic factors:
 Better spatial overlapping between metal and ligand orbitals
 Similar energy of ligand orbitals and orbitals of the 5d metals in comparison to first-
row and 4d metals (see “Ionization energies”, chapter 2.1.3.1.)
 Higher covalent bond content with heavier metals via better orbital overlapping and lower
energy difference
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 30

Ligand characteristics

2.2.1. Ligand classifications: Dative/covalent, neutral/anionic, even/uneven number of


electrons, L type/X type

Fundamental classification based on assigning the charges to the ligand or to the oxidation
number of the metal ion:
“Covalent” ligands: “Dative” ligands:
“Charged” ligands “Neutral” ligands
Bonding takes place by metal and ligand Bonding takes place by the ligand donating
sharing an electron each two electrons to the metal ion
If the electrons from the bond are assigned
If the electrons from the bond are assigned to the ligand  “neutral” ligand
to the ligand  “anionic” ligand

“X type” ligands: “L type” ligands:


They donate one electron to the metal ion They donate two electrons to the metal
ion
Distribute the electrons of the metal-ligand bond in such a way that neutral ligands are
formed as fragment

X type ligands frequently bridge two metal atoms.


Bridging halide, alkoxide, thiolate and amide ligands: 1 x X type and 1 x L type

Bridging hydride ligands: Three-center two-electron bond

2.2.2. Ligand classifications: Number of electrons

Neutral or ionic model and counting method:


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 31

 ligands:
Ligands with an even number of C atoms  neutral, donate an even number of electrons
when using the neutral counting method
Ligands with an uneven number of C atoms  charged, donate an uneven number of
electrons when using the neutral counting method
The number of electrons donated by a cyclic  ligand using the ionic counting method
corresponds to the number of electrons in the most stable aromatic system.
Examples:

Metal-ligand bond

2.3.1. Electron accounting

The valence electron number is physically measurable, in contrast to the rather arbitrary
formalisms "oxidation number" and "ligand charge" discussed in chapters 2.1.2. and 2.2.1.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 32

This number is important to understand and predict the characteristics and the reaction
behavior of transition organometallic compounds.
Two systems of counting electrons:
1) Covalent model of electron accounting
Neutral central atom and neutral ligands
Valence electron number = group number of the metal atom + number of X ligands + 2x
number of L ligands  Overall charge of the complex
2) Ionic model
Charged central ion and charged ligands
Valence electron number = number of d-electrons of the metal ion + 2x number of X ligands
+ 2x number of L ligands
Relationship between charge, oxidation number and ligand charge in the ionic model:
Complex charge = oxidation number + overall charge of the ligands
Number of d-electrons = group number - oxidation number
Examples:
[CpFe(CO)2(C2H5)]
a) Covalent model: b) Ionic model:
1. Iron Group 8  8e- 1. Ligands
2. CO = L type ligand: 2e- per CO  2 x CO: Neutral, 2 x 2e- = 4e-
3. Cp: 5e-  Cp-: Anion, 6e-
 C2H5: Anion, 2e-
4. C2H5 (ethyl): 1e-  electron number of the ligands = 12e-
Overall valence electron number = 2. Metal ion
8 + (2 x 2) + 5 + 1 = 18e-
Oxidation number: Neutral complex, two
anionic ligands  Fe2+
3. Number of d-electrons for Fe(II): 6e-
4. Overall valence electron number =
12 + 6 = 18e-
[CpMo(CO)3(H2O)]+
a) Covalent model: b) Ionic model:
1. Molybdenum: Group 6  6e- 1. 3x CO: 6e
2. CO: 3x 2e- 2. H2O: 2e
3. H2O: 2e- 3. Cp-: 6e-
4. Cp: 5e-  overall: 14e-
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 33

5. Positive charge: -1 4. Oxidation number: Cationic complex, an


Overall valence electron number = anionic ligand  Mo2+

6 + (3 x 2) + 2 + 5  1 = 18e- 5. Number of d-electrons: Group 6, Mo(II) 


d4
6. Overall valence electron number =
14 + 4 = 18 e-
[Ru(NO)(PPh3)2Cl3]
a) Covalent model: b) Ionic model:
1. Ruthenium: Group 8  8e- 1. 2x PPh3: 2x 2e-
2. 2x PPh3: 2x 2e- 2. NO+: 2e-
3. NO (linear): 3e- 3. 3x Cl-: 3x 2e-
4. 3x Cl: 3e-  overall: 12e-
Overall valence electron number = 4. Oxidation number: Neutral complex, a
8 + (2 x 2) + 3 + 3 = 18e- cationic and 3 anionic ligands  Ru2+, d6
5. Overall valence electron number= 12 + 6 =
18e-
14 + 4 = 18 e-
[Pt(PEt3)2(nBu)2]
a) Covalent model: b) Ionic model:
1. Platinum: Group 10  10e- 1. 2x PEt3: 2x 2e-
2. 2x PEt3: 2x 2e- 2. 2x nBu: 2x 2e-
3. 2x nBu: 2x 1e-  overall: 8e-
Overall valence electron number: 10 + (2 x 2) 3. Oxidation number: Neutral complex, two
+ (2 x 1) = 16e- anionic ligands  Pt2+, d8
4. Overall number of electrons = 8 + 8 = 16e-

2.3.2. Structures

IUPAC definition: Coordination number as number of the atoms that are directly bound to the
metal atom
Convention in organometallic chemistry: Coordination number corresponds to the number of
donated electron pairs so that the coordination number in the above mentioned complexes
[Pt(PEt3)2(nBu)2], for example, equals four, in [CpFe(CO)2(C2H5)] and [Ru(NO)(PPh3)2Cl3] equals
six each and in [CpMo(CO)3(H2O)]+ seven.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 34

Electronic effects (overall energy of the occupied orbitals, in particular of the HOMO) can
stabilize a sterically unfavorable structure. Impact of certain d-electron configurations: Four-
fold coordinated complexes of the 2nd and 3rd transition metal series are nearly always
square planar  eight d-electrons fill low-lying MOs
Most organometallic complexes have strong donor ligands  usually strong crystal-field
splitting, low-spin complexes

2.3.3. Molecule orbital theory

Construction of an MO diagram:
 Symmetry-adapted linear combinations (SALCs) of metal and ligand orbitals
 MOs are formed by combining metal and ligand orbitals
Example: Octahedral complex [PtH6]2-
 Ligand orbitals: 6 SALCs, eg, t1u, and a1g
 Metal orbitals: t1u for (n+1) p orbitals, a1g for (n+1)s orbital, t2g for three d orbitals and eg
for two additional d orbitals  no SALC for three d orbitals (t2g symmetry)
 Metal and ligand orbitals of the same symmetry type form binding MOs with lower
energy than the isolated metal and ligand orbitals as well as anti-binding MOs, which
have a higher energy than the isolated orbitals.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 35

• Non-bonding and anti-bonding MOs with high d character often control the reaction
behavior  for the octahedral complex, usually t2g and 2eg orbitals. These are usually
considered to be metal d orbitals, which is not really true since they are MOs (with a high
d percentage). Due to the higher energy, the d percentage for 2eg is higher than for t2g
because the energies of the d orbitals are usually higher than those of the ligand orbitals.
 18-valence electron rule: 18 electrons fill the 1a1g, eg, t1u and t2g; t2g is the HOMO, 2eg is
the LUMO of the complex. Background of the 18-valence electron rule: A complex of any
symmetry can accept exactly 18 electrons because 9 valence orbitals of the metal are
available, which can form at the most 9 binding and 9 non-binding MOs. When accepting
more than 18 electrons, they must be accepted by anti-binding MOs  energetically
unfavorable

2.3.4. The  bond in metal complexes

2.3.4.1. CO and CO analogs

The  acceptor interaction is extremely important to stabilize complexes in low formal


oxidation states.
Question: Which are the strongest  acceptor ligands?
Answer: Carbon monoxide, NO+, isocyanides
For the  acceptor interaction, unoccupied * orbitals of ligands often play a big role. To put
it differently: Occupied MOs have * orbital parts of the ligand (e.g. of CO), which results in a
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 36

delocalization of electron density of the ligand. Delocalization leads to a decrease of the C-O
bond order and compensates the σ donation of the electron density  “backbonding”.
Backbonding is physically noticeable by a longer C-O bond length and a lower C-O vibrational
frequency.

Digression: Tolman electronic parameter (TEP)


The interaction between trialkyl phosphine ligands PR3 with transition metal atoms can be
interpreted as Lewis-acid-base interaction between the diffuse free electron pair of the
phosphorus and the Lewis-acidic metal atom. Tertiary phosphines and, in particular,
phosphites, however, also have  acceptor ability due to the interaction between occupied
metal d orbitals and P-X-σ* orbitals of the ligands.

A classical approach in order to classify the electronic influence of phosphine ligands is to


determine the Tolman electronic parameter: Value of the highest C-O bond in the IR
spectrum of [Ni(CO)3L].
The value of CO is dependent on the the strength of the metal-ligand backbonding. It, in turn,
depends on the electron density at the metal center and thus on the donor strength of the
phosphine.
Examples:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 37

L CO / cm-1 L CO / cm-1


PtBu3 2056 PPh3 2069
PCy3 2056 P(OMe)3 2079
PMe3 2064 P(OPh)3 2085
P(C6H4-4-Me)3 2066 PF3 2110
 PtBu3 and PCy3 are the strongest donors, PF3 is the weakest donor in this selection.

2.3.4.2. Carbene and carbyne complexes

Examples:

Despite this convention of counting electrons, the M-C bond differs only slightly. It is, of
course, by no means the case that the negative charge sits on the carbene ligand for
tantalum complexes.
From the perspective of the metal fragment, the symmetry of the border orbitals of carbene
and carbyne ligands are similar:
 Carbene ligand: One σ and one  orbital
 Carbyne ligand: One σ and two  orbital
Important distinction between Fischer and Schrock carbene or carbyne complexes:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 38

Carbyne complexes, by analogy:

Fischer carbenes (e.g. C(OMe)Me, C(NMe2)Me and CPh2): Neutral ligands, two-electron
donors in the covalent and ionic models (see table in section 2.2.1.)
Schrock carbenes: Dianionic ligands, two-electron donors in the covalent model, four-
electron donors in the ionic model
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 39

Solutions:

2.3.4.3.  bond in alkene complexes

Dewar-Chatt-Duncanson model, the interaction between σ donor and  acceptor again plays
an important role:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 40

While for CO the  acceptor ability is predominant over σ donor characteristics, the relative
σ donor and  acceptor strength strongly depends on the metal ion for alkenes:

Metal atoms in high oxidation states (frequently in cationic complexes): Alkenes act as
excellent σ donors  resonance structure A
 Similar C-C bond length as in the free alkene
 Complexes often react as nucleophiles
Electron-rich metal atoms in formal low oxidation states: Alkenes act as excellent 
acceptors  resonance structure B
 Electron density from the metal to the ligand
 Significantly longer C-C bond length than in the free alkene
 Hybridization increases from sp2 to sp3  angular structure
Examples:

K[PtCl3(2-C2H4)]H2O: Zeise's salt, similar bond length as in the free ethene (1.339 Å) 
structure A
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 41

[Os(2-C2H4)(CO)4]: Similar bond length as in cyclopropane (1.512 Å)  structure B


Iron complex: σ donor quality is predominant (for moderately electron-rich and electron-poor
metals)  structure A
Zirconium and ruthenium complex: Strong backbonding, metal cyclopropene structure B

2.3.5.  donor ligands

Coordinated polyenes and arenes act in metal complexes as σ donor and  donor ligands.
Example: Bis(benzene)chromium
Exercise in the lecture: Draw the  orbitals of benzene

MOs of Cr(C6H6)2:
 SALCs of the benzene ligands: 2x σ (2x a2u  a1g + a2u), 2x  (e1g  e1g + e1u), 2x  (e2u 
e2g + e2u), 2x b2g  b2g + b2u
 Metal orbitals: s, dz2  a1g, pz  a2u, px,py  e1u, dxz,dyz  e1g, dx2-y2,dxy  e2g
 MOs by overlapping with the following metal orbitals:
 a1g + s orbital, dz2 orbitals  σ symmetry
 a2u + pz  σ symmetry
 e1u + px,py; e1g + dxz,dyz   symmetry
 e2g + dxy, dx2-y2   symmetry
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 42

Many other cyclic and acyclic polyenes can also have σ,  and  interactions with the metal
orbitals. The question whether these are “donor” or “acceptor” interactions cannot be
answered clearly beforehand but is dependent on the formal charge allocated to the ligands
and the occupation of the orbitals in the free ligand.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 43

Some simple ligands can also act as  donors, for example:


Alkynes:

 Alkynes act as four-electron ligands in complexes with less than 18 electrons (see table in
section 2.1.2)
Alkoxides and amides:

 Alkoxide ligands: Linear M-O-C angle implies sp hybridization  two electron pairs form
a  bond with the metal atom
 Amide ligands: Planarity implies sp2 hybridization at the nitrogen  free electron pair
can form a  bond with the metal atom
Other good  donor ligands: Halide, oxide and nitride ions
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 41

3.  donor ligands

3. Elementary steps in organotransition metal catalysis

3.1. Dissociation and coordination of ligands

Ligand substitution is an important step in the synthesis of many organometallic bonds. Ligand
substitutions consist of both a ligand dissociation and a ligand addition.

Various mechanisms:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 42

The ligand dissociation is the first step in many catalytic reactions (see chapter 1.4.4.) - “a
site of coordinative unsaturation is perhaps the most important property of a homogeneous
catalyst” (J. P. Collman, Acc. Chem. Res. 1968, 1, 136).
Associative mechanism, A:
Species formed in the first, associative step is in organometallic reactions in most cases an
intermediate; for SN2 reactions, however, it is a transition state  A parallel can be drawn
with base-catalyzed transesterifications, where an intermediate can actually be observed.

Associative and dissociative mechanisms are extremes in a continuum. The truth is often
somewhere in between  interchange mechanisms where both a partial bond break and a
partial bond formation take place in the transition state. When the bond formation is more
pronounced in the rate-determining state, it is referred to as associative interchange IA; when
the bond break is more pronounced, it is called dissociative interchange ID.
Exchange mechanisms are particularly widespread in coordination chemistry when charged
complexes react with ligands of opposite charge  Formation of an encounter complex where
complex and ligand form an intermediary ion pair.

The associative interchange mechanism does not play a major role in organometallic
reactions, because they usually take place with neutral complexes, non-polar organic solvents
and neutral ligands such as alkenes and CO. Nevertheless, this option needs to be kept in mind.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 43

Dissociative interchange mechanisms are quite frequently observed in organometallic


chemistry.

3.2. Oxidative addition and reductive elimination

Oxidative addition: A substrate molecule X-Y is added to a low-valent metal complex after a
bond break occurred  increase of the coordination number, the oxidation state and the
valence electron number by two units
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 44

Prerequisites for oxidative additions:


 metal atom is coordinatively and electronically unsaturated
 metal atom has non-bonding electrons
 metal atom is able to make a transition to an oxidation state that is higher by two units
Oxidative additions are frequently encountered with d8 and d10 complexes where a transition
occurs from a square planar to an octahedral coordination geometry. Example:
MII  MIV (M = Pd, Pt), MI  MIII (M = Rh, Ir), M0  MII (M = Pd, Pt)
Typical substrates - 3 classes:
 non-polar or weakly polar, e.g. H-H, X-X (X = Cl, Br, I), C-H and Si-H
currently “hot”: C-H additions! See e.g. S. R. Neufeldt, M. S. Sanford, "Controlling Site
selectivity in Palladium-Catalyzed C–H Bond Functionalization" Acc. Chem. Res. 2012, 45,
936-946.
 polar, e.g. H-X and R-X (R = alkyl)
 somewhat polar, e.g. R-X (R = aryl, X = Cl, Br, I)

Qualitative trends for oxidative addition:


Oxidative addition is preferable at
 electron-rich metal centers because most oxidative additions result in an increase of the
partial charge of the metal atom
 sterically unhindered metal centers because the coordination number increases in
oxidative additions
 oxidative addition of non-polar substrates require coordinative unsaturation and a
valence electron number of 16; for polar substrates, a 16 valence electron metal atom is
not required when a cationic metal center is formed with an uncoordinated anion
(“oxidative ligation”)
 The rate and equilibrium constants for the ligand association or -dissociation also have
an influence on the rate of the oxidative addition.

Oxidative additions have a rich mechanistic variety: radical, ionic and concerted
(synchronous) mechanisms, also SN2-type mechanisms are known – more information will
be given below in section 4.3.2. dealing with transition metal-catalyzed hydrogenation
reactions
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 45

3.3. Oxidative coupling and reductive cleavage

Oxidative coupling:
 Alkenes or alkynes form metallacycles following  complex formation (C-C bond linkage)
 Metal makes the transition to an oxidation state that is higher by two units
 Reversal: Reductive cleavage or fragmentation
3.4. Migratory insertion and -elimination

Key steps in complex catalysis, e.g. for hydrogenation, hydroformylation etc.


Migratory insertion and -elimination can be seen not only with metal hydrides but also with
CO, alkynes, carbenes and other ligands → also referred to as hydro- and carbometallation
Two steps for the insertion:
1. Olefin coordination: Coordination number and valence electron number increase
2. Insertion step: Coordination number and valence electron number decrease
The oxidation state of the metal atom does not change!
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 46

3.5. -hydrogen elimination and carbene insertion reactions

The -hydrogen elimination generates either a carbene and a hydrido ligand from an alkyl
ligand, or a carbene and an alkane from two alkyl ligands.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 47
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 48

Factors encouraging -hydrogen elimination:


 -hydrogen elimination is typically slower than -elimination  usually with benzyl,
neopentyl and methyl complexes that do not have any -hydrogen atoms
 Frequently d0 complexes of the groups 4 and 6  access to high oxidation states, which
are important for forming alkylidene hydride; in the absence of d-electrons, there is a
weaker tendency to form alkene complexes due to competing -H elimination
 sterical hindrance of the complexes
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 49

3.6. Addition of nucleophiles and heterolytic fragmentations

By coordination to a metal atom, an olefin, for example, is activated so that it can react upon
attack by a nucleophile, producing an alkyl complex. The reverse reaction is called
heterolytic fragmentation.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 50

3.7. One-electron reduction and oxidation

Oxidized and reduced metal complex merely differ by one electron. The electron injected
upon reduction may populate a metal-centered or a ligand-centered molecular orbital. In case
a metal-centered orbital is populated, the oxidation state of the metal atom decreases by one
unit. When a ligand-centered orbital is involved, the ligand is oxidized or reduced.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 51

4. Homogeneous hydrogenation

4.1. Introduction

 One of the most important reactions for the synthesis of bulk and fine chemicals
 Heterogeneous reactions (palladium and platinum as catalysts) are advantageous, but
homogeneous reactions are more selective  in particular stereoselectivity
 Diverse functional groups: Alkynes, aldehydes, ketones, esters, nitroarenes, arenes and
heteroarenes
 Multitude of catalysts, in particular neutral and cationic rhodium, ruthenium and iridium
complexes
2001 Nobel Prizes:

W. S. Knowles: Industrial synthesis of the rare amino acid L-DOPA, which can be used to treat
Parkinson's disease. In the key step, a cationic rhodium complex of the chiral diphosphine
DIPAMP catalyzes enantioselectively the hydrogenation of the enamide A (quantitative yield
and 95% ee). A simple acid-catalyzed hydrolysis step leads to L-DOPA.
The Monsanto process was the first commercial catalytic asymmetric synthesis with a chiral
metal complex.
Noyori and Takaya discovered the atropisomeric, chiral diphosphine BINAP in 1981.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 52

BINAP is considerably effective in a great variety of asymmetric reactions:


 Enantioselective hydrogenation of -(acylamino)acrylic acids and/or esters to amino acids
 Enantioselective isomerization of allylamines to enamines
 Hydrogenation of ,- and , -saturated amino acids
Example: Synthesis of (S)-naproxen (pain-relieving, fever-reducing and anti-inflammatory)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 53

Revolutionary: Discovery of “Noyori-type” complexes that selectively hydrogenate carbonyl


groups of unsaturated ketones (switch of selectivity)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 54

4.2. Selected examples of hydrogenation catalysts

4.2.1. Rhodium

Prototype of the hydrogenation catalyst: Wilkinson's catalyst


Synthesis:

Reactivity of Wilkinson's catalyst


 Rapid hydrogenation of non-conjugated alkenes and alkynes (1 bar H2, room
temperature); important: polar solvents such as ethanol facilitate migratory insertion
and thus accelerate the catalysis
 Relative rates for different alkenes cover a range of about 50-fold
 Cis addition of H2
 No isotope scrambling between H2 and D2 or between D2 and substrate or solvent
 Selective alkene hydrogenation in the presence of ester, keto and nitro groups
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 55

Cationic rhodium complexes [RhL2(H)2(S)2]+ (S = polar solvent molecule, e.g. THF or MeOH,
L2 = tertiary phosphine molecule or more commonly bidentate phosphine ligand)
 important for the development of enantioselective hydrogenation of alkenes!
 Usually more selective than neutral Wilkinson-type ligands - little alkene isomerization

Initial rate constant kinit for hydrogenation of 1-hexene by cationic rhodium complexes
with phosphine ligands (nbd = norbornadiene)
Complex Substrate concentration / kinit × 104 /s-1
mM
[Rh(nbd)(PPh3)2]+ 5.3 0.1
[Rh(nbd)(PPh2Me)2]+ 3.7 3.0
[Rh(nbd)(PPhMe2)2]+ 3.5 6.0

 ligands with stronger electron donation generate more active complexes


 catalysts with bidentate ligands are more active than with monodentate ligands;
particularly active: [Rh(nbd)(dppb)]+ (dppb = 1,4-bis(diphenylphosphino)butane)
 hydrogenation of the coordinated alkene forms the active catalyst
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 56

 The reaction with H2 forms the cationic dihydride [RhL2(H)2S2]+ with acidic hydride
ligands  adding a base such as triethylamine leads to neutral Wilkinson-type ligands

4.2.2. Iridium

“Crabtree’s catalyst”: [Ir(cod)(py)(PCy3)]+ (cod = 1,5-cyclooctadiene)


This is actually a precatalyst; the actual catalyst is created by hydrogenation of the COD
ligand with formation of a cationic iridium(III) hydrido complex
 Crabtree’s catalyst is not sensitive to air, unlike most hydrogenation catalysts!
 binds polar groups (cation!); non-polar solvents are therefore required, e.g. CH2Cl2
 weakly coordinating anions such as [B(ArF)4]- (ArF = C6H3-3,5-(CF3)2)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 57

 Crabtree's catalyst is more reactive than neutral, sterically more hindered rhodium and
ruthenium catalysts
 Very high activity with tetra-substituted alkenes
 Dramatic influence of the solvent (entries 2 and 3)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 58

4.2.3. Ruthenium

 As old as Rh-catalyzed hydrogenations


 A wide range of applications today, especially with asymmetric hydrogenations etc.
 First catalyst: [Ru(PPh3)3(H)Cl] (Wilkinson, 1965), very active for terminal alkenes, 1000
times slower for internal alkenes

 Commercial use of Ru-catalysts with chiral ligands, e.g. [Ru(BINAP)(OAc)2)] and


[Ru(BINAP)Cl2]2, for enantioselective hydrogenation of alkenes and ketones

 Clean, simple, economical


 100 mg to 100 kg range, high substrate concentration (up to 50%)
 Up to 100% enantiomerically pure products  significantly superior to enzyme reactions
 Both enantiomers can be synthesized effectively
 Wide range of industrial applications

4.3. Mechanisms

Countless studies (e.g. the legendary studies by Halpern et al. on the mechanism of Rh-
catalyzed hydrogenations) - two very important conclusions:
1. The spectroscopically observable species during catalytic reactions are frequently not
located in the catalytic cycle (unproductive equilibria, off-cycle intermediate or irreversibly
formed, unproductive side products)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 59

2. Determine rate laws for catalytic sub-steps independently whenever possible and correlate
with the rate law for the overall reaction! Studies examining the influence of countless
variables on the kinetics of the overall reaction are misleading! (too many variables)

4.3.1. Typical mechanisms of alkene hydrogenation


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 60

Similar mechanisms can be observed with ruthenium and iridium complexes, whereby
modifications may occur, e.g. concerted oxidative addition and migratory insertion or
heterolytic cleavage of H2 with ruthenium acetates.
Another variation is σ-bond metathesis:

Hydrogenation of ketones and imines often does not happen via a reaction path that includes
oxidative addition, migratory insertion and reductive elimination. Instead, an outer-sphere
mechanism can often be observed in such cases, i.e. metal-ligand bifunctional catalysis
(Noyori)
Classical: Chiral, non-racemic diamine complexes [Ru(BINAP)(diamine)(H)2]
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 61

 Initially, formation of H2 complexes as short-lived intermediate


 Heterolytic H-H bond cleavage by simultaneous transfer of a proton and an H ion to the
ketone or imine with formation of a dihydride
 Simultaneous transfer of a hydride ion and proton via a six-membered transition state -
Indications:
o Kinetic isotope effect for hydride and N-H positions.
o Overall isotope effect is approximately the sum of the individual isotope effects.
o Complexes without N-H function do not show catalytic activity.
 Nature of the product of the hydride transfer
o Recent results: Direct formation of the alcohol that is still bound to the amine ligand
via H-bridge, or direct formation of the alkoxide complex
o No dissociation of the free alcohol
Elimination of the alcohol takes place by protonation with an acid  reminiscent of the
dissociative ligand exchange mechanism (chapter 3.1)
Asymmetric transfer hydrogenations
Advantages:
 No hydrogen pressure
 Simple catalysts consisting only of aminoalcohol and arene or of Cp ligands
 Can easily be performed in a lab
Mechanism:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 62

Additional bifunctional hydrogenation catalysts:


Shvo catalyst and chiral diamine complexes by Noyori
The Shvo catalyst was one of the first hydrogenation catalysts for ketones reacting via an
outer-sphere mechanism. Closely related is the Knölker complex having iron as the central
atom.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 63

Few years ago, the reaction principle of the Shvo catalyst was transferred to iron:

Additional mechanisms of alkene hydrogenation, which we cannot cover due to time


limitations:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 64

 Radical mechanism (porphyrin and carbonyl complexes)


 Ionic mechanism with sequential proton and hydride transfer via an acidic dihydride or H2
complex; example: [CpW(CO)2(PPh3)(O=CEt2)]+

4.3.2. Oxidative addition as a key step in hydrogenation catalysis

4.3.2.1. Oxidative addition in Wilkinson's catalyst

The “classic” catalyst

Kinetic studies:
 Oxidative addition either at the starting complex or (104 times faster) at [RhCl(PPh3)2]
 Contribution of the two species depends on their relative concentration and thus on the
concentration of free PPh3
 In the absence of a high concentration of added PPh 3 this reaction mainly takes place via
the 14e complex [RhCl(PPh3)2], although it exists only in low concentration
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 65

How does the addition of H2 occur?  Numerous theoretical studies:


 H2 approaches side-on
 Cis addition with small barrier preferred (usually <10 kcalmol-1)
 Direct trans-addition of H2 “prohibited” (high barrier)
 “Early” transition state with slightly elongated H-H bond
 Various mechanistic scenarios:
 H2 complex as intermediate (reaction coordinate diagram 1, top)
 H2 complex merely as transition state (reaction coordinate diagram 2, center)
More information about stable H2 complexes shortly (reaction coordinate diagram 3, bottom)

4.3.2.2. Vaska's complex

…is accessible to numerous oxidative addition reactions:


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 66

 Oxidative addition of H2 via a dihydrogen complex


H2 complex can be isolated, but loses H2 again in the vacuum  Addition is reversible!
Relative rate constant for the addition X = I > Br > Cl (100:14:1); reason: Better  donation
X  Ir for X = I
 Oxidative addition of MeI and other good electrophiles (benzyl bromide, allyl halide,
chloromethyl ether) via an SN2 reaction - the two new ligands are trans-coordinated!
Higher halides add to Vaska's complex via more complex mechanisms (see below).
First step in analogy to the alkylation of an amine, reaction is faster in polar solvents,
inversion of the configuration at the C atom, kinetics of second order,
relative rate constant for the addition: X = I > Br > Cl >> I
Relative rate constants for various alkyl halides: Me > primary > secondary >> tertiary

4.3.2.3. Oxidative addition of alkyl halides RX to [Pt(PPh3)2]

 Radical mechanism

The reaction takes place with homolytic cleavage of the R-X bond - Transfer of an electron
(SET - single electron transfer) of Pt(0) to the σ*-C-X orbital  Increase of the oxidation state
and coordination number of M by one unit each, subsequently recombination of the radical
pair that formed
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 67

4.3.2.4. Oxidative addition of CH bonds

Iridium(I) complexes: A. H. Janowicz, R. G. Bergman, J. Am. Chem. Soc. 1982, 104, 352-354.
Activation of methane: J. K. Hoyano, A. D. McMaster, W. A. G. Graham, J. Am. Chem. Soc. 1983,
105, 7190-7191.
Rhodium(I) complexes: W. D. Jones, F. J. Feher, Organometallics 1983, 2, 562-563.
The most difficult substrate for CH activation is without doubt methane because it has a very
high CH bond energy (435 kJmol-1), no dipole moment and no functional group.
Recent example: Methane functionalization with ethyl diazoacetate as carbene source in
supercritical CO2 by Peréz et al.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 68

Literature:
A. Caballero, E. Despagnet-Ayoub, M. Mar Díaz-Requejo, A. Díaz-Rodríguez, M. E. González-
Núñez, R. Mello, B. K. Muñoz, W.-S. Ojo, G. Asensio, M. Etienne, P. J. Pérez, Science 2011, 332,
835-838.
V. N. Cavaliere, D. J. Mindiola, Chem. Sci. 2012, 3, 3356-3365. (Review)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 69

CH activation of methane in nature: Soluble methane monooxygenase (sMMO)


Cytochrome P450 is not able to hydroxylate. Nevertheless, there are (so-called
methanotrophic) bacteria, which use methane as the only carbon source. These bacteria have
the methane-monooxygenase enzyme.
Overall reaction:

sMMO has a fascinating structure, which consists of several components.


 Hydroxylase unit MMOH: 251 kDa heavy, heart-shaped 222 heterodimer. Each 
subunit contains a helix bundle consisting of four  helices. Between the helices, there
is the carboxylate- and hydroxo-bridged diiron center, where the activation of O2 and
the hydroxylation of methane takes place.
 Reductase MMOR: Provides the electrons for the reaction, 38.5 kDa heavy iron-sulfur
flavoprotein, which transports electrons from NADH via FAD and [2Fe-2S] cofactors to
the active center of the hydroxylase.
 Third components MMOB: Acts as regulator of the reaction, 19.5 kDa, it does not
contain any prosthetic groups, but modulates the reactivity of sMMO by forming
specific protein-protein complexes with the hydroxylase, which indirectly influence the
structure and reactivity of the diiron center.
The oxidation of methane by oxygen takes place in the  subunit of the hydroxylase MMOH
at a dinuclear iron center. The diiron unit contains two six-fold coordinated iron atoms with
bridging glutamate and hydroxo ligands. The active center of MMOH was characterized in
utmost detail by numerous spectroscopic techniques: EXAFS, ESR, ENDOR/ESEEM, UV-vis,
MCD and Mössbauer spectroscopy.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 70

The mechanism of sMMO was investigated on model compounds in quantum chemical studies
and examinations.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 71

Mechanistic steps:
1) Resting state is the form FeIIIFeIII (MMOHox)
2) Reduction by the reductase protein MMOR makes the FeIIFeII (MMOHred) form
3) Reaction with O2 leads to the superoxo intermediate (not yet spectroscopically observed
but plausible based on theoretical and kinetic studies). It is of a mixed-valent FeIIFeIII
structure.
4) Formation of a peroxo intermediate. FeII ion is oxidized to FeIII.

4.3.2.5. Digression: σ complexes and agostic interactions

The first step of an oxidative addition is often the formation of a σ complex or an agostic
interaction.
First example of a stable σ complex: Kubas complex
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 72

G. J. Kubas, R. R. Ryan, B. I. Swanson, P. J. Vergamini, H. J. Wasserman J. Am. Chem. Soc. 1984,


106, 451-452.
1) H2 complexes:
Reviews: G. J. Kubas, PNAS 2007, 104, 6901-6907. G. J. Kubas, Chem. Rev. 2007, 107, 4152-
4205.
 Stable H2 complexes can also be described as arrested dihydrides (Kubas, 1984)
Bonding situation:
 Bonding: Filled ligand σ orbital with unoccupied metal d orbital
 Backbonding: Occupied metal d orbital and unoccupied ligand σ* orbital:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 73

There is actually a continuum ranging from H2 complexes to "non-classical" hydrido complexes


and classical hydrides - classification:

How can H2 complexes be distinguished from hydrido complexes?


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 74

 Neutron diffraction: See above


 NMR spectroscopy
1) H-D coupling constant: An H-D coupling is 1/6.5 of an H-H coupling due to the
gyromagnetic ratios of H and D. H-D coupling in free H-D is 43 Hz, the H-D coupling in
H2complexes is also considerable (e.g. JH-D = 34 Hz) in the Kubas complex.
2) Time T1:
Time T1: Spin-lattice relaxation, time T2: Spin-spin relaxation
Time T1 correlates with the H-H distance if the relaxation of the respective H atom is
dominated by the dipolar relaxation of the neighboring H atom, in which case the following
applies: T1(min)  [d(H-H)]6 (sixth power of the H-H distance!).
H2 complexes: T1min < 35 ms (at 250 MHz), nearly all hydrido complexes: T1min > 50 ms
T1(min) must be determined by temperature-dependent measurements because T1 is highly
dependent on temperature (and also on the measurement frequency!).
Time T1, however, is also influenced by many other factors (e.g. the gyromagnetic ratio of
specific metals, e.g. Mn, Re and Co); nevertheless, the T1 criterion is often applied (except
with the metals mentioned)
 IR spectroscopy: H-H stretching vibration usually between 2,400 and 3,100 cm-1,
frequently weak, though, and difficult to identify

What are the prerequisites for stabilizing H2 complexes?


 Stabilization of backbonding, but not too much (otherwise: Oxidative addition), therefore:
 Low oxidation state of the metal
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 75

 Frequently cationic complexes, for example isoelectronic Wolfram's and rhenium


complexes:
[W(H2)(PiPr3)3(CO)3]: Mix (equilibrium) of H2 complex and dihydrido complex
[Re(H2)(PiPr3)3(CO)3]+: H2 complex exclusively
 With neutral complexes, frequently  acceptor ligands, in particular CO with neutral H2
complexes because they weaken the backbonding to the H2 ligand
 First-row and second-row metals favor H2 complexes, 5d metals rather form classical
hydrides, e.g.
[RuH2(H2)(PPh3)3] (H2 complex) vs. [OsH4(PPh3)3] (hydrido complex)
[Mo(H2)(CO)(dppe)] (H2 complex) vs. [WH2(CO)(dppe)] (dihydride)

Reactivity of H2 complexes:
 Oxidative addition of H2 ( catalysis)
 H2 complexes are frequently at equilibrium with the corresponding hydrido complexes
Energy barrier for the transformation results from structural changes (coordination
geometry, ligands), which must take place when the dihydrido complex is formed
In some cases, H2 complexes were observed as kinetic products when reacting with H2 at
low temperatures before the transformation into dihydride takes place.
 Dissociation of the H2 ligand
H2 is usually easily displaced by simple dative ligands, e.g. by N 2, and rarely by H2O and
ether
Example:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 76

 Deprotonation through external bases


Many cationic H2 complexes are strong acids, e.g.
Complex pKa value Comment
[RuH2(H2)(PPh3)3] 36 highly electron-rich complex
[Cp*Re(CO)(H2)(NO)]+ -2 cation
[Cp*Ru(CO)2(H2)] + -2 cation
[Os(CO)(H2)(dppp)] 2+ -5.7 dication
free H2 49

Relevance of H2 complexes for catalysis - Example of hydrogenases


Hydrogenases catalyze the production/oxidation of dihydrogen

Some anaerobic microorganisms use protons as terminal electron acceptors. Others use H2 as
reducing agent in order to reduce inorganic species such as CO2 (acetogenic, methanogenic
and photosynthetic bacteria), Fe3+ ions, N2, NO3 and SO42 ions.
Most organisms with H2 metabolism are anaerobic prokaryotes. There are also some
eukaryotes and aerobic bacteria that produce energy from the oxidation of H 2 with O2, with
the formation of water.
The production and oxidation of H2 is catalyzed by hydrogenase enzymes (H2ases).
Interestingly, they are one of the most primitive enzymes in biology. Some hydrogenases are
related to subunits of the complex I of mitochondria, an important component of the
respiratory chain. This indicates that these hydrogenases and complex I are evolutionarily
related.
Three types of hydrogenases are known and were also isolated from sulfate-reducing bacteria
and structurally characterized:
[NiFe] hydrogenase: Heterodinuclear Ni-Fe unit in the active center, usually involved in H2
oxidation
[FeFe] hydrogenase (iron-only hydrogenase): Homodinuclear Fe-Fe unit in the active center,
usually involved in hydrogen production
[Fe] hydrogenase: Only a single iron atom in the active center
All three types are sensitive towards atmospheric oxygen. [NiFe] and [FeFe] hydrogenases
both have similar dinuclear centers but are not evolutionary related.
Example: [NiFe] hydrogenase
The activity of the [NiFe] hydrogenase is lower by one to two times the power of ten than the
[FeFe] hydrogenase. However, H2 binds about 100 times better. [NiFe] hydrogenase is less
sensitive towards O2 and CO.
The structures of a row of [NiFe] hydrogenases were crystallographically determined. The
structures are very similar: Globular heterodimers of 3 nm radius, two subunits with different
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 77

sizes (26 and 63 kDa). The large subunit contains the active center with the dinuclear Ni-Fe
unit. The small subunit contains three iron-sulfur clusters for the electron transport: [4Fe-4S]
cluster at a distance of 1.4 nm from the active center, [3Fe-4S] cluster further away and a [4Fe-
4S] cluster on the surface of the protein. The protein still consists of a large cavity and several
hydrophobic channels.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 78

b) [FeFe] hydrogenase
For the Fe-only hydrogenase, crystal structures of two bacteria (C. pasteurianum and
Desulfovibrio desulficans) are known. Both show the same active center with a dinuclear diiron
complex and terminal CO and CN ligands. The two iron atoms are bridged by a
di(thiomethyl)amine ligand. Similar to the [NiFe] hydrogenases, the protein consists of three
[4Fe4S] clusters, which are located at relatively regular intervals of about 1.2 nm between
the active center and the surface. The [4Fe4S] cluster that is closest to the dinuclear center
is bound to one of the Fe atoms via a thiolate bridge.
The unit from the dinuclear reaction center, the thiolate bridge and the adjacent [4Fe4S]
cluster is also called “H cluster”. Here again, a hydrophobic channel connects the dinuclear
reaction center with the surface.

Crystallographic and IR studies show differences in the oxidized and reduced form of the
[FeFe] hydrogenase of D. desulfuricans. In the oxidized form, a CO molecule bridges the two
iron atoms asymmetrically. Fe1 has six ligands, which are distorted octahedrally coordinated.
Fe2 has five ligands and a free coordination site, which is also free in the reduced form or
which is occupied by a water molecule.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 79

2) Additional examples of catalytically relevant σ complexes:


 Silane and borane complexes
 Alkane complexes:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 80

Stable CH complexes are hardly known, but there are indications (isotope effects, NMR
studies)
Example: H-D exchange between hydrido ligands and alkanes almost certainly takes place
via an alkane σ complex (R. G. Bergman)

First crystallographic characterization of a σ alkane complex by Weller et al. in 2012:


S. D. Pike, A. L. Thompson, A. G. Algarra, D. C. Apperley, S. A. Macgregor, A. S. Weller,
Science 2012, 337, 1648-1651.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 81

3) Agostic coordination of CH bonds (“agostic complexes”):

= intramolecular coordination of aliphatic CH bonds


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 82

Unlike CH σ-complexes, there is no “entropic punishment” when coordinating the CH bond.
The strength is specified as 4-40 kJmol-1. The bonding situation may be described in a way
similar to H2complexes. The agostic interaction may be described as “arrested” CH bond
activation.
Different types of agostic interaction:

 -agostic interaction: very frequent; interaction with H atom at the  carbon atom 
Intermediate on the path towards -H elimination
 -agostic interactions: In complexes with highly electrophilic metal atoms, also in carbene
complexes with metal atoms in high oxidation states
 Agostic interactions between unsaturated metal atoms and ligand-CH bonds, examples:
Kubas‘ H2 complex, [RuCl2(PPh3)3], unsaturated Pd(II) and Rh(I) complexes
Agostic interactions are particularly frequent in cationic complexes. Agostic complexes can
thus be synthesized via protonation of alkene complexes.

Compare 1J(CH) of CH4 and C2H6 = 125 Hz


Methods to identify agostic interactions:
 X-ray crystal structure analysis: Not very reliable, but short M-C distances and distortions
in bond lengths and angles may provide clues
 IR spectroscopy: Reduced CH stretching vibration due to reduced CH bond order
 NMR spectroscopy: High-field shift of the NMR signals, reduced 1J(CH) couplings - see
above

Recommended literature on H2 complexes:


G. J. Kubas, “Dihydrogen complexes as prototypes for the coordination chemistry of saturated
molecules”, PNAS 2007, 104, 6901-6907.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 83

G. J. Kubas, “Fundamentals of H-2 binding and reactivity on transition metals underlying


hydrogenase function and H2 production and storage”, Chem. Rev. 2007, 107, 4152-4205.
Recommended reviews about CH agostic interactions:
M. Brookhart, M. L. H. Green, “Carbon-hydrogen-transition metal bonds”, J. Organomet.
Chem. 1983, 250, 395-408.
M. Brookhart, M. L. H. Green, L.-L. Wong, “Carbon-Hydrogen-Transition Metal Bonds”, Prog.
Inorg. Chem. 1988, 250, 395.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 84

4.3.3. Mechanism of the asymmetric alkene hydrogenation

Hydrogenations with Wilkinson's catalyst take place stereoselectively in the sense of a syn
addition. If the alkene to be hydrogenated is prochiral, a racemate of two enantiomers forms.
Enantioselective hydrogenations of styrene derivatives were described for the first time by
Leopold Horner and William S. Knowles in 1968 as [RhCl{(S)-(+)-PPhMePr}3]; however, only ee
values up to 15% were obtained.
Enantiomeric excess up to 90% with cationic rhodium complexes [Rh(L2)(S)2] with chiral
diphosphine ligands L2:
Ligands:
No general connection between position of the chirality centers and the catalytic efficiency.
Chirality may be centered at the P atom (DIPAMP), and also at the backbone (DIOP, GLUP;
DuPhos): DIOP was the first efficient P,P chelating ligand (H. Kagan, 1971). Helical chirality with
BINAP. Decisive factors: Form and conformational stability

Example: Synthesis of L-DOPA (L-(3,4)-dihydroxyphenylalanine)


Amino acid L-DOPA is an efficient drug to treat Parkinson's disease - it reduces the lack of
dopamine in the brain.
Conventional synthesis by Hoffmann-LaRoche is complicated: Palladium-catalyzed
hydrogenation of a cinnamic acid amide, racemate separation and deprotection
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 85

Considerable simplification by means of enantioselective, homogeneous catalytic


hydrogenation: L-DOPA in only one step with 95% ee.

Olefin mechanism: Acetamido cinnamic acid esters react with cationic rhodium complexes
[Rh(L2)(S)2]+, forming diastereomer complexes.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 86

Two effects are crucial:


 Stability of the diastereomers (thermodynamic stability): With the L-DOPA synthesis, the
“D-diastereomer” 44b is more stable and is produced in excess (steric effect of the shape
of the coordination pocket)
 Reactivity of the diastereomers (kinetic stability): Oxidative addition of H2 is the rate-
determining step and is faster with “L-diastereomer” 44a since the Rh atom in 44b is
sterically more shielded
 kinetically controlled enantioselectivity
This is caused by the Curtin-Hammett principle: The difference of the free enthalpies of the
transition states is crucial for the product composition, not the position of the equilibrium
between the conformers of the reactant

The Curtin-Hammett principle is the basis for understanding why a conformer present in a low
concentration can be product-determining:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 87

Key publication for independent study:


A. C. S. Chan, J. J. Pluth, J. Halpern, "Identification of the Enantioselective Step in the
Asymmetric Catalytic Hydrogenation of a Prochiral Olefin", J. Am. Chem. Soc. 1980, 102, 5952‒
5954.
Additional examples for enantioselective hydrogenations:
(S)-Naproxen (anti-inflammatory, pain-relieving, fever-reducing)
 (R) enantiomer liver-damaging
 Monohydride mechanism where the alkene inserts into a monohydride species
 Release of the product via hydrogenolysis of the alkyl compound that formed
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 88

Metolachlor:
 Impressive productivity and activity make the use of iridium catalysts economically viable
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 89

5. Catalytic carbonylations

5.1. Introduction

 Important for industry and used by industry on a huge scale:


o Many millions of tons of aldehyde and ketone by hydroformylation of propene
o Millions of tons of acetic acid by carbonylation of methanol
 Good availability of CO and CO/H2 mixes from coal (water-gas shift reaction: CO + H2O 
CO2 + H2)
 Earlier, 3d metals, particularly cobalt; now, platinum metals: Rh, Ir and Pd
5.2. Rhodium-catalyzed carbonylation of methanol: Monsanto acetic acid process

 One of the largest and most successful applications of homogeneous catalysis


 In 2004, an estimated 7.7106 t of acetic acid produced, about 80% of which by
carbonylation of MeOH
 In combination with iodine, all metals of groups 9 and 10 are active catalysts
 BASF high pressure process (770 atm, 210 °C) in 1965 on the basis of cobalt and iodide
 1970, Monsanto acetic acid process: Rhodium and iodide (both expensive!) at 180 °C and
30-40 atm, economical because hardly anything gets lost from the catalyst
 Variation: Tennessee-Eastman acetic acid anhydride process, formation of acetic acid
anhydride
 BP Cativa process since 1996: Iridium and iodide

Monsanto acetic acid process:


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 90

The mechanism of the Monsanto process is very well studied:


 Nearly any rhodium and iodide source may serve as starting material  Formation of
[Rh(CO)2I2] under catalysis conditions (proof of high pressure IR, CO = 2055 and 1985 cm-
1)

 Linear correlation of the reaction rate with the concentration of rhodium and methyl
iodide, no dependency of the reaction rate from CO pressure and MeOH concentration
 Oxidative addition of CH3I as rate-determining step

Catalysis cycle in five steps:


1) Reaction of MeOH or methyl acetate with HI - Formation of CH3I
2) Oxidative addition of CH3I to Rh(I) (or Ir(I))
3) Migratory insertion of the methyl group - Formation of an acetyl complex
4) Elimination of acetyl iodide through nucleophilic attack of I
5) Hydrolysis of the formed acyl iodide  Formation of acetic acid and HI

Mechanism of the oxidative addition of CH3I and the reductive elimination of CH3C(O)I:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 91

 Oxidative addition: SN2 attack by [Rh(CO)2I2] on CH3I (see section 3.2b), reason for good
nucleophilicity: Negative charge of the complex, good donor characteristics of I, 105 times
more reactive than neutral complex [Rh(AsPh3)2(CO)I]
 Reductive elimination: Typical reaction path for carboxylic acid derivatives

5.3. Hydroformylation of alkenes

5.3.1. Introduction

Otto Roelen, Ruhrchemie, 1938:


Roelen examined the influence of alkenes on the Fischer-Tropsch synthesis in the presence of
a heterogeneous cobalt-thorium catalyst  Formation of aldehydes. Later, it was discovered
that it is a homogeneous catalyzed reaction with HCo(CO)4 as catalyst. The process was quickly
developed to application maturity. Using terminal olefins results in the formation of n-
aldehydes and, as by-products, undesired iso-aldehydes are generated.
H CHO
HCo(CO) 4, H2, CO
CHO + H n/i = 3-4:1
R R
120-180 °C, 200-300 bar R

n-aldehyde, linear iso-aldehyde,


branched

 Hydroformylation: Formally, addition of a formyl residue and an H atom, also referred to


as “oxo synthesis”
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 92

 One of the most important homogeneous catalyzed reactions: Approx. 7106 t annually
(approx. 1 kg per person on earth)
 Aldehydes: Starting materials for alcohols, carboxylic acids, esters, etc.
 Slide: Aldol condensation of n-butyraldehyde and subsequent hydrogenation results in 2-
ethyl-hexane-1-ol  Ubiquitous as softener for PVC

5.3.2. Cobalt-catalyzed hydroformylation

 Typically high temperatures (120-170 °C) and pressures (syngas H2/CO 1:1, 200-300 atm)
 e.g. propene, 1-octene or internal C10 and C12 alkenes  SHOP process
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 93

 HCo(CO)4: Highly acidic “hydride” formed from Co2(CO)8 and H2


 High temperatures necessary for cost-effective reaction rate
 High CO pressure suppresses the formation of cobalt clusters and cobalt metal
 Mechanism goes back to Heck and Breslow, who examined reactions with model
complexes (organocobalt carbonyl complexes)
Key publication (independent study):
R. F. Heck, D. S. Breslow "The Reaction of Cobalt Hydrotetracarbonyl with Olefins" J. Am.
Chem. Soc. 1961, 83, 4023‒4027
Mechanism:

 Sequence of 18- and 16-electron species as intermediate


 Steps 1-6 are potentially reversible!
 Precatalyst: HCo(CO)4, 18 VE, d8, trigonal bipyramidal
 Step (2): Dissociation of CO, then alkene coordination (3)
 Migratory insertion (4) leads to the formation of n-/iso-RCo(CO)3 species (linear vs.
branched)
 CO addition (5) leads to the formation of 18e species RCo(CO)4, which was independently
isolated and characterized
 Next step: Alkyl migration (6)  Formation of acyl cobalt complex R(CO)Co(CO)3. It is in
equilibrium with R(CO)Co(CO)4 (so-called off-cycle intermediate), which was
independently synthesized. Both acyl cobalt complexes are the only cobalt species that
are detected when hydroformylating 1-octene.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 94

 Formation of the aldehyde by reaction of R(CO)Co(CO)3 with H2 (oxidative addition and


migratory insertion, not shown)  conversion-limiting step

Isomerization of the alkenes used:


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 95

Regioselectivity and reaction rate can be improved by adding phosphines:


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 96

 PBu3 in model studies, under technical conditions, phobanes as particularly stable ligands
 Stronger Co-CO bond due to stronger backbonding by coordination of the phosphine
ligand  More stable complexes; therefore, correspondingly lower CO pressure and
higher temperatures are possible without observing any noteworthy decomposition to
form elementary cobalt.
 Higher n:i ratio
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 97

 Phosphine donors accelerate the rate of hydrogenation of the aldehydes formed 


Formation of alcohols
 Problem: Alkene hydrogenation as side reaction (10-15%)

5.3.3. Rhodium-catalyzed hydroformylation

Second generation of hydroformylation catalysts, developed in the 1970s by Celanese and


Union Carbide

 Wilkinson 1965: [RhCl(PPh3)3] as hydroformylation catalyst at room temperature and


atmospheric pressure, highly selective at high catalyst concentrations
 Lower CO pressure  Low-pressure oxo process
 Very rapid in the absence of added phosphorus ligands: 10 3-104 times faster than with
cobalt
 Nonselective in the absence of phosphine ligands; selective with phosphines
 No alcohol formation, alkene hydrogenation or isomerization as side reactions
 Catalytically active species: [Rh(CO)(H)(PPh3)2]; as catalyst precursors, [Rh(CO)(H)(PPh3)3]
and [Rh(acac)(CO)2] are usually used
 Many rhodium species observed in solution
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 98

Mechanism: Catalytic system is very complex: Many steps, formation of at least two isomeric
products, many intermediates form several isomers, ...
 NMR studies for the reaction with [Rh(CO)2(H)(PPh3)2] as precatalyst (Brown): Two isomers
of [Rh(CO)2(H)(PPh3)2] as resting state (diequatorial:apical-equatorial in the ratio of 85:15);
the equilibrium is reached quickly at room temperature (Berry pseudorotation or turnstile
mechanism)
 In situ IR measurements: These two isomers as dominating species under catalysis
conditions for the hydroformylation of 1-octene (60-100 °C, 5-20 bar)
 Mechanistic steps are similar to the hydroformylation catalyzed by [Co(H)(CO)4]; however,
there is a larger number of intermediates and of stereoisomers of these intermediates
 Migratory alkene insertion, CO insertion and hydrogenolysis of the acyl complex can
proceed both reversibly and irreversibly and controls the stereoselectivity
 Kinetically dependent on the reaction conditions
 Reaction order under technical conditions (T = 170-210 °C, p(CO/H2) = 5-25 bar, [Rh]  1
M, [alkene] = 0.1-2 M): [Alkene] 1st order, [Rh] zero order, [PR3], [CO] negative reaction
order  Conclusion: Migratory insertion of the alkene into the Rh-H bond of
[Rh(alkene)(CO)(H)] as rate-determining step

Generally accepted mechanism:


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 99

The ligands have a great influence on selectivity. Consideration: Diphosphines are supposed
to lead to high regioselectivity since [RhH(CO)(PPh3)2] has a high regioselectivity. Simple
diphosphine such as dppe, dppp and dppb, however, are disappointing (n:i  4:1). Better are
selectivities (n:i  5-8:1) with ligands with a large bite angle  both P atoms can bind in
equatorial position.
High n:i selectivities with diphosphines with large bite angle, well-known e.g. Xantphos (van
Leeuwen et al.)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 100

Digression: Phosphorus ligands in catalysis


Phosphines and related ligands with trivalent phosphorus belong to the most important
control ligands in homogeneous catalysis. All transition metals and in particular the more
electron-rich transition metals form stable phosphine complexes. The high stability results
from the fact that the soft P donor atoms bind well to soft metal atoms in low oxidation states,
with the substituents at the phosphine being able to dramatically influence the characteristics
and reactivity of the metal atom.
The number of phosphine complexes is innumerable. A few particularly important examples
are shown below:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 101

Of particular importance are chelating ligands where a P donor is connected to another


(identical) P donor. However, also asymmetrical diphosphine with two different P donors are
becoming more and more important. In addition, P donors are frequently coupled with other
donors (N heterocycles or charged donors such as Cp, OR and NR2). Due to the length of
the “bridge”, diphosphines usually coordinate in cis configuration. Trans-configurations can
be achieved with very large bridges:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 102

Characteristics of free phosphines:


Free trialkyl phosphines have similar pKa values than the corresponding amines (compare
(HPEt3)+ pKa(DMSO) = 9.1, (HNEt3)+ pKa(DMSO) = 9). Because most transition metals are soft
Lewis acids, phosphines usually bind more strongly to the metal atoms than do amines with
harder N atoms. Steric effects have less impact than with tertiary amines due to the
comparably long MP bond.

Phosphines are easier to oxidize than amines because the oxidation state +V is more stable
for phosphorus than for nitrogen. Due to this reason, many phosphines, in particular alkyl
phosphines, are very sensitive to air. The less electron-rich aryl phosphines and phosphites
are less sensitive to air than are the alkyl phosphines and are often even air-stable. For air
sensitivity, the steric hindrance of the phosphine also plays a role.
The inversion barrier of phosphorus is considerably higher than with amines: Typically, it is in
the range of 2935 kcal mol1. While chiral amines racemize, P chiral phosphines can usually
be isolated as optically active compounds. The P chiral phosphine PMeCy(o-anis) was part of
the first catalyst for enantioselective alkene hydrogenation. (W.S. Knowles et al., J. Chem. Soc.,
Chem. Commun. 1972, 10; Angew. Chem., Int. Ed. Engl. 2002, 41, 1999).
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 103

Bonding and electronic properties of phosphines:


Tolman electronic parameter (TEP)
The interaction between trialkyl phosphine ligands PR3 with transition metal atoms can be
interpreted as Lewis-acid-base interaction between the scattered free electron pair of the
phosphorus and the Lewis-acidic metal atom. Tertiary phosphines and, in particular,
phosphites, however, also have  acceptor ability due to the interaction between occupied
metal d orbitals and P-X-σ* orbitals of the ligands.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 104

A classical approach in order to classify the electronic influence of phosphine ligands is to


determine the Tolman electronic parameter: Value of the highest C-O bond in the IR
spectrum of [Ni(CO)3L].
The value of CO is dependent on the strength of the metal-ligand backbonding. It in turn
depends on the electron density at the metal center and thus on the donor strength of the
phosphine.

 PtBu3 and PCy3 are the strongest donors, PF3 is the weakest donor in this selection.
Steric properties: Tolman cone angle
In addition to electronic properties, the steric properties frequently control the reactivity of
organometallic compounds. The steric demand, however, is difficult to quantify. Referring to
this, Tolman coined the term “cone angle” in the 1970s.
Tolman cone angle: Angle defined by the outermost edge of the substituents on phosphorus
and the metal center in a space-filling model. Since there were no computer models, geometry
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 105

optimization etc. at that time, the angle used to be determined by using homemade models
made of wood and a protractor.
Literature for further reading: C. A. Tolman, Chem. Rev. 1977, 77, 313.

The simple cone angle model has two significant disadvantages:


1. The cone angle is dependent on the conformation, which in turn is dependent on the
complex. This means that the cone angle actually is slightly different at every structure.
2. The cone angle is badly defined with asymmetrical ligands, which means that only an
average can be given
 The Tolman cone angle gives the impression of steric trends; however, the quantitative
value should not be taken too seriously. Sometimes, it is better to give a range for the values,
see PCy3 and P(o-Tol)3 in the table.
In order to eliminate deficits, the term “solid angle” was introduced, which is based on a three-
dimensional angle.
Solid angle (): Normalized shadow area that is cast by the ligand onto a sphere encompassing
the entire complex. The metal center serves as a light source.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 106

Impact of steric and electronic properties on structure and reactivity:


 The tendency towards dissociation increases with the Tolman cone angle
Ex. NiL4 complexes and related phosphine complexes:
Tendency towards dissociation: PMe3 < PMe2Ph < PMePh2 < PPh3 < PiPr3 < PCy3 < PPh(tBu)2
 Effect on the geometry: Sterically charged phosphines favor to bind in a trans order
 There are distortions of the structures due to a high steric demand; Wilkinson's catalyst,
for example, has a non-planar arrangement of the four P atoms in the crystal structure
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 107
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 108

Big problem: Separation and recycling of the rhodium catalyst, in particular when the
aldehydes need to be distilled because this will destroy the catalyst. Solution by Kuntz (Rhone-
Poulenc): Water-soluble triarylphosphine catalysts. The reaction is done in a two-phase
system  the catalyst in the aqueous phase, products in the organic phase. Propene is
sufficiently water-soluble; long-chain alkenes, however, are not sufficiently soluble in order to
be used effectively in the Rhone-Poulenc process.

5.3.4. Enantioselective hydroformylation

Enantioselective hydroformylation has been studied for more than 30 years and have great
synthesis potential for the enantioselective formation of CC bonds. Hydroformylation of vinyl
arenes is of particular interest. The aldehydes that form can easily be oxidized to pain-relieving
drugs such as ibuprofen and naproxen.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 109

Challenges:
 Usually, linear products are preferred with hydroformylation - vinyl arenes, vinyl acetates
and alkenes with electron-withdrawing substituents, however, have a strong preference
for the formation of the branched product.
 Alkene binds exclusively via the  system - no two-point interaction as with many
asymmetrical hydrogenations, which could lead to enantioselectivities that are too high
 Phosphine ligands in the trigonal bipyramidal catalyst are located further away from
alkene than with square planar hydrogenation catalysts
 Hydroformylation must take place faster than the racemization of the enolizable
stereocenter
Breakthrough: Development of phosphite-phosphine ligands by Nozaki, Takaya et al.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 110
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 114

6. Oxidations

Oxidative functionalization of alkenes:


 access to aldehydes, ketones, vinyl ethers, vinyl acetates, allyl acetates, allyl ethers,
enamides, allyl amines, etc.
 of great interest for both industry and academia
6.1. The Wacker process

= Oxidation of ethene to acetaldehyde with palladium and copper


The Wacker oxidation is the most important industrial synthesis of acetaldehyde (low price of
ethene as starting material). The reaction goes back to the year 1894 when F. C. Phillips
demonstrated that palladium(II)-chloride oxidizes ethene to acetaldehyde. In the process, the
divalent palladium is reduced to metallic palladium. In the late 1950s and early 1960s, Smid
and Hafner of Wacker-Chemie in Munich developed a catalytic process, which is still used
today on a large scale:

6.1.1. Mechanism of the Wacker oxidation

Countless studies have dealt with the mechanism of the Wacker oxidation. The sequence of
the elementary steps during the Wacker oxidation seems to depend on the reaction
conditions. Key question: Formation of the C-O bond through nucleophilic attack of H2O on
the coordinated alkene or insertion of the alkene in an M-OH bond??
Two mechanisms, dependent on the Cl- concentration:
 Low Cl- (<1 M) and CuCl2 concentration (<1 M, catalysis conditions): Exclusive formation of
aldehydes
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 115

𝑑[𝐶𝐻3 𝐶𝐻𝑂] 𝑑[𝐶2 𝐻4 ] [𝑃𝑑𝐶𝑙4 2− ][𝐶2 𝐻4 ]


 =− =𝑘 [𝐻 + ][𝐶𝑙− ]2
𝑑𝑡 𝑑𝑡
 Suggestion: Inner-sphere mechanism
1. ligand exchange 2x faster
2. Preceded by a protonation/deprotonation equilibrium
3. Intramolecular syn-hydroxypalladation
4. 2x oxidative addition and reductive elimination of C-H bonds lead to acetaldehyde
 High Cl- (> 3 M) and CuCl2 concentration (> 2.5 M)
𝑑[𝐶𝐻3 𝐶𝐻𝑂] 𝑑[𝐶2 𝐻4 ] [𝑃𝑑𝐶𝑙4 2− ][𝐶2 𝐻4 ]
=− =𝑘
𝑑𝑡 𝑑𝑡 [𝐻 + ][𝐶𝑙− ]
Formation of aldehyde and chlorohydrin through external nucleophilic anti addition
Key step: Anti addition by H2O on the coordinated alkene; the corresponding transition
state is stabilized by CuCl2
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 116

Key publication for independent study:


J. A. Keith, R. J. Nielson, J. Oxgaard, W. A. Goddard, III. J. Am. Chem. Soc. 2007, 129,
1234212343.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 117

6.2. Epoxidations

The epoxidation of ethene to ethylene oxide is done on an industrial scale by using oxygen as
oxidizing agent. A heterogeneous silver contact catalyst is used.

Already with propene, direct oxidation no longer works because too many side products result
from the oxidation of the methyl group. Instead, the oxirane process is used. During this
process, an alkane, e.g. isobutane, is converted into tert-Butyl hydroperoxide (TBHP) by using
oxygen. It is catalytically converted with propene to propylene oxide and tert-Butyl alcohol.

Propylene oxide is used to produce glycols, glycol ethers, alkanolamines, polyesters and
urethanes. Production of 3.106 t∙a-1 (2003)
Catalysts for the oxirane process:
 Heterogeneous Ti(IV)/SiO2 catalyst (Shell)
 Homogeneous molybdenum catalyst (ARCO)
The transfer of the oxygen atom to the propene takes place through direct attack of the
alkene on the electrophilic oxygen atom of an alkyl peroxo complex.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 118

A special variation of catalytic epoxidations is the Sharpless epoxidation (epoxidation of allyl


alcohols with titanium(IV)-tartrate complexes). The catalyst is formed in situ from titanium
isopropanolate and tartaric acid diethyl ester (= diethyl tartrate, DET). Starting from the most
simple educt allyl alcohol, glycidols are formed, which are valuable starting compounds for
pharmaceuticals and agrochemicals  Technical process for glycidols by K.B. Sharpless in
cooperation with the ARCO company.

The production of 1 t glycidol requires 200 kg titanium isopropanolate, 150 kg tartrate and
400 kg of molecular sieves. Allyl alcohol, catalyst and solvent are tempered in the reactor at -
20 °C. Cumol hydroperoxide is added slowly with stirring. When the reaction is completed, the
molecular sieves are removed by filtration and the glycidol is extracted with water. The
organic phase is treated with a reducing agent to destroy solvent residues. The aqueous phase
is then concentrated and the glycidol is obtained with a purity of 88-90% by vacuum
distillation.
The Sharpless epoxidation has the disadvantage that it can only be done with allyl alcohols
and not with unsubstituted alkenes. Simple alkenes can best be oxidized by Jacobsen-Katsuki
epoxidation. Cationic manganese(III)-salen complexes serve as catalysts.
In order to perform an asymmetrical Jacobsen epoxidation, chiral salen ligands must be used.
One of the most successful catalysts is the “Jacobsen catalyst” with trans-cyclohexane-1,2-
diamine as salen bridge. This “Jacobsen catalyst” is used in the industry preferably with a two-
phase system: The alkene is dissolved in dichloromethane; an aqueous bleach solution (NaOCl)
serves as oxidizing agent.
Sterically demanding groups are important for high selectivity. The sterically more demanding
alkene residue is thus being positioned on the back of the catalyst.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 119

Mechanism: Spectroscopic and kinetic studies as well as quantum chemical calculations show
that the “lock-and-key principle” is at work in these reactions, i.e. the transition state resulting
from the energetically lowest intermediate with the lowest steric repulsion of catalyst and
substrate leads to the product. This is an important distinction to the rhodium-catalyzed
hydrogenation where the diastereomer, which is present at lower concentrations, leads to the
product.
Asymmetric dihydroxylation is done stoichiometrically with potassium permanganate or
catalytically with osmium tetroxide as oxidizing agents. Osmium tetroxide itself reacts slowly
with alkenes. However, already Criegee found that adding amine ligands accelerates these
reactions (ligand-accelerated catalysis, Sharpless). Enantioselective reactions can be induced
with chiral amines.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 120

Mechanism:
Osmium acid ester as an intermediate stage, which is hydrolyzed with the dissociation of the
cis diol and the formation of Os(OH)2O2. The latter one is oxidized with tert-Butyl
hydroperoxide or N-methylmorpholine N-oxide (NMO) to OsO4. In the enantioselective
variation, naturally occurring cinchona alkaloids from cinchona bark are used as auxiliaries.

6.3. Iron-catalyzed oxidations

6.3.1. Nature as a model

In nature: Very high substrate specificity, good regioselectivity and stereoselectivity, mild
conditions, O2 as oxidizing agent.
Common characteristic of various oxygenases is the formation of iron(III)-peroxo species as
intermediate. They can either react directly with the substrate or - as takes place in most cases
- form highly oxidized iron-oxo species, which oxidize the substrate. Biomimetic strategies in
oxidation catalysis aim at generating similar intermediates.
Key publication for independent study:
L. Que Jr., W. B. Tolman, „Biologically inspired oxidation catalysis“, Nature 2008, 455, 333 340.

1: Active center of the soluble methane monooxygenase, 2: ketoglutarate-dependent


monooxygenase, 3: Naphthalene dioxygenase, 4: Oxy tyrosinase, 5: O2 adduct of the
peptidylglycine- hydroxylating monooxygenase
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 121

6.3.2. Gif chemistry: Radical reactions

In the lab: The iron salt/H2O2 system to oxidize organic compounds has already been used in
the 19th century. However, mainly highly unselective reactions, e.g.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 122

Fenton reaction:
Oxidation, catalyzed by iron salts (Fenton's reagent), of organic substrates with hydrogen
peroxide in acidic medium, which was discovered at the end of the 19th century by Henry John
Horstman Fenton. In biological systems, Fenton's reaction is seen as the main source of
reactive oxygen species (ROS) in the cell, which lead to oxidative stress. Also on a large
industrial scale, Fenton's reaction and the production of Fenton's reagent are used. In
addition to synthetic applications, it serves as a very strong oxidizing agent, for example to
treat waste water and seepage water from landfills.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 123

6.3.3. Bioinspired oxidation catalysis in the lab

BPMEN (= MEP) = N,N'-dimethyl-N,N'-bis(2-pyridylmethyl)ethane


Many groups developed porphyrin-iron complexes, which serve as biomimetic models for
cytochrome P450.
Iron-porphyrin complexes are models for cytochrome P450 (monooxygenase). Their easy
oxidative decomposition poses an issue, but can be reduced by introducing electron-
withdrawing groups. Electron-poor, aryl substituted porphyrins with Ar = C6H5 or 2,6-Cl2-C6H2-
3-SO3 are particularly suitable.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 124

The mechanism resembles the one used for cytochrome P450 and for related heme
peroxidases. Selectivity depends on the ring size - cyclopentene and cyclohexene give
significant amounts of allyl alcohol and enone as side products.
Explanation for the formation of side products: Oxygen rebound mechanism
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 125

Complexes with tetradentate N-donor ligands mimic the non-heme peroxidases. Important
for high catalytic activity: Two cis-oriented, labile coordination sites similar to the Rieske
dioxygenase. At the same time, there is an important difference to the heme proteins where
only one free coordination site is available. The mechanism of such reactions was explained
for catalysts of the [Fe(TPA)L2]X2 type.
Methods and findings of the mechanistic studies:
 Spectroscopic studies at low temperature
o FeIII-OOH intermediate at 40 °C intercepted in acetonitrile and characterized by
various spectroscopic techniques  mechanistic suggestion: Cleavage of FeIII OOH to
FeV(O)(OH), which reacts with the organic substrate
 18O labeling at room temperature supports this mechanistic support

o Incorporation of H218O into oxidation products with retention of stereochemistry (no


hydrolysis, but oxygen transfer via Fev(O)(OH))
o cis diol contains one oxygen atom of H2O2 and one of H2O!
 Conclusion: Heterolytic bond cleavage FeIII-OOH  Fev(O)(OH)  critical role of proton
donors; big effect by adding acids, e.g. acetic acid

Current example from literature:


Proof of iron(V)-oxo species through Cryo ESI-MS and quantum chemical calculations
I. Prat, J. S. Mathieson, M. Güell, X. Ribas, J. M. Luis, L. Cronin, M. Costas, Nat Chem 2011, 3,
788-793; also see: A. Company, Y. Feng, M. Güell, X. Ribas, J. M. Luis, L. Que, M. Costas, Chem.
Eur. J. 2009, 15, 3359-3362.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 126

The non-heme iron complex [Fe(OTf)2(Me,HPytacn)] catalyzes the hydroxylation of alkanes and
the cis dihydroxylation of alkenes with H2O2 as oxidizing agent functional model of the non-
heme oxygenases. Isotope labeling and DFT calculations allowed to draw the conclusion that
a FeV(O)(OH) species is responsible for the catalytic activity. An iron(V)-oxo species was
observed earlier with the epoxidation of alkenes with peroxides (L. Que Jr., e. L. Bominaar, E.
Münck, T. C. Collins et al., Science 2007, 315, 835-838).
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 127

a, The formation of the oxo-iron(V) species {[Fe(V)(O)(OH)(Me,HPytacn)](OTf)}+ (3O) via a water-


assisted heterolytic cleavage of the O–O bond in {[Fe(III)(OOH)(OH2)(Me,HPytacn)](OTf)}+
(3P·H2O).
b, Cryospray ionization mass spectra (CSI-MS) of the species formed when
[Fe(II)(Me,HPytacn)(OTf)2] (1) was reacted with H2O2 and H2O in acetonitrile solution at -40 C. i,
{[Fe(V)(O)(OH)(Me,HPytacn)](OTf)}+ (3O) generated with H216O2 (10 equiv.) in the presence of
H216O (1,000 equiv.).
ii, {[Fe(V)(O)(18OH)(Me,HPytacn)](OTf)}+ (3Oa) (the additional descriptor a refers to the isotopic
composition of H2O and H2O2 reagents used in the generation of 3O) generated with H216O2
(10 equiv.) in the presence of H218O (1,000 equiv.).
iii, {[Fe(V)(18O)(OH)(Me,HPytacn)](OTf)}+ (3Ob) generated with H218O2 (10 equiv.) in the
presence of H216O (1,000 equiv.).
iv, {[Fe(V)(18O)(18OH)(Me,HPytacn)](OTf)}+ (3Oc) generated with H218O2 (10 equiv.) in the
presence of H218O (1,000 equiv.). All calculated peaks fit the statistical treatment of
experimental error. Red bars correspond to the simulated data (see Supplementary
Information for the full analysis of the isotopic envelope) and black lines correspond to the
experimental data. • = 18-labeled oxygen, a.u. = arbitrary units.

a, The reaction between the oxo-iron(V) species {[Fe(V)(O)(OH)(Me,HPytacn)](OTf)}+ (3O) with


an olefin to form the hydrogenglycolate species {[Fe(III)(C8H14(O)(OH))(Me,HPytacn)](OTf)}+ (5)
and glycolate species [Fe(III)(C8H14O(O))(Me,HPytacn)]+ (4).
b, CSI-MS spectra of the species formed when {[Fe(V)(O)(OH)(Me,HPytacn)](OTf)}+ (3O) was
reacted with an olefin in acetonitrile solution at −40 °C.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 128

i, {[Fe(III)(C8H14(O)(OH))(Me,HPytacn)](OTf)}+ (5) formed by reaction of cyclooctene (100 equiv.)


with {[Fe(V)(O)(OH)(Me,HPytacn)](OTf)}+ (3O) (generated with H216O2 (3 equiv.) in the presence
of H216O (1,000 equiv.)).
ii, {[Fe(III)(C8H14(O)(18OH))(Me,HPytacn)](OTf)}+ (5a) (the additional descriptor a refers to the
isotopic composition of H2O and H2O2 reagents used in the generation of 3O, which in turn
form 5) formed by reaction of cyclooctene (100 equiv.) with
18 Me,H + 16
{[Fe(V)(O)( OH)( Pytacn)](OTf)} (3Oa) (generated with H2 O2 (3 equiv.) in the presence of
H218O (1,000 equiv.)).
iii, {[Fe(III)(C8H14(18O)(OH))(Me,HPytacn)](OTf)}+ (5b) formed by reacting cyclooctene (100
equiv.) with {[Fe(V)(18O)(OH)(Me,HPytacn)](OTf)}+ (3Ob) (generated with H218O2 (3 equiv.) in the
presence of H216O (1,000 equiv.)).
iv, {[Fe(III)(C8H14(18O)(18OH))(Me,HPytacn)](OTf)}+ (5c) formed by reacting cyclooctene (100
equiv.) with {[Fe(V)(18O)(18OH)(Me,HPytacn)](OTf)}+ (3Oc) (generated with H218O2 (3 equiv.) in
the presence of H218O (1,000 equiv.)).
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 129

Jacobsen:
 Example for using acetic acid (30 mol%) as additive for efficient alkene epoxidation
 Terminal alkenes can be epoxidized with a yield of up to 90% by using only 3 mol% catalyst
(50% excess H2O2).
Beller:
 Enantioselective catalyst with high ee's and very high yields
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 130

 Generating the catalyst in situ from FeCl3, chiral ligands and two equiv. pyridine-2,5-
dicarboxylic acid
 No enlightening insights into the mechanism so far!
Iron catalysts can also be used to hydroxylate hydrocarbons. Classical test systems are
cyclohexane and adamantane, but also complex substrates with tertiary CH bonds can be
selectively hydrated.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 131

Literature

Excellent overview: “Bioinspired Nonheme Iron Catalysts for C−H and C=C Bond

Oxidation: Insights into the Nature of the Metal-Based Oxidants”, W. N. Oloo, L. Que, Jr. Acc.
Chem. Res. 2015, 48, 2612−2621.

7. Metathesis

Greek:  = displacement, transposition, change in place


7.1. Alkene metathesis

7.1.1. Introduction

The metathesis of olefins is a catalytic reaction during which a redistribution of the alkylidene
groups takes place by the scission and regeneration of double bonds.

Answer: Product ratio 1 : 2 : 3 = 50 : 25 : 25


Six types of alkene metathesis have been studied in particular detail:
1) Metathesis in the SHOP process: Equilibration in order to increase the molecular weight
distribution in combination with the distillation of the desired fraction from the equilibrium in
the SHOP process
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 132

2) Ring-opening metathesis (ROMP)  polymers, precursors for elastomers after cross-


linking
3) Addition metathesis polymerization (ADMET)  polymers with precise architectures
4) Ring-closing metathesis (RCM)  complex natural products, thermodynamically favored
(entropy), frequently in open systems with elimination of ethene
5) Cross metathesis: Driving force is formed due to the elimination of ethene. Principally, a
statistical mixture of alkenes is formed. Selectivity is possible through kinetic preference of
the reaction of the sterically unhindered carbene complex with an unhindered alkene or of
the reaction of an electron-rich carbene complex with an electron-poor alkene.
6) Enyne metathesis: Metathesis between alkenes and alkynes
7) Alkyne metathesis: Conjugated polymers, ring-closing metathesis with subsequent Z-
selective hydrogenation for the synthesis of natural products (Fürstner)
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 133
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 134

Answer: While preserving orbital symmetry, there is no direct way from the ground state of
the ethene molecules to the ground state of the cyclobutane. For thermal reaction control, a
very high barrier must be overcome due to the symmetry. Thus, the reaction is symmetry
forbidden.

7.1.2. Historical development

 Metathesis of acyclic alkenes, in particular of propene, was realized with heterogeneous


catalysts such as MoO3 (reduced with AliBu3) on Al2O3 as early as the late 1950s
 Catalysts (rhenium, molybdenum and tungsten compounds) are poorly defined, but are
cheap and long-lived; high temperatures are required
 Technical process for the metathesis of propene, which was then available in excess, in
ethene and butene for the first time in 1966 (“Phillips triolefin process”)
 First homogeneous catalyst system (N. Calderon, 1966): WCl6/AlEtCl2/EtOH, highly active
at room temperature, equilibrium is reached within seconds
 Y. Chauvin and J.-L. Hérisson postulated the olefin mechanism with successive [2+2]
cycloadditions in 1971
 Then: Mechanistic insights with alkylidene complexes by Grubbs and Schrock
 2005: Nobel Prize for Y. Chauvin, R. H. Grubbs and R. R. Schrock
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 135
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 136

7.1.3. Catalysts

Two important types of homogeneous catalysts:


1) Schrock catalysts: Molybdenum complexes
2) Grubbs catalysts: Ruthenium complexes
 Both types contain metal-carbon multiple bonds!
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 137

The Hoveyda-Grubbs catalyst was originally developed for immobilization, but has proven to
be particularly successful. Advantages:
 Absence of phosphine ligand, which can take part in decomposition reactions
 Particularly stable thermally; therefore, higher reaction temperatures are possible
 Numerous modifications are possible (works by Blechert, Grela and others)
Immobilized catalysts (Schrock, Coperet, Basset) are also particularly successful:
 Isolated metal atoms  no dimerization as important deactivation path possible
 Asymmetric environment favors higher activity because the electronic requirements of
carbene intermediates and metallacyclic structures are best met  low barriers for [2+2]
and retro-[2+2] addition steps
 Silica is usually used as carrier material; binding occurs via protonation of reactive alkyl or
pyrrole residue through the OH groups of the carrier material
 Up to 10x higher turnover frequencies for the immobilized catalyst
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 138

Synthesis of the Schrock catalyst:

 Various alkoxides, e.g. also fluoride-alkoxides


 Anti conformation of the alkylidene ligand preferred because the alkylidene hydrogen
atom can in this case interact well with the tungsten atom
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 139

 Schrock catalysts with molybdenum even more active


 Anti conformation of the alkylidene ligand preferred because the alkylidene hydrogen
atom can in this case interact well with the tungsten atom
 Again, a multitude of alkoxide ligands is possible
Synthesis of the Grubbs catalyst:
Reaction of sulfur ylides with low valent metal complexes with M =Ru, Os or Ir is a good
method to synthesize carbene complexes
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 140

Free, unsaturated carbenes (imidazolin-2-ylidene):


1) Condensation of glyoxal or  diketones with amines
2) Reaction of the formed diimine with paraformaldehyde to form imidazolium salt
3) Alternatively: Reaction with triethyl orthoformate to form the alkoxide
4) Deprotonation of the imidazolium salts with strong base, e.g. sodium amide in liquid
ammonia or reaction of the EtOH adduct with acid
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 141

Saturated N-heterocyclic carbenes (imidazolidine-2-ylidene, dihydro-imidazol-ylidene):


Similar sequence
 Diamine is frequently generated by reducing glyoxal dimine
 Subsequently, formation of the imidazolium salt triethyl formate for generation of the
alkoxy-substituted imidazolidine and for protonation
The reaction of the free carbene is not always a strategy that will lead to success. Therefore,
various methods for synthesizing NHC complexes were developed:
1) Reaction of an imidazolium salt with a complex with basic ligands
2) Transmetalation of silver carbene complexes with metal halogenides
3) Decarboxylation reaction of imidazolium-2-carboxylates

7.1.4. Mechanism

The mechanism of the alkene metathesis is a result of [2+2] cycloadditions and [2+2]
cycloreversions.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 142

The mechanism of the [2+2] reactions was studied in detail with both Schrock and Grubbs
complexes. In general: At first, there is the coordination of the alkene, then the [2+2]
cycloaddition to the metallacyclobutane, followed by [2+2] cycloreversion and dissociation of
the new alkene.
Mechanism of the [2+2] cycloadditions, taking the Grubbs catalyst as an example:
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 143

Important experiments by Katz:


Experiment A
 Ring-opening of cyclooctene with a mixture of 2-butene and 4-octene: pairwise
mechanism would only form products containing 12 and 16 carbons in the absence of any
methathesis between acyclic olefins.
Experiment B
 Statistical 1:2:1 ratio of the isotopomers of ethylene observed, consistent with the “non-
pairwise” mechanism and inconstant with the “pairwise” mechanism
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 144

Because the [2+2] and retro-[2+2] additions are reversible, the principle of microscopic
reversibility needs to be considered. This principle would be violated if the intermediates and
transition states existed only as square pyramidal complexes. In this hypothetical case, the
[2+2] and retro-[2+2] reactions cannot take place via the same intermediates and transition
states. Therefore, at least one intermediate or transition state must possess a trigonal
bipyramidal structure.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 145

An important “complication”: Intermediates with the alkene in cis position to the carbene
ligand have a similar, sometimes even lower energy as compared to the trans isomers.
The metallacyclobutane intermediates were detected in several studies, and their dynamic
behavior and their reaction behavior were analyzed. Piers and Romero succeeded in
synthesizing a ruthenacyclobutane at low temperature and in doing an NMR spectroscopic
characterization:
 Use of 13C marked ethene shows that the C atoms are replaced both in  and in  position
 Energy barriers for the replacement are NMR spectroscopically determined
 Replacement is done without dissociation of the alkene
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 146

The mechanism of the alkene metathesis, which is catalyzed by Schrock complexes, was
studied for a complex with perfluorinated tert-butanolate ligands. The reaction mechanism
takes place via [2+2] cycloadditions, which require the coordination of the alkene  similar to
the Grubbs complexes. Due to this reason, five or six times coordinated molybdenum and
tungsten complexes are inactive in metathesis. Pseudo-tetrahedral complexes are, however,
active catalysts. The syn isomer is more stable than the anti isomer because the carbene-
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 147

hydrogen atom can form an agostic interaction (hyperconjugation C-H*Mo-N) with the metal
atom in the first case. This also shows in the structure in a larger Mo-C-C bond angle (higher
triple bond character of the Mo=C bond). This stabilization gets lost when the alkene is
coordinated. Therefore, the anti isomer reacts about 1,000 times faster than the syn isomer.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 148

Kinetic studies with Grubbs catalysts lead to additional important mechanistic findings.
Consideration regarding the mechanism:
 The starting complex is a 16e complex; however, 14e intermediates play a crucial role
 Faster ligand dissociation of the ligand enables a faster reaction rate
 Ligand L', which is left behind, is supposed to encourage [2+2] addition and to make it
harder for the dissociated L to coordinate again

Conclusions:
1. Weakly donating and sterically demanding ligands increase the reaction rate because they
dissociate more easily
2. N-heterocyclic carbenes make the dissociation of PCy3 harder by two units, but make [2+2]
cycloaddition easier
3. The k2/k-1 ratio (cycloaddition vs. reassociation) is about four units better than when using
tricyclohexylphosphine or triphenylphosphine. The acceleration of the reaction when
using an N-heterocyclic carbene in the Grubbs II catalyst thus results from the strong
acceleration of the [2+2] cycloaddition.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 149

8. Olefin Oligomerization
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 150

8.1. The Ziegler reaction

Ziegler and co-workers (MPI for Coal Research, Mülheim/Ruhr) found that the reaction of
triethylaluminium with ethene (90−120 °C, 100 bar) affords long-chain alkyl aluminium
species. β-Hydride elimination occurs at higher temperatures (320 °C) leading to α-olefins and
aluminium hydrides. AlEt3 is regenerated by ethene insertion into Al−H bonds.

The Aufbau reaction was coupled with an oxidation generating aluminium alcoholates that
were directly converted into fatty alcohols. This process is merely of historical interest
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 151

nowadays. However, it is noteworthy that high molecular weight, completely linear


polyethylene (M = 9·109 g mol−1, melting point 140−143 °C) is formed in the absence of O2.
The Gulf Oil Company and the Ethyl Corporation (Texas) developed technical processes for the
synthesis of α-olefins in the 1960s that are still used today.
Investigations in Ziegler’s group found in 1953 that the dimerization of ethene to butenes was
exclusively observed when certain autoclaves were used. This resulted in the discovery of the
“nickel effect”. Nickel compounds were formed during the purification of the autoclaves used,
which were identified as catalysts. Traces of acetylene in the ethene used were also necessary
for stabilizing the catalyst.
The catalytic cycle involves the insertion of ethene into an Al−C bond followed by a
nickel-catalyzed butene elimination, which occurs as a “transalkylation” at a nickel(0) complex
which also contains an ethene ligand.

These findings inspired the development of nickel-catalyzed olefin oligomerization.

8.2. Olefin Dimerization

The dimerization of ethene is also catalyzed by nickel(II) compounds in the presence of alkyl
aluminium co-catalysts. Nickel(II) hydride complexes are the catalytically active species.
Wilke and Bogdanovic investigate the mechanism of nickel(II)-catalyzed olefin dimerization
using nickel(II) allyl halide complexes as precatalysts. Ethen coordination, insertion and β-
hydride elimination afford a nickel(II) hydride chloride species. The selectivity of the reaction
can be modified by the phosphane. The highest selectivity for dimerization was achieved with
PR3 = PMe3. More sterically demanding phosphanes lead to higher oligomers, because
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 152

β-hydride elimination is increasingly inhibited with sterically demanding groups. Even


polyethylene can be obtained with PR3 = PtBu3.

The catalytic cycle involves coordination and insertion of ethene into an Ni−H bond and β-
hydride elimination. All reaction steps are reversible except for ethene insertion into the Ni−C
bond of a nickel(II) alkyl intermediate. Aluminium alkyl compounds serve as Lewis acid,
forming [AlEtCl2Br]− anions which may serve as hemilabile ligands.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 153

Nickel-catalyzed butene formation may also occur via a nickel(0)-based mechanism involving
metallocyclopentane intermediates. A co-catalyst is not required in this case. However, the
activity of the catalyst is low.
Moreover, Wilke’s nickel(II) catalyst also catalyzes propene dimerization.

The structure of the products is determined by the phosphanes used. Branched dimer
(2,3-dimethylbutene) is favored by sterically highly demanding phosphanes. The catalysts are
extremely active (TOF = 2.1·108 propene / mol Ni · h): 150 t product / g Ni · h can be formed!
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 154

8.3. Ethene trimerization

This reaction is important because 1-hexene is needed as a co-monomer for the synthesis of
LLDPE. Titanium and chromium catalysts are used.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 155

8.4. Ethene tetramerization


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 156

8.5. Shell Higher Olefin Process (SHOP)

Linear α-olefins have various uses. Short chain olefins (C4‒C8) are co-monomers for the
production of polyethylene (LLDPE) and other polymers. α-Olefins with an intermediate chain
length (C8‒C12) are used for producing synthetic lubricants, while those with a chain length of
C12‒C18 are used for the synthesis of detergents and cleaning agents.
The Shell Higher Olefin Process (SHOP) involves the oligomerization of ethylene by
homogeneous nickel catalysts. SHOP is a process for producing linear α-olefins (C12‒C18) from
ethene. This process can be coupled with hydroformylation for the synthesis of aldehydes and
alcohols. SHOP is a combination of oligomerization, isomerization and metathesis.
SHOP involves the oligomerization of ethene by nickel catalysts in solvents such as butane-
1,4-diol at 80‒120 °C and 7‒14 MPa. The resulting linear α-olefins are immiscible with the
solvent. This leads to a biphasic system where the product can be separated conveniently
through a simple phase separation process (biphasic catalysis). The olefins are separated into
three molecular weight ranges by distillation: C4-C8, C10-C18, and >C19. The lighter and the
heavier fractions are then isomerized to a mixture of internal olefins over a heterogeneous
catalyst. The lighter and the heavier fractions of the resulting internal olefins are then
recombined over a heterogeneous olefin metathesis catalyst (MoO3/Al2O3). Linear C4 to C10+
olefins are obtained selectively (99% linear olefins, 1‒2% internal olefins). Anionic nickel
complexes with P,O-ligands are used as catalysts.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 157

High linearity is observed (approx. 99%, only 1-2% internal olefins).


The oligomerization of olefins can form alkenes of a single length, or they can form a
distribution of alkenes of various chain lengths. Dimers and trimers are typically formed
selectively. Higher α-olefins are typically formed as mixtures. The distribution of olefin chain
lengths produced from oligomerization depends on the relative rates of growth of the chain
and termination of the chain. The distribution of molecular weights resulting from these
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 158

relative rates is called Schultz-Flory distribution and obeys the following quantitative
relationship:
K = [Cn+2]/[Cn] = 1/(1 + β)
β…is the relative rate for chain transfer and propagation
Olefin polymerization occurs when when chain transfer is much slower than chain growth.
Olefin oligomerization occurs when chain transfer occurs at rates similar to chain growth.
Olefin dimerization occurs when chain termination occurs after every insertion. Trimerization
typically occurs by a mechanism that is distinct from dimerization, oligomerization, or
polymerization.

Real olefin mixtures from oligomerization reactions mainly contain olefins with short chain
lengths. The longer the olefin chain length the shorter is its molar fraction in the mixture.

Isomerization and olefin metathesis are used to modify the α-olefin product distribution in a
desirable way. These processes afford a mixture of internal olefins (including olefins with an
uneven number of C atoms) with desired chain lengths, that directly converted into
n-aldehydes by hydroformylation. The desired fraction, e.g C10‒C14, can be isolated by
distillation.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 159

8.6. Acetylene Tetramerization

8.7. Butadiene Di- and Trimerization

Nickel(0) complexes catalyze a variety of cyclooligomerizations of butadiene. This process can


lead to a myriad of products that depend on reaction conditions. The main cyclodimerization
products are 4-vinylcyclohexene (VCH), cis,cis-cycloocta-1,5-diene (COD), and cis-1,2-
divinylcyclobutane (DVCB).
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 160

Dimerization is achieved by the presence of a phosphane ligand. The ratio dimer/trimer can
be controlled by the steric properties of the P donor. The ratio COD:DVCB is strongly
influenced by the electronic properties of the ligand on nickel. COD is an important
intermediate in the chemical industry.

The cyclotrimerization of butadiene to cyclododeca-1,5,9-triene (CDT) is another very


important oligomerization reaction. This reaction was one of the first organometallic reactions
subjected to careful mechanistic study. Three out of four isomeric cyclododeca-1,5,9-trienes
are obtained. The all-trans isomer t,t,t-DCT is the main product. Minor amounts of c,t,t-CDT
and c,c,t-DCT are also observed, while c,c,c-DCT is not formed. Cyclodimers and higher
oligomers are formed as by-products.

The reaction is catalyzed by butadiene complexes of nickel(0) that are void of any other ligands
(‘naked’ nickel). The catalyst are preferably formed by reducing nickel(II) precursors such as
Ni(acac)2 with AlEt2(OEt) in the presence of butadiene, by reductive elimination from nickel(II)
allyl complexes or from nickel(0) precursors such as [Ni(cod)2].
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 161

‘Naked’ nickel forms an equilibrium of several butadiene complexes [Ni(C4H6] in butadiene.


The energy differences between these complexes are small. DFT calculations have given
detailed insight into the reaction mechanism.
The mechanism contains the following steps:
A) Oxidative coupling/reductive cleavage. C-C bond formation via the coupling of two
terminal C atoms from two butadienes; formation of a nickel(II) bis(allyl) complex (16VE)
which contains an additional butadiene ligand. The reaction is reversible.
B) Allyl insertion. Butadiene inserts into an η3-allyl group; formation of an η3, η3-
dodecatrienediyl nickel(II) complex (18VE); the bis(η3-anti) isomer was isolated and fully
characterized by NMR
C) Reductive elimination. Formation of a C-C bond between to terminal C atoms. This results
in the formation of a cyclododecatriene nickel(0) complex (18VE).
D) Ligand substitution. This step reforms the bis(butadiene) nickel(0) catalyst.
Reversible allyl isomerizations (not shown) also play an important role in the mechanism. They
are responsible for the remarkable trans-selectivity of the reaction.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 162
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 163

9. Polymerization
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 164

9.1. Olefin Polymerization

Herrmann Staudinger developed the basic understanding of synthetic polymers in the 1920s
(Staudinger received the Nobel prize in 1953.) The first industrial processes for polymer
production were developed in the 1930s: synthetic rubber, radical polymerization of vinyl
chloride, vinyl acetate and styrene.
There are four basic mechanisms for polymerizations, which are named according the the
natur of the reactive intermediates:
1) Radical Polymerization are started by a radical initiator.
2) Cationic Polymerizations are initiated by Brønsted or Lewis acids.
3) Anionic polymerizations of dienes or vinyl compounds with acceptor groups are initiated
by by bases such as sodium amide and alkyl lithiums.
4) Coordinative polymerizations. Transition metal complexes serve as catalysts.

Karl Ziegler serendipitously discovered Ziegler-Natta polymerization when he investigated the


catalytic properties of transition metal compounds with aluminium alkyls in order to
understand the “nickel effect” more deeply. In 1953, He discovered that the combination of
TiCl4 and AlEt3 in petroleum catalyzes the synthesis of highly crystalline polyethylene (HDPE:
high density polyethylene. Polyethylen was previously only obtained by radical
polymerizations at 200 °C and 1000-2000 bar ethylene pressure. Giulio Natta discovered the
synthesis of highly stereoregular polypropylene using very similar catalysts in 1954. Ziegler
and Natta received the Nobel prize for chemistry in 1963. These discoveries marked the
beginning of modern plastics production. Ziegler catalysts are used to produce roughly 99% of
the world’s polyethylene.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 165

Polyethylenes produced from different catalysts can have different structures. Polyethylene
materials can contain branches from side reactions or the incorporation of comonomers. High-
density polyethylene (HPDPE) is strictly linear polyethylene. Low-density polyethylene (LDPE)
contains extensive branching along the polymer chain and “branches on branches”. Linear
low-density polyethylene (LLDPE) is a linear polymer prepared from ethylene and an α-olefin
that creates a material possessing a small number of branches.

Most HDPE is prepared by heterogeneous Ziegler catalysts containing titantium or by


supported chromium oxide catalysts. HDPE has a higher melting point and is much more
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 166

crystalline than LDPE. Many transition metal catalysts polymerize ethylene to form HDPE, but
few catalysts match the activity of the heterogeneous titanium systems, particularly
considering the low cost of the Ziegler catalysts.

LDPE is synthesized by a free-radical process using peroxide initiators at high temperatures


and pressures. LDPE has a highly branched structure due to hydrogen atom shifts that convert
a primary alkyl radical into a secondary alkyl radical. LLDPE is typically formed from ethylene
and short-chain olefins (typically about ten weight percent of monomers with 4-8 carbons)
using Ziegler catalysts.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 167

Titanium catalysts can contain TiIII or TiIV. Classical Ziegler catalysts are based on the formation
of solid TiCl3 via the reduction of TiCl4 with AlEt3 in petroleum. TiCl3 precipitates as a fibery
solid. The co-catalysts AlEt3 alkylates TiCl3 at the surface. Thereby Lewis acidic AlClxEt3-x (x =
0,1,2) are formed. The catalyst is a heterogeneous catalyst with typical organometallic species
on a surface.
Homogeneous catalysts act via a similar mechanism. TiIII or TiIV species are involved. The first
homogeneous system was already presented by Breslow in the late 1950s. The active species
is a TiIV species in this case. Subsequently, investigations on related cationic zirconium species
showed that cationic complexes of type [ZrCp2P]+ (P = polymer chain) are the catalytically
active species. Thus, weakly Lewis basic anions are required that do not interfere with olefin
coordination. Frequently used anions are [B(C6F5)4]- and [MAO-Me]-. The latter is obtained
from the commercially used co-catalyst methylaluminoxane (MAO).
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 168

Organometallic catalysts are able to catalyze propene polymerization. This is particularly


significant, because it is impossible to reach technically relevant molecular weights of
polypropylene with radical polymerizations due to the relatively high stability of the allyl
radical.

Different regio- and stereoisomers may be formed:


 Regioselectivtivity. Head-to-tail connections may be formed selectively. The formation of
head-to-head and tail-to-tail connections has not been achieved yet.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 169

 Steroselectivity. Polymerization is stereoselective, when all stereocenters in a polymer


chain have the same relative stereochemistry. Isotactic PPP has an …RRRR… or …SSSS…
configuration; syndiotactic PPP has an alternating …RSRSRS… configuration. Atactic PPP
is formed by non-stereoselective polymerizations.

Isotactic PPP is formed when the alkene always coordinates through the Re and the Si face of
the polymer. Syndiotactic insertion is formed by alternating insertion into the Re and the Si
face. Stereoselective Polymerization requires asymmetric induction. Thereby, either the last
asymmetric C atom form is stereoregulating (chain end control) and/or a chiral catalyst
complex (enantiomorphic site control).
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 170
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 171
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 172
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 173
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 174
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 175

Stereoselective polymerization of 1,3-butadiene was first achieved by G. Natta in the 1950s


using “Ziegler Natta catalysts” (vide supra). This is an important reaction because
cis-1,4-polybutadiene has similar properties as natural rubber (cis-1,4-polyisoprene). Hence,
polybutadiene are used for manufacturing tyres, but also for many other applications. In
addition to Ziegler-Natta catalysts, chromium, nickel, cobalt, molybdenum and neodymium
complexes are used as catalysts for the production of butadiene rubber nowadays. In addition,
BuLi is used as a catalyst for anionic polymerizations.
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 176

9.2. Ring Opening Metathesis Polymerization (ROMP)

Ring-opening metathesis polymerization is a reaction with enormous value for polymer


synthesis. This reaction can be used to generate hydrocarbon polymers with favourable
properties, e.g. polymers containing pendant olefins for cross-linking, and polymers with
pendant polar functionalities. Cyclic olefins possessing ring strain are used as monomers
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 177

because thermodynamics will favor ring-opened materials. Dicyclopentadiene is a particularly


favored material. Other common monomers are cyclooctene and norbornadiene.
Many complexes and heterogeneous materials catalyze ROMP. Polymerization processes that
meet the requirements for technical applications of ROMP have been achieved through
proper design or selection of the catalyst. Important parameters are the average molecular
weight and the dispersity of molecular weights (polydispersity index, PDI). Materials of low
polydispersity are formed when the rate of initiation (kinit) is much faster than the rate of chain
propagation (kprop).
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 178
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 179
Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 180

9.3. Epoxide and Lactide Polymerization


Prof. Dr. Robert Wolf, Organotransition Metal Chemistry and Catalysis WS 2017/18 181

9.4. Atom Transfer Free Radical Polymerization

S-ar putea să vă placă și