Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống
1
/ 67 trang
THÔNG TIN TÀI LIỆU
Thông tin cơ bản
Định dạng
Số trang
67
Dung lượng
0,97 MB
Nội dung
CHAPTER 1
Introduction
Transition metal complexes are one of the most important compounds in synthetic
chemistry due to their ability to catalyse many reactions. Of these many reactions,
organic transformations have been the focus of many research groups in the past
decades. Transition metals have incomplete d-shells which are easily accessed
energetically, which allow them to exhibit a variety of oxidation states and
coordination numbers. This property makes these metal complexes ideal candidates
for catalysts.
A catalyst provides an alternative pathway with lower activation energy to
accelerate a reaction. In theory, it should also remain chemically unchanged at the
end of the reaction. Catalysis allows reactions to proceed efficiently and usually
reduces the wastage of chemicals, thus providing atom economy in most cases.
Catalysis can be classified into two main groups which are homogeneous and
heterogeneous catalysis. In homogeneous catalysis, all substrates, including the
catalyst, are in the same phase. However, in heterogeneous catalysis, at least one of
the substrates is not in the same phase.
One of the most common classes of catalysts is organometallic compounds.
Organometallic compounds are defined as complexes containing metal-carbon bonds,
of which a group of such compounds are the metal-carbonyl complexes. Metal1
carbonyl complexes are useful catalysts as the carbonyl ligand can be easily
characterized using spectroscopic techniques such as Infrared and
13
C-NMR. Both
techniques allow the identification of the symmetry, arrangement of ligands, as well as
the functional groups present in the complex.
Metal-carbonyl complexes have distinct carbonyl stretching frequencies within the
range of 1700 – 2100 cm-1, which is a region usually free of most other molecular
vibrations. By monitoring the shifting of the carbonyl stretching frequencies, one is
able to deduce if processes such as ligand association and dissociation have taken
place, or if the symmetry of the complex has been changed. IR spectroscopy is hence a
very useful tool in the investigation of reaction intermediates, and eventually the
mechanism, of a metal-carbonyl-catalyzed reaction.
1.1. Organometallic Compounds in Catalysis
1.1.1. Organometallics
An organometallic compound is usually defined as a substance that contains metalcarbon bonds. Nowadays, this definition is often relaxed to include other non-carbon
ligands such as nitrosyls, cyanates, hydrides and metals with ligands of groups 15 and
16.1-3 Besides that, main group metals such as silicon, boron, germanium, and
tellurium have also been included into the vast family of organometallics.
Organometallics can be sub-divided into various categories such as catalysis,
synthetic organometallic chemistry and bioorganometallic chemistry, with many
2
established works done in these fields. Industrial processes such as the Suzuki
Reaction,4 Reppe Syntheses,5 Fischer-Tropsch Reactions.6,7 the Sonogashira
Reaction,8 and olefin metathesis9 are but a few examples of successful application
using organometallic compounds as catalysis. It is thus the focus of this work to
concentrate on silane transformations catalyzed by rhenium and ruthenium
organometallic complexes.
1.1.2. Organo-Rhenium Complexes
In the periodic table, rhenium is the last natural occurring stable element to be
found in 1925 by Noddack, Tacke and Berg.10 Of the variety of rhenium complexes
that were synthesized, rhenium carbonyls make up a significant part of these known
complexes. Re(CO)5Cl, Re(CO)5Br, Re2(CO)10 and CpRe(CO)3 are but a few examples
of commercially available rhenium carbonyl complexes. Due to their cost, many
rhenium complexes have not been as well studied in terms of their catalytic abilities.11
However, catalytic reactions involving oxo-rhenium complexes (MeReO3)12,13 have
been reported by a few research groups (Scheme 1).
R"
R
R'"
MeReO3
R'
Oxidizing agent
R"
O
R
R'"
R'
Scheme 1. Olefin Epoxidation using MeReO3 as catalyst
Recent work done by various groups explored the catalytic capabilities of rhenium
carbonyl complexes towards organic reactions such as C-C bond cleavage and
3
formation reactions. Kusama et. al.14 reported the use of Re(CO)5Br as a catalyst
towards Friedel-Crafts acylation (Scheme 2). In normal Friedel-Crafts reactions, the
Lewis Acid is usually added in stoichiometric amounts with respect to the substrates.
However, with Re(CO)5Br as a catalyst, the reaction proceeded smoothly albeit with a
mixture of regio-isomers. It is only until recently that Kuninobu et. al.15 reported the
use of Re2(CO)10 as catalyst for regioselective alkylation of phenols.
CH3
O
O
+
Ph
CH3
Re(CO)5Br (10 mol %)
Cl
reflux, 2hr
Ph
91%
Solvent
Scheme 2. Friedel-Crafts acylation catalyzed by Re(CO)5Br
Besides Friedel-Crafts reactions, reactions involving nucleophilic addition to
unsaturated bonds were also explored. Re(CO)5Br was reported16 to be able to
catalyze the condensation of aldehydes and activated methylene complexes with high
stereoselectivity (Scheme 3). The stereoselectivity of the reaction decreases as
unsymmetrical ketones are introduced as the reactant. Besides addition onto carbonyl
compounds, the nucleophilic addition onto C-C double and triple bonds were also
reported.17
NC
O
Ph
R
1
+
NC
R
2
R
2
Re(CO)5Br (0.5mol %)
neat
Ph
R
1
Scheme 3. Nucleophilic addition of activated nitrile onto a carbonyl substrate
substrate
Of more interest are the reported work of Chen et. al.18 and Kuninobu et. al.19 on CH bond activation catalyzed by rhenium (I) carbonyl complexes. The work by Chen et.
4
al. involving the coupling of an alkene with borane complexes (Scheme 4) has since
then been modified by various groups using other transition metal catalysts; whereas
the work by Kuninobu et. al. provided different insights to the reaction pathways and
mechanistic studies of the insertion of unsaturated molecules into a C-H bond
(Scheme 5). Other interesting reactions involving rhenium (I) catalysts like C-Si bond
formation,20 C-N bond formation21 and C-O bond formation22 were also reported. All
the reactions stated above show that rhenium (I) carbonyl complexes have huge
potential in catalyzing a variety of interesting reactions.
t
t
Bu
Bu
NH
N
H
+
[Re(CO)3Br(thf)]2 (3mol%)
Ph
Ph
toluene, reflux, 24hr
Scheme 5. Insertion of unsaturated molecules into a C-H bond
In this project, the following commercially-available, air-stable rhenium (I)
compounds, Re(CO)5Br and Re2(CO)10 were chosen to catalyze the hydrosilylation of
a wide range of carbonyls including aldehydes, ketones and esters. The products
formed from these reactions have wide applications in industrial processes like the
production of polymeric materials.23 Rhenium (I) complexes in the form Re(CO)5X (X
= Br, Re(CO)5) were hence chosen because it was reported to be capable of activating
Si-H bonds. However, their activity towards carbonyl hydrosilylation has not been
fully explored.24
5
1.1.3. Organo-Ruthenium Complexes
Over the past two decades, organo-ruthenium catalysts have provided vast
contributions to developing new efficient organic syntheses such as olefin metathesis
and carbon-carbon bond formation.25 Organo-ruthenium complexes such as [Ru(3allyl)Br(CO)3] 26 are usually used as catalysts for organic transformation reactions as
the ruthenium metal centre is able to adopt a wide range of oxidation states.27 As
such, they have a very good ability to promote catalytic cycles (example of silane
hydrolysis28 depicted in Scheme 6).
CO
Ph3P
Br
Br
Br
Ph3P
Ru
Ru
OC
CO
CO
CO
Br
Br
Br
Ru
Ru
OC
PPh3
Br
CO
CO
Br
PPh3
THF
CO
CO
Br
Ph3P
Et3SiH
OC
THF
Br
CO
Br
Br
Ph3P
Ru
Ru
OC
-Et3SiBr
Br
Ph3P
Ru
H
Si Et3
OC
H
H2O
THF -Et3SiOH
CO
CO
Br
Ph3P
Ru
OC
HO
Br
Et3SiBr
Si Et3
-H2
Br
Ph3P
Ru
OC
H
HO
H
Scheme 6. Silane hydrolysis catalyzed by a ruthenium dimer
Due to their ability to accommodate different oxidation states and hence, different
coordination number, organo-ruthenium complexes containing many different ligands
such as phosphines, cyclopentadienyl, arenes and dienes have been synthesized and
their applications towards catalytic transformations investigated.29 Among these
transformations include bond activation and formation involving C-C,30 C-H31 and Cheteroatom32 (Scheme 7), oxidation reactions33 and hydrogenation.34
6
[Cp*RuCl(Y Me)]2
Ph
NH4BF4
+
O
OH
Ph
CH3
reflux, 3hr
+
H2O
O
Si(OEt) 3
O
O
RuH2(CO)(PPh3)3
+
Si(OEt)3
toluene, 125°C
NMe 2
NMe 2
R'SeSeR'
+
RuCl3
2 R"
X
Zn, DMF, 60-100 °C
2 R'Se
R"
Scheme 7. Examples of C-C, C-H, C-heteroatom bond activation
and formation catalyzed by organoruthenium complexes
Recent interest have been shown in synthesizing ruthenium complexes via oxidative
addition of the allyl halide onto Ru3(CO)1235 to obtain products of the form [Ru(3allyl)(CO)3X] (X = halide). Such complexes were reported to be useful catalysts for
regioselective and steoreospecific allylation reactions under mild conditions.26
Modification of the ligands around the catalyst have also been done to better
understand the mechanism of such allylation reactions towards nucleophilic attack.36
Besides allylruthenium complexes, polynuclear-ruthenium complexes have also
been widely synthesized and their catalytic capabilities towards many organic
reactions were extensively explored. One of the most classic examples of a
polynuclear-ruthenium catalysed reaction would be the Water Gas Shift Reaction, in
which Ru3(CO)12 was utilised as the catalyst.37 Other interesting reactions include the
use of a ruthenium-phosphine dimer as catalyst for the hydrogenation of nitriles to the
7
corresponding amines38 and also oxidation of alkanes to alcohols and ketones.39 Of
more recent interest would be the use of ruthenium dimers containing bridging halides
to catalyze the Kharasch reaction with high efficiency.40
As ruthenium metal has the ability to adopt a variety of oxidation states, many
modifications have been made to organo-ruthenium catalysts to further fine tune their
electronic and steric properties. Ruthenium-NHCs41 have recently been synthesized
and they were shown to be able to catalyze reactions such as Ring-Opening
Metathesis Polymerization,42 and cyclopropanation reaction41 (Scheme 8).
[RuCl2(p-cymene)]2
imidazol(in)ium salt
KO-t-Bu or NaH
n
PhCl, 2hr, 60 °C
Visible light
n
EtO2C
n
EtO2C
[RuCl2(p-cymene)]2
imidazol(in)ium salt
KO-t-Bu
PhCl, 24hr, 60 °C
+
+
CO 2Et
Scheme 8. Examples of ROMP and Cyclopropanation reaction catalyzed by Ru-NHCs
All the above reported organo-ruthenium catalyzed reactions showed the versatility
of ruthenium chemistry in the progression of various fields of research. In this project,
the commercially available and air-stable ruthenium carbonyl, Ru3(CO)12 has been
utilised as a catalyst in the investigation of its efficiency towards catalytic cross
coupling between silanes and other heteroatoms. It was previously reported43 that
ruthenium was able to catalyse the cross coupling between C-S, C-Se and C-Te.
8
+ N2
However, little has been done in this area to further understand the mechanism and
applications of using ruthenium for cross coupling reactions.
1.2. Organosilane Chemistry
Organosilane compounds contain mainly a carbon-silicon bond and its chemistry is
similar to that of normal organic compounds containing carbon-carbon bonds. But as
the silicon atom being more electropositive, the carbon-silicon bond is actually polar.
This leads to certain distinctions from normal organic compounds, ie, nucleophilic
attack usually occurs at the silicon atom, silicon-oxygen, silicon-nitrogen and siliconhalide bonds are stronger than their carbon-heteroatom counterpart, and addition to
olefins occur in an anti-Markovnikov manner, with the ability to hydride transfer.44
Organosilanes are formed initially by reacting elemental silicon with chlorine gas
or hydrogen chloride. Subsequently, reaction with alkali metal alkyl compounds (alkyl
lithium, alkyl sodium etc), with Grignard reagents45 or hydrosilylation of an olefin
under catalytic conditions would yield the desired organosilane (Scheme 9).
Si
2Cl2
Si
3HCl,
RLi
SiCl4
HSiCl3
-LiCl
+RMgCl
RHSiCl2
-MgCl2
-SiCl4, -H2
Si
RSiCl3
RHC=CH2
RCH 2CH 2SiCl3
catalyst
Scheme 9. Formation of organosilanes using different methods
9
Of these methods, the more popular method to derive the required organosilane is
to perform a Grignard reaction with a suitable Grignard reagent while
hydrosilylation of olefins under catalytic conditions would afford very efficient yields.
Chloro derivatives of organosilanes are perhaps one of the most useful classes of
organosilanes as they are frequently utilized as precursors for silicon-heteroatom
compounds such as Si-N and Si-OH. These silicon-heteroatom compounds have great
application in industrial processes and one of the most important processes involves
silicone manufacturing, which requires the reaction of water with chlorosilanes to
form a silanol precursor.44
Of the many reactions involving organosilane compounds, the hydrosilylation
reaction is the only reaction that provides the most application in inorganic,
organometallic, organic, bioorganic, materials and polymer chemistry.46 Especially in
organic chemistry, hydrosilylation provides the most efficient and direct route of
introducing silyl functional groups into unsaturated substrates. The resultant products
are useful intermediates that can participate in a variety of synthetically valuable
organic reactions, with some of them capable of participation in one-pot processes.47
Some of these organic reactions include the Tamao-Fleming protocol,48 which
involves the steoreospecific oxidation of an organosilane into an alcohol (Scheme 10),
and the Hiyama coupling,49 which involves the cross-coupling of organosilanes with
organo halides to form carbon-carbon bonds (Scheme 11).
H
R
1
SiR'2X
R
2
H2O2, KHF2
DMF, r.t.
H
OH
1
R
R
2
H3C
HBF4.Et2O
m-CPBA, Et3N, Et2O
H
Si
CH3
R
1
R
2
X=F,Cl,OR,H,NR2
Scheme 10. Reaction scheme depicting the Tamao oxidation (left to right) and the
Fleming Oxidation (right to left).
10
FR
SiR'3
+
R" X
R
R"
Pd cat.
R = Aryl, Alkenyl, Alkynyl
R' = Cl, F, Alkyl
R" = Aryl, Alkenyl, Alkynyl, Alkyl
X = Cl, Br, I, OTf
Scheme 11. The Hiyama Cross-Coupling reaction.
Besides hydrosilylation reactions, cross-coupling reactions involving organosilanes
also provide many useful applications in organic and bioorganic chemistry.
Transition metal-catalyzed silane hydrolysis and alcoholysis involving the reaction
between organosilanes and water/alcohol produces silanols, silyl ethers and
siloxanes,50 which provide useful starting materials for the manufacture of
pharmaceuticals51 and also as excellent research examples for bioactivity.52 In
addition, such reactions also provide a good source of hydrogen (Scheme 12),50 which
can be utilized in a variety of reactions, with the most classic example being the
hydrogenation of unsaturated compounds catalyzed by palladium (II)53 or iridium (I)54
complexes.
R4-xSi Hx
+
xH 2O
oxorhenium catalyst
xH 2
+ R4-xSi(OH)x
R = alkyl or aryl
x = 1-3
Solv = CH3CN or H2O
Scheme 12. Silane hydrolysis producing hydrogen as the by-product.
Besides silicon-oxygen coupling, silicon coupling with other heteroatoms such as
sulfur and nitrogen also provide many uses. Although organosilanes are useful
11
protecting groups for carbonyl compounds, organosulfur derivatives of these silanes
provide more selectivity in protecting certain functional groups like aldehydes and
ketones (Scheme 13).55 However, the methods of synthesizing these compounds
require harsh conditions or air-sensitive catalysts such as B(C6H5)3 (Scheme 14).56
Also, there are cases whereby sulfur poisoning of the catalyst is a problem as it
caused the drop in efficiency in the system, and eventually the catalytic yield of these
organosulfur derivatives of silanes.57
R
O
R'
+
R" 3Si
R
OSiR"3
R'
SR"'
R"'S
Scheme 13. Protecting groups containing a Silicon-Sulfur bond used
for protecting aldehydes and ketones.
Other than silicon-sulfur compounds, silicon-nitrogen compounds also provide
many useful applications in organic chemistry. The formation of iminosilanes58a,
silazanes58b and diaminosilanes58c has been reported recently by coupling silanes such
as trichlorosilane and trimethylsilane with various amino substrates like
aminofluorosilanes. Besides employing expensive or dificult to synthesize substrates,
harsh reducing agents like butyl lithium were required. However, these reactions were
significant as they provide much insight into the role of amine ligands in processes
such as alkene polymerisation59 and alkene metathesis.60
1.3. Main Objectives
1.3.1. Catalytic hydrosilylation of carbonyls via Re(CO)5Cl photolysis
It was reported that the following commercially-available rhenium (I) complexes,
Re(CO)5X (X = Br, Cl, Re(CO)5) are capable of activating Si-H bonds. However, their
12
catalytic capabilities towards organosilane-related reactions such as hydrosilylation
has not been explored.18 As it was also reported19 that the air and moisture-stable
Re(CO)5Cl dissociates a CO ligand readily upon photolysis (quantum yield of 0.06 to
0.44 from 366nm to 313nm), it shows that Re(CO)5X compounds have the capability
to generate a vacant site for ligand/substrate association.
As it is desirable for any catalyst to be able to perform its catalytic task in the
presence of air and moisture, it is the aim of this part of the project to photoactivate
Re(CO)5Cl and other rhenium carbonyl derivatives in order to carry out
hydrosilylation reactions on a variety of carbonyl substrates at room temperature. The
relative efficiency of the catalysts towards hydrosilylation of different carbonyls will
also be investigated and identification of possible key catalytic intermediates will be
done. Eventually, the mechanism of the hydrosilylation reaction will also be
elucidated using a combination of IR and NMR spectroscopy as well as computational
studies.
1.3.2. Ruthenium carbonyl-catalysed Si-X coupling (X = S, O, N)
In this part of the project, the following air-stable, commercially-available
ruthenium carbonyl compound, Ru3(CO)12, will be used as a catalyst for the coupling
reaction between silanes and various substrates such as thiols, amines and alcohols. It
was previously reported that Ru3(CO)12 can be used to activate Si-H bonds31 and in
the process form dimeric ruthenium silyl species of the form (R3Si)Ru-Ru(SiR3). The
silanes used in this case range from chlorosilanes to alkylsilanes. However, their
catalytic capabilities were not explored. Ru3(CO)12 was also known to be able to
13
activate benzylic C-H bonds in the presence of silanes to afford benzylsilanes,32 while
silylation of silanols with vinylsilanes catalyzed by Ru3(CO)12 to afford alkenes and
siloxanes was also reported. 33
It is hence the aim of this part of the project to utilize Ru3(CO)12 as a coupling
catalyst for Si-heteroatom (S, N and O) coupling. The scope of the catalysis as well as
the kinetics and detection of reactive intermediates will be investigated using FTIR,
NMR and ESI. Computational chemistry using the Gaussian suite of programs will
also be employed for further elucidation of the reaction mechanism.
14
1.4 References
[1] Elschenbroich, C. Organometallics, 3rd ed.; Wiley-VCH, Weinheim, 2005
[2] Crabtree, R. H. The Organometallic Chemistry of the Transition Metals, 4th
ed.; Wiley-WCH, Weinheim, 2005
[3] Hunt, L. B. Platinum Metals Rev. 1984, 28, 76
[4] Chemler, S. R.; Trauner, D.; Danishefsky, S. J. Angew. Chem., Int. Ed. Engl.
2001, 40, 4544
[5] Hanazawa, T.; Sasaki, K.; Takayama, Y.; Sato, F. J. Org. Chem. 2003, 68,
4980
[6] Maitlis, P. M. J. Organomet. Chem. 2004, 689, 4366
[7] Herrmann, W. A. Angew. Chem., Int. Ed. Engl. 1982, 21, 117
[8] Ljungdahl, T.; Bennur, T.; Dallas, A.; Emtena H.; Martensson, J.
Organometallics. 2008, 27, 2490
[9] Hong, S. H.; Sander, D. P.; Lee, C. W.; Grubbs, R. H. J. Am. Chem. Soc. 2005,
127, 17160
[10] Kuninobu, Y.; Takai, K. Chem. Rev. 2011, 111, 1938
[11] Hua, R.; Jiang, J. -L. Maitlis, P. M. Curr. Org. Synth. 2007, 4, 151
[12] Kühn, F. E.; Santos, A. M.; Herrmann, W. A. Dalton Trans. 2005, 2483
[13] Espenson, J. H. Adv. Inorg.Chem. 2003, 54, 157
[14] Kusama, H.; Narasaka, K. Bull. Chem. Soc. Jpn. 1995, 68, 2379
[15] Kuninobu, Y.; Matsuki, T.; Takai, K. J. Am. Chem. Soc. 2009, 131, 9914
15
[16] Zuo, W. -X.; Hua, R.; Qiu, X. Synth. Commun. 2004, 34, 3219
[17] Kuninobu, Y.; Kawata, A.; Yudha, S. S.; Takata, H.; Nishi, M.; Takai, K. Pure.
Appl. Chem. 2010, 82, 1491
[18] Chen, H.; Hartwig, J. F. Angew. Chem. Int. Ed. 1999, 38, 3391
[19] Kuninobu, Y.; Kawata, A.; Takai, K. J. Am. Chem. Soc. 2005, 127, 13498
[20] Zhao, W. -G.; Hua, R. Eur. J. Org. Chem. 2006, 5495
[21] Ouh, L. L.; Müller, T. E.; Yan, Y. K. J. Organomet. Chem. 2005, 690, 3774
[22] Hua, R.; Tian, X. J. J. Org. Chem. 2004, 69, 5782
[23] Ojima, Y.; Yamaguchi, K.; Mizuno, N. Adv. Synth. Catal. 2009, 351, 1405
[24] Hua, R.; Jiang, J. Cur. Org. Synth. 2007, 4, 151
[25] Murahashi, S. -I. Ruthenium in Organic Synthesis, Wiley-VCH:Weinheim 2004
[26] Renaud, J. -L.; Demerseman, B.; Mbaye, M. D.; Bruneau, C. Cur. Org. Chem.
2006, 10, 115
[27] Bruneau, C.; Dixneuf, P.H. Ruthenium Catalysts and Fine Chemistry,
Springer: Berlin, 2004
[28] Tan, S. T.; Kee, J. W.; Fan, W. Y.; Organometallics 2011, 30, 4008
[29] Naota, T.; Takaya, H.; Murahashi, S. -I. Chem. Rev. 1998, 98, 2599
[30] Nishibayashi, Y.; Uemura, S. Cur. Org. Chem. 2006, 10, 135
[31] Trost, B. M.; Toste, D.; Pinkerton, A. B. Chem. Rev. 2001, 101, 2067
[32] Zhao, X.; Yu, Z.; Yan, S.; Wu, S.; Liu, R.; He, W.; Wang, L. J. Org. Chem.
2005, 70, 7338
16
[33] Murahashi S. -I.; Komiya, N.; Hayashi, Y.; Kumano, T. Pure Appl. Chem.
2001, 73, 311
[34] Blaser, H. -U.; Malan, C.; Pugin, B.; Spindler, F.; Steiner, H.; Studer, M. Adv.
Synth. Catal. 2003, 345, 103
[35] Sbrana, G.; Braca, G.; Piacenti, F.; Pino, P. J. Organomet. Chem. 1968, 13,
240
[36] Mbaye, M.D.; Demerseman, B.; Renaud, J.-L.; Toupet, L.; Bruneau, C.
Angew. Chem. Int. Ed., 2003, 42, 5066
[37] Ford, P. C. Acc. Chem. Res. 1981, 14, 37
[38] Grey, R. A.; Pez, G. P.; Wallo, A. J. Am. Chem. Soc.1981, 103, 7536
[39] Neumann, R.; Khenkin, A. M.; Dahan, M. Angew. Chem., Int. Ed. Engl. 1995,
34, 1587
[40] Quebatte, L.; Solari, E.; Scopelliti, R.; Severin, K. Organometallics 2005, 24,
1404
[41] Delaude, L.; Demonceau, A.; Noels, A. F. Cur. Org. Chem. 2006, 10, 203
[42] Maj, A. M.; Delaude, L.; Demonceau, A.; Noels, A. F. J. Organomet. Chem.
2007, 692, 3048
[43] Beletskaya, I. P.; Ananikov, V. P. Chem. Rev. 2011, 111, 1596
[44] Auner, N.; Weis, J. Organosilicon Chemistry IV, Wiley-VCH: Weinheim 2000
[45] Arkles, B. Grignard Reagents and Silanes, reprinted from: Silverman, G.S.;
Rakita P.E. Handbook of Grignard Reagents, Marcel Dekker: New York 1996
[46] Roy, A. K. Adv. Organomet. Chem. 2008, 55, 1
17
[47] Marciniec, B.; Maciejewski, H.; Pietraszuk, C.; Pawluć, P. Hydrosilylation, A
comprehensive review on recent advances, Springer Science+Business Media B. V.
2009
[48] Jones, G. R.; Landais, Y. Tetrahedron 1996, 55, 7599
[49] Hatanaka, Y.; Hiyama, T. J. Org. Chem. 1988, 53, 918
[50] Ison, E. A.; Corbin, R. A.; Abu-Omar, M. M. J. Am. Chem. Soc. 2005, 127,
11938
[51] Srimani, D.; Bej, A.; Sarkar, A. J. Org. Chem. 2010, 75, 4296
[52] Dai, X.; Strotman, N. A.; Fu, G. C. J. Am. Chem. Soc. 2008, 130, 3302
[53] Tour, M. J.; Cooper, J. P.; Pendalwar, S. L. J. Org. Chem. 1990, 55, 3452
[54] Wang, D. W.; Wang, D. S.; Chen, Q. A.; Zhou, Y. G. Chem. Eur. J. 2010, 16,
1133
[55]
Evans, D. A.; Truesdale, L. K.; Grimm, K. G.; Nesbitt, S. L. J. Am. Chem.
Soc. 1977, 99, 5009
[56] Harrison, D. J.; Edwards, D. R.; McDonald, R.; Rosenberg, L. Dalton Trans.
2008, 26, 3401
[57] Chauhan, B. P. S.; Boudjouk, P. Tetrahedron Lett. 2000, 41, 1127
[58] a) Jendras, M.; Klingebiel, U.; Niesmann, J. Organosilicon Chemistry V, Wiley
VCH, 2000, 264; b) Abele, S.; Becker, G.; Eberle, U.; Oberprantacher, P.; Schwarz,
W. Organosilicon Chemistry V, Wiley VCH, 2000, 270; c) Oprea, A.; Mantey, S.;
Heinicke, J. Organosilicon Chemistry V, Wiley VCH, 2000, 277
[59] Bolton, P. D.; Mountford, P. Adv. Synth. Catal. 2005, 347
[60] Hoveyda, A. H.; Schrock, R. R. Chem. Eur, J. 2001, 7, 945
18
Chapter 2
Catalytic hydrosilylation of carbonyls via Re(CO)5Cl
photolysis
2.1 Introduction
The reaction of silanes with a wide range of carbonyl compounds remains
important in the generation of useful protecting groups for organic synthesis. For
example, hydrosilylation across the C=O bond of aldehydes and ketones produces
protected alcohols in a single step.1 In the presence of a catalyst, these reactions
could be carried out under mild conditions in contrast to using harsh reducing agents
such as lithium aluminium hydride. Hydrosilylation catalysis has been reported for
various transition metal complexes containing iron,2-3 tungsten,4 molybdenum,5
manganese,6-7 rhenium,1,8 rhodium,9 silver10 and most recently ruthenium.11 Recent
works were also done on the mechanistic studies of hydrosilylation using oxo-rhenium
complexes.12-13
While Berke et al.8 has reported the use of different rhenium (I) catalysts for the
hydrosilylation of aldehydes and ketones, one of the more interesting processes is the
CpFe(CO)2Me-catalysed reaction between silanes and dimethylformamide (DMF)
reported by Pannell et al.14 As shown in Scheme 1, hydrosilylation across the C=O
bond of DMF first afforded a silyl ether which further reacted with another molecule
19
of silane to give siloxane and trimethylamine as final products. Cutler et al.15 has used
a manganese complex (CO)5MnC(O)CH3 to reduce the much less reactive ester to
ether via a silyl acetal intermediate. Silyl esters are another important class of siliconcontaining compounds as they find wide use in the production of polymeric
materials.16 Mizuno et al.16 reported that a range of ruthenium catalysts enables the
conversion of carboxylic acids to silyl esters to take place. Furthermore, Yamamoto’s
group showed that a boron catalyst, B(C6F5)3, reduces carboxylic acids further to the
corresponding alkanes as depicted in Scheme 1.17 However, most of the reported
carbonyl hydrosilylation catalysts require careful handling as they are relatively
unstable under ambient conditions.
Although the following commercially-available rhenium (I) complexes, Re(CO)5X
(X = Br, Cl, Re(CO)5) have been reported to activate Si-H bonds, their activity
towards carbonyl hydrosilylation has not been explored fully.18 It was previously
reported19 that the air and moisture-stable Re(CO)5Cl dissociates a CO ligand readily
upon photolysis (quantum yield of 0.06 to 0.44 from 366nm to 313nm). In light of
efforts in utilizing energy from light rather than from fossil fuel sources, much work
would have to be carried out to further explore light-driven catalytic processes. As it
is also desirable for the catalysts to be air and moisture-stable and easy to handle, we
have photoactivated Re(CO)5Cl and other rhenium carbonyls in order to carry out
hydrosilylation reactions on a variety of carbonyl substrates at room temperature. The
objectives of the work are to determine the relative efficiency of the catalysts towards
different carbonyls, identify possible key catalytic intermediates and explore the
mechanism of hydrosilylation using a combination of IR and NMR spectroscopy as
well as computational studies.
20
O
O
+
R3SiH
Me 2N
OSiR3
[CpFe(CO)2Me]
H
r.t., h
Me 2N
R = H, Alkyl, Aryl
O
PhSiH3
+
O
R
3% (L)(CO)4MnC(O)CH3
C6D6
OEt
O
Et3SiH
OH
H
B(C6F5)3, r.t.
R
R3SiH
Me 2N
OSiEt 3
PhSiH3
SiR3
+
H
CH3
O(SiH 2Ph)2
OEt
H
+
OEt
OSiEt3
Et3SiH
OSiEt3
R3Si
2 Et3SiH
R CH3
R
H
OSiEt3
- 2O(SiEt3)2
Scheme 1: Catalytic hydrosilylation of carbonyl compounds
2.2 Results and Discussion
2.2.1 Hydrosilylation of silanes with aldehydes and ketones
Preliminary experiments have been carried out to assess the efficiency of UV
photolysis of Re(CO)5Cl towards carbonyl hydrosilylation using Et3SiH and acetone
as substrates (Table 1, Entry 1). The quantum yields for Re(CO)5Cl dissociation at
313 nm and 366 nm have been determined to be 0.44 and 0.06 respectively19. The light
source used in our work is broadband, ranging from 300nm to 800 nm. Since
Re(CO)5Cl absorbs in the 300-350 nm region, the quantum yield would also be
expected in the similar range as quoted. 1H NMR analysis of the mixture after 4 hours
of photolysis showed isopropoxytriethylsilane, Et3SiOCH(CH3)2 (1a) as the sole
21
product (Figure 1), with a turnover number determined to be 70 to 80 with respect to
Re(CO)5Cl (1% mol loading).
OSiEt3
HB C
HA
CH B
Figure 1. 1H NMR peaks of (1a)
Peaks Assignment:
HA - 3.99ppm (septet)
HB - 1.15ppm (doublet)
The reaction appears to proceed most efficiently if an excess of silane is used to
achieve full conversion of acetone to the product. A ratio of silane : carbonyl of 3:1
has been used for further comparison with other carbonyl substrates. However the
silyl ether (1a) does not undergo further reaction with the excess silane. Such is also
the case in the Re(CO)5Cl-catalysed reactions of Et3SiH with other carbonyl
substrates (Table 1, Entries 2-6). When the mixture was heated from 30 to 120°C in
the absence of light, the yield of the resultant silyl ether was less than 10%. No
hydrosilylation product was observed when the mixture was photolysed in air or upon
removal of the rhenium catalyst.
Re2(CO)10 has been found to be an effective catalyst as well (Table 1, Entries 8-9)
with similar TON ratio to Re(CO)5Cl. However hydrosilylation proceeded only
sluggishly when HRe(CO)5 was used (Table 1, Entries 10-11). Compare to previously
reported catalysts7 such as [Re(CH3CN)3Br2(NO)] (TON = 99 for aldehydes/ketones),
similar efficiency has been observed for Re(CO)5Cl. However, hydrosilylation with
22
substrates such as 1,3-diketones, anthraquinones and acetylacetone did not occur
even after extended photolysis of up to 12 hours.
Entry
Substrate
Product[a]
Catalyst used
TOF[b](hr1
)
OSiEt3
O
Re(CO)5Cl
1
(1a)
18
(1b)
20
H
O
2
Re(CO)5Cl
H
O
OSiEt3
O
3
H
H
Re(CO)5Cl
OSiEt3
Et3SiO
20
(1c)
O
4
Re(CO)5Cl
OSiEt3
(1d)
21
(1e)
23
(1f)
11
(1g)
4
H
OSiEt3
O
5
Re(CO)5Cl
OSiEt3
O
6
Re(CO)5Cl
OSiEt3
O
7
Re(CO)5Cl
O
8
Re2(CO)10
(1a)
15
Re2(CO)10
(1b)
18
O
9
H
23
O
10
HRe(CO)5
(1a)
HRe(CO)5
(1b)
3
O
11
H
5
When the substrate was changed from acetone to benzaldehyde, an increase in the
amount of product formed over the same period was observed (Table 1, Entries 1 and
2). This can be attributed to the increased reactivity of aldehydes when compared to
ketones. When different silanes such as Ph2SiH2 and Ph3SiH were used, a decrease in
the corresponding silyl ether yield was observed (Table 2). The relative lower rates
towards hydrosilylation of acetone appear to bear a strong relation to the steric
hindrance of the silane. In addition, when the carbonyl substrate was changed from
aliphatic to a hindered aromatic carbonyl, i.e. benzophenone, acetophenone and
benzaldehyde, a decrease in the TOF was observed.
2.2.2 Hydrosilylation of esters and carbonates
In the presence of Re(CO)5Cl, esters and carbonates have been reduced with
silanes, although at a lower rate compare to the aldehydes. Under our experimental
conditions, the reaction of ethyl acetate and Et3SiH gave ethoxytriethylsilane as the
main product with trace amount of diethyl ether. Acetaldehyde has been detected as
an intermediate whereby its signal intensity decreases upon further photolysis while
that of ethoxytriethylsilane continues to increase (Figure 2). Similar results have been
Table 1. The reactions of Et3SiH with various aldehydes and ketones.
[a]
Product
obtained
after 4hrs of photolysis
in vacuo,
with 1% mol loading of catalyst.
obtained
forwas
methyl
phenylacetate
and methyl
formate.
Ratio of silane to carbonyl substrate is 3:1.
[b]
TOF calculated based on 4hrs of photolysis in vacuo.
24
Entry
Substrate
Silane
Time
Product
TOF[b](hr1 [a]
)
25
1
OSiPh3
O
Ph3SiH
10hrs
(2a)
1
H
OSiPh3
O
2
H
Ph3SiH
10hrs
(2b)
1.5
OSiPh2H
O
3
Ph2SiH2
5hrs
(2c)
8
(2d)
10
(2e)
8
(2f)
14
H
OSiPh2H
O
4
H
Ph2SiH2
5hrs
OSiMe2Ph
O
Me2PhSiH
5
5hrs
H
O
6
OSiMe2Ph
H
Me2PhSiH
5hrs
Table 2. The reactions of Acetone and Benzaldehyde with various silanes in ratio of silane : carbonyl = 3:1
[a]
TON calculated based on 1% mol loading of Re(CO) 5Cl. Yield of product determined based on the
integration of the 1H NMR spectra using toluene as standard.
26
The reaction of Et3SiH with diethyl carbonate afforded a variety of products with
Et3SiOEt (1) and EtOCH2OSiEt3 (2) being the major ones. A trace amount of ethyl
formate EtOCHO (3) has been detected. When the reaction mixture was left overnight,
1
H NMR peaks belonging to (EtO)2CH2 (4) were detected along with a decrease in (1)
and (2).
Figure 2. 1H NMR intermediate studies on the amount of intermediates and
products produced during the reaction of ethyl acetate and Et3SiH
2.2.3 Comparison of hydrosilylation rates for different carbonyls
By comparing the ratio of the products within the same photolysis period, the
following general relationship has been determined for the relative rate using
Re(CO)5Cl and Et3SiH : Aliphatic Aldehydes > Aromatic Aldehydes > Aliphatic
Ketones > Aromatic ketones ~ Esters (Table 3). Comparing within each functional
group, the alkyl analogues are more reactive than their aryl counterparts. Using
pyruvic acid which has both C=O and O-H groups, it was found that Et3SiH
preferentially reacted with the O-H group to produce the silyl ester.
27
Entry
Substrate
1
Ethanal
2
Propanal
3
Benzaldehyde
Ratio[a],[b]
Product
OSiEt3
OSiEt3
OSiEt3
Ph
90
90
9
OSiEt3
4
Acetone
1
OSiEt3
5
Acetophenone
0.5
Ph
OSiEt3
6
Benzophenone
0.03
Ph
7
Ph
Ethyl Acetate
OSiEt 3
0.25
O
8
Pyruvic Acid
OSiEt 3
480
O
Table 3. Ratio of hydrosilylation products formed. Relative rates were determined by adding 3:1:1
ratio of Et3SiH with various carbonyl substrates and acetone.
[a]
Ratio with respect to Acetone.
[b]
Yield calculated on the basis of the integration of the 1H NMR spectra using toluene as standard
2.2.4 Mechanism
The identification of the organic intermediates and products suggests that the first
step of hydrosilylation proceeds via addition of R3SiH across the carbonyl group
except when an O-H group is present. The formation of both H2 and a strong Si-O
bond provides much of the driving force behind the reaction between the alcohol and
28
silane. However our focus is on the addition reaction, hence the complex reaction
pathways for the diethylcarbonate case are highlighted as an example (Scheme 2).
The products of the reaction between Et3SiH and diethylcarbonate were observed
and identified in
triethylethoxysilane
1
H NMR (Experimental Section) and were attributed to
(1),
ethylformate
(2),
ethoxymethoxytriethylsilane
(3),
diethoxymethane (4) and hexaethylsiloxane (5). The formation of these products led us
to postulate the following mechanism: An unstable silylacetal intermediate is first
formed, which undergoes ‘condensation’ to afford (1) and (2). With excess silane, the
ethylformate (2) can be further reduced to (3). This explains the relatively little
amount of formate left in the mixture. In addition, the experimental data also
suggested that compound (1) undergoes further reaction with (2) to generate the ether
(4) and siloxane (5).
O
O
2 Et3SiH
Et3SiOEt (1)
[Re]
+
EtO
SiEt3
OEt
Step 1
EtO
O
OEt
(2)
EtO
Step 2
SiEt3
Et3Si
(5)
+
OEt
EtO
(4)
Et3SiOEt (1)
[Re]
Et3SiH
OSiEt3
[Re]
O
H
(3)
EtO
H
Scheme 2: Proposed reaction scheme of the step-wise reaction
As one of the main objectives of this work is to elucidate the mechanism of
hydrosilylation, FTIR spectroscopy has been utilised for the detection of any metal
carbonyl intermediates present in the catalytic mixture. Upon photolysis of Et3SiH
29
with Re(CO)5Cl, a dimeric rhenium carbonyl species with a bridging hydride
(henceforth known as complex (A)) has been identified in the mixture (Figure 3). The
identification of the complex was based on the IR spectral resemblance to rhenium
complexes20 of formula HRe2(CO)9(SiR3). Its bridging hydride signal at -9.03ppm has
also been recorded in the 1H NMR spectrum. The diphenylsilyl analogue, complex (B)
has also been prepared upon Re(CO)5Cl photolysis in the presence of diphenylsilane.
Other than (A), another hydride signal at -5.77 ppm found in the reaction mixture has
been attributed to HRe(CO)5.
1.2
Abs
1.05
0.9
0.75
0.6
0.45
0.3
0.15
-0
2115
2100
2085
2070
2055
2040
2025
2010
1995
1980
Cy
1965
1950
1/cm
2100
2080
2060
2040
2020
2000
1980
Figure 3. IR and 1H Hydride NMR Spectrum of (A) obtained from photolysis of Re(CO)5Cl with Et3SiH.
IR peaks (cm-1): 2102, 2094, 2085, 2030, 2026, 2020, 2007, 2000, 1978
Lit Values13 for HRe2(CO)9SiCl3(cm-1): 2150, 2095, 2085, 2047, 2019, 2012, 1999, 1978
1
H NMR peaks (ppm): -9.03
The formation of complex (A) can be rationalised as follows (Scheme 3). Upon UV
photolysis of Re(CO)5Cl, a CO ligand dissociates to form the 16-electron Re(CO)4Cl
intermediate followed by Et3SiH coordination. Upon reductive elimination, either HCl
or Et3SiCl would be formed together with the corresponding Et3SiRe(CO)4 or
HRe(CO)4 species. A free CO molecule coordinates back to HRe(CO)4 to give
30
HRe(CO)5. The bridging of its hydride ligand to Et3SiRe(CO)4 would then result in the
formation of (A).
When (A) was isolated and tested for aldehyde hydrosilylation, the silylether was
generated about two to three times faster compared to Re(CO)5Cl. UV irradiation is
still essential for the catalysis to occur. The IR signals of (A) persisted even after the
end of catalysis, with a recovery of about 70-80%.
SiEt3
CO
OC
Re
CO
Et3SiH
- HCl
CyH, h
Re(CO) 5Cl
- CO
CO
+
Re(CO) 4Cl
- Et 3SiCl
Et3SiH
H
CO
OC
Re
CO
CO
CO
CO
OC
SiEt3
CO OC
Re
CO
Re
H
OC
CO
(A)
CO
CO
Scheme 3: Formation of (A) from Re(CO)5Cl photolysis in Et3SiH
Interestingly, complex (A) can also be generated upon Re2(CO)10 photolysis in
triethylsilane, which would explain the similarity in the reaction rate to Re(CO)5Cl.
From these observations, there are reasons to believe that complex (A) acts as a
resting state in carbonyl hydrosilylation. One of the crucial steps in the catalysis
involves the photocleavage of (A) to form HRe(CO)5 and a 16-electron rhenium
31
carbonyl Et3SiRe(CO)4 . As it was shown earlier that catalysis with HRe(CO)5 is
sluggish, we believe that the rhenium silyl species is most likely the active catalytic
species instead.
A mechanism for the hydrosilylation of carbonyl compounds is proposed to account
for the experimental observations (Scheme 4). Upon photolysis, (A) dissociates to
afford Et3SiRe(CO)4. Then the carbonyl substrate undergoes coordination onto the
vacant site and facilitates the silyl ligand shift onto the oxygen atom. This process
results in the formation of an alkyl ligand bound to the Re centre (Steps 1-2). Another
silane undergoes coordination via a 2 silyl-complex4,7 or a sigma (σH) silyl-complex
(Step 3). The H atom migrates from the silane to the alkyl group, thus regenerating
the catalyst and releasing the silyl ether product (Step 4). When either the carbonyl or
silane has been depleted, the R3SiRe(CO)4 would coordinate back to HRe(CO)5 and
becomes part of the resting state (A).
The activity of (A) in the catalytic cycle can be tested by using its Ph2SiH2
analogue, (B) to catalyse the hydrosilylation of acetone and Et3SiH. Upon completion
of catalysis, 1H NMR analysis showed the presence of iPrO(SiEt3) as the expected
main product with a small amount of iPrO(SiHPh2). More importantly, the IR
spectrum has changed from that of (B) to (A). These observations show that silyl
exchange has occurred between (B) and the free Et3SiH during catalysis and lent
support to (A) participating in the catalysis (Scheme 5).
32
CO
OC
SiEt3
CO OC
CO
Re
CO
+
Re
CO
CO
CO
(A)
OC
CO
OC
h
Re
H
OC
CO
SiEt3
CO
Re
OC
H
CO
CO
SiEt3
O
R
- RCH 2COSiEt3
CO
OC
Re
H
CO
CO
Step 1
Step 4
SiEt3
OC
H
CO
OC
CO
R
H
Et3SiO
Re
O
CO
R
SiEt3
CO
Re
CO
CO
H
Step 2
CO
OC
Step 3
Re
R
H
H SiEt3
CO
Et3SiO
CO
Scheme 4: Proposed mechanism of the hydrosilylation reaction
OC
SiPh2H
CO
Re
CO
CO
O
OC
O
SiPh2H
CO
Re
CO
CO
CO
OC
Re
CO
CO
HPh2SiO
H SiEt3
H
SiEt3
CO
OC
- iPrOSiPh2H
SiEt3
OC
Re
CO
Re
CO
CO
CO
Et3SiO
CO
Scheme 5: Silane exchange from (B) to (A) during reaction
33
2.2.5 Calculations
Computational studies have been carried out to show the thermodynamic feasibility
of the catalytic cycle in Scheme 4 and to provide a general explanation for the relative
rates of hydrosilylation among different carbonyl substrates. The energetics and
structures of the complexes together with any transition states have been calculated
using mp2/lanl2dz level of theory and basis set in Gaussian 03.21 Vibrational
frequencies were calculated for the optimised structures. In order to save
computational time, formaldehyde CH2O and silane SiH4 have been used as the
substrates as shown in Scheme 6. No solvent effect has been computed as well.
O
H
SiR3
CO
OC
H
- H 3COSiR3
Re
CO
I
-96.5
CO
IV
-11.61
SiR3
OC
H
CO
OC
CO
H
H
R3SiO
Re
Intermediate 1
O
CO
H
SiR3
CO
Intermediate 3
Re
CO
CO
H
74.27
II
CO
OC
H
Re
CO
H
R3SiO
-45.50
III
H SiR3
CO
Intermediate 2
Scheme 6: Calculated catalytic cycle for SiH4 and CH2O.
Values quoted are for free energies in kJ/mol
34
Table 4 shows the relevant thermodynamic parameters and co frequencies for the
intermediates (restricted to singlet states only) and transition states found for the
cycle while the calculated structures of the relevant intermediates and transition
states are found in Figure 4. The calculations showed that Steps 1 and III in Scheme 6
are exogernic due to coordination of substrates to the relevant rhenium complexes.
In contrast Step II is endoergic as activation is required for the silyl transfer to the
carbonyl ligand in order to form Intermediate II. The transition state TSI for this step
has been found and optimized with an activation energy of 74.3 kJ/mol relative to
Intermediate I or 9.6 kJ/mol relative to CH2O and Re(CO)4SiH3. The reaction
coordinate correctly depicts the migration of the silyl to the O atom of formaldehyde
at an imaginary frequency i of 281 cm-1. Step IV turns out to be exogernic with a
transition state TSII at only 26.0 kJ/mol relative to Intermediate 3. The reaction
coordinate for this transition state shows an almost linear migration of an H atom
from silane to the alkyl group at an imaginary frequency of 798 cm-1.
The formation of the strong C-H bond is the driving force behind the reaction and
accounts for the relatively lower activation barrier. By comparison of the energies,
Step II represents the rate-determining step in the catalytic cycle. The computations
show that this proposed catalytic cycle is indeed feasible as the highest point in the
catalytic cycle represented by TS1 is only 9.6 kJ/mol (or 2.3 kcal/mol) above
Re(CO)4SiH3 and CH2O, the starting pair of the cycle. Thus, once the rhenium silyl
species has been generated, the catalysis should proceed readily even at room
temperature.
35
Complex
Structures for CH2O
substrate
H3SiRe(CO)4
Intermediate I
Intermediate II
Intermediate III
TS I
Enthalpy H
Free energy G
CO or i
-535.751262
-535.806047
1767, 1807, 1809, 1908
-649.819307
-649.791523
-655.891354
-649.880743
-649.85245
-655.960371
-649.783078
-649.840337
-655.882012
-655.950457
1775, 1784, 1818, 1908
1779, 1792, 1820, 1916
1767, 1784, 1808, 1921
1789, 1799, 1817, 1900
i = 281
1797, 1820, 1832, 1921
i = 798
-763.804788
-763.868593
-727.961207
-728.02476
-879.689834
-879.758904
TS II
Transition states for
different substrates
TS I for HCOOCH3
TS I for (CH3)2CO
TS I for PhCHO
1781, 1793, 1821, 1907
i = 291
1777, 1793, 1805, 1900
i = 279
1780, 1792, 1809, 1902
i = 265
Ea
9.6
67.7
24.0
2.0
Table 4. Calculated enthalpies (Hartree), free energies (Hartree), carbonyl vibrational and
imaginary frequencies (cm-1), and the activation barrier to TS1 (kJ/mol, relative to
Re(CO)4SiH3 and respective carbonyl) for the rhenium intermediates and transition states of
the catalytic cycle. The silane used is SiH4. Method and basis set used : mp2/lanl2dz.
The relative hydrosilylation rates can be explained by comparing the activation
barrier of the rate-determining step for the different substrates. We have focused on
optimising the respective transition states TS1 of four representative carbonyls;
formaldehyde, benzaldehyde, acetone and methyl formate. The value of the activation
barrier is calculated relative to the starting pair in the cycle i.e. Re(CO)4SiH3 and the
corresponding carbonyl. The various transition states together with the essential
molecular parameters have been depicted in Figure 5.
Comparing the values of Ea in Table 4, it can be seen that the barriers for the
aldehydes are lower than the ketone which itself is much lower than the ester. In the
36
ester case, electronic effects due to the OR group destabilise the transition state much
more than the aldehyde or ketone counterpart. Interestingly, the calculated activation
energy for the benzaldehyde system is the lowest despite experimental kinetic studies
suggesting a faster reaction for aliphatic aldehydes. Although the difference is slight
(~7 kJ/mol), the inclusion of solvent effects may be able to reverse this trend.
37
2.61A
2.21A
Intermediate 1
2.09A
1.64A
1.79A
Intermediate 2
2.26A
Intermediate 3
Figure 4. Calculated structures of intermediates and transition states for Scheme
6.
Essential bond lengths (Ao) are shown.
38
2.61A
Re(CO)4(SiH3)
2.80A
2.05A
2.24A
TS 1
2.31A
2.63A
1.66A
2.39A
TS 2
Figure 4 continued.
39
2.05
A
2.80
A
2.24
A
2.83
A
TS (Formaldehyde)
2.01
A
2.26
A
TS (Acetone)
2.82
A
2.04
A
2.23
A
2.84
A
2.00
A
2.26
A
TS (Methyl Formate)
TS (Benzaldehyde)
Figure 5. Optimised transition states for the rate-determining steps of different substrates.
Bond lengths in Ao
40
2.3 Conclusion
Re(CO)5Cl has been found to catalyse the hydrosilylation reaction between various
silanes and carbonyl substrates such as aldehyde, ketone, ester and carbonate
effectively with a turnover frequency of 20 to 25 hr-1 for aldehydes. The rates of
hydrosilylation are highest for the least sterically-hinderd silanes with aldehydes,
followed by aliphatic ketone. Aromatic ketone, ester and carbonate react slowest with
the silanes. A rhenium dimer complex HRe2(CO)9 has been detected using IR
spectroscopy and is believed to be the resting state for the catalysis. Upon its
photolysis, the active catalyst Re(CO)4SiH3 is released and participate in the catalytic
cycle. Computational studies have shown that the proposed cycle can be carried out
readily at room temperature and are able to explain the relative hydrosilylation rates
among the various carbonyl substrates.
2.4 Experimental Section
All chemicals were used as purchased. All reactions were carried out under vacuum
conditions unless otherwise stated. Proton nuclear magnetic resonance (1H NMR)
spectra were obtained on a Bruker Avance 500 (AV500) spectrometer at room
temperature. The chemical shifts were recorded relative to tetramethylsilane for
spectra taken in CDCl3.
Infrared (IR) spectra were obtained on a Shimadzu IR Prestige-21 Fouriertransformed infrared spectrometer (1000-4000cm-1, 1cm-1 resolution, 4 scans co-
41
added for spectra averaging) using a 0.05mm path-length CaF2 cell. Mass spectra of
the organic products are recorded with a Finnigan Mat 95XL-T spectrometer.
Catalysis using Re(CO)5Cl: Re(CO)5Cl (0.011mmol) was mixed with Et3SiH
(0.33mmol) and the carbonyl substrate (0.11mmol) in a quartz tube. The mixture was
photolysed in vacuo, using a Legrand Broadband Lamp (200-800nm, 11W) for 4
hours. Products were identified using 1H NMR.
Catalytic reduction of diethylcarbonate: Re(CO)5Cl (0.011mmol) was mixed with
Et3SiH (0.33mmol) and diethylcarbonate (0.11mmol) in a quartz tube. The mixture
was photolysed in vacuo, using a Legrand broadband Lamp (200-800nm, 11W) for 4
hours. 1H NMR (CDCl3): Triethylethoxysilane (1) - 0.60(m), 0.98(m), 1.20(m),
3.65(m); Ethylformate (2) - 1.31 (t), 4.24 (q), 8.04 (s); Diethoxymethane (4) - 0.60(m),
0.98(m), 1.20(m), 3.65(m), 4.86(m).
Synthesis of HRe(CO)5: HRe(CO)5 was synthesized via a modification of reported
methods.22-24 0.011mmol of Re(CO)5Cl was dissolved in methanol and the mixture was
stirred for 1hr after which the mixture was cooled down using liquid nitrogen. Excess
Zinc and acetic acid were then added to the cooled mixture before evacuation of the
air inside the rbf was done. The reaction was then left to stir for 24hr at room
temperature. The HRe(CO)5 formed was extracted using 3x3mL of hexane and
recrystallised using CHCl3. CO(Cyclohexane): 2015(vs), 2005(m), 1983(w); 1H
NMR: -5.77ppm
42
Comparison of rate of hydrosilylation reaction between different substrates:
Re(CO)5Cl (0.011mmol) was mixed with Et3SiH (0.33mmol) and two different
carbonyl substrates (0.11mmol each) in a quartz tube. The mixture was photolysed in
vacuo, using a Legrand Broadband Lamp (200-800nm, 11W) for 4 hours. Products
were identified using 1H NMR and yields were calculated with toluene (0.011mmol) as
standard.
Synthesis of HRe2(CO)9(SiR3) from Re(CO)5Cl: Re(CO)5Cl (0.011mmol) and
R3SiH (0.022mmol) were dissolved in cyclohexane in a quartz tube. The mixture was
photolysed in vacuo, using a Legrand Broadband Lamp (200-800nm, 11W) until all
(CO) of Re(CO)5Cl have disappeared (2-4hours). The yellow solution was
concentrated and passed through a column using a 1:1 mixture of hexane and
chloroform. The resultant yellow band was then collected and the solvent removed
using vacuum and recrystallised in hexane solution to afford a yellow powder (Yield
with respect to Re(CO)5Cl is 65%). The products were identified based on comparing
their IR and 1H NMR spectra to reported values.15
HRe2(CO)9(SiEt3)
CO(Cyclohexane):
2030(m),
2021(s),
2016(sh),
2008(m),
2000(m), 1982(sh), 1978(s); 1H NMR: -9.03ppm; UV(nm): 200-400 nm (broad with
max = 218 nm), ESI mass spectrum : m/z = 742
HRe2(CO)9(SiPh2H) CO(Cyclohexane): 2029(m), 2020(s), 2014(sh), 2008(m),
2000(m), 1981(sh), 1976(s) ; 1H NMR: -9.71ppm
Synthesis of HRe2(CO)9(SiEt3) from Re2(CO)10: Re2(CO)10 (0.011mmol) and Et3SiH
(0.022mmol) were dissolved in cyclohexane in a quartz tube. The mixture was
photolysed in vacuo, using a Legrand Broadband Lamp (200-800nm, 11W) until all
43
(CO) of Re(CO)5Cl have disappeared (2-4hours). The yellow solution was
concentrated and passed through a column using a 1:1 mixture of hexane and
chloroform. The resultant yellow band was then collected and the solvent removed
using vacuum and recrystallised in hexane solution to afford a yellow powder (Yield
with respect to Re(CO)5Cl is 68%).
Silyl exchange between HPh2SiRe(CO)4 and Et3SiH: Et3SiH (0.22mmol) and
Acetone
(0.11mmol)
were
added
to
a
cyclohexane
solution
containing
HRe2(CO)9(SiR3) (0.011mmol) in a quartz tube. The mixture was photolysed in vacuo
and an IR spectrum of the reaction was taken at every 30 minute interval.
2.5 References
[1]
Abu-Omar, M. M.; Du, G. Organometallics 2006, 25, 4920
[2]
Morris, R.H. Chem. Soc. Rev. 2009, 38, 2282
[3]
Shaikh, N. S.; Junge, K.; Beller, M. Org. Lett. 2007, 9, 5429
[4]
Fernandes, A.C.; Fernandes, R.; Romão, C.C; Royo, B. Chem. Comm. 2005,
213
[5]
Gądek, A.; Szymańska-Buzar, T. Polyhedron 2006, 25, 1441
[6]
Son, S. U.; Paik, S.; Chung, Y. K. J. Mol. Catal. A: Chem. 2000, 151, 87
[7]
Cavanaugh, M.D.; Gregg, B.T.; Cutler, A.R. Organometallics 1996, 15, 2764
[8]
Dong, H.; Berke, H. Adv. Synth. Catal. 2009, 351, 1783
[9]
Ojima, I.; Nihonyanagi, M.; Kogure, T.; Kumagai, M.; Horiuchi, S.;
Nakatsugawa, K. J. Organomet. Chem. 1975, 94, 449
44
[10] Wile, B. M.; Stradiotto, M. Chem. Commun. 2006, 4104
[11] Do, Y.; Han, J.; Rhee, Y. H.; Park, J. Adv. Synth. Catal. 2011, 353, 3363
[12] Chung, L. W.; Lee, H. G.; Lin, Z.; Wu, Y. J. Org. Chem. 2006, 71, 6000
[13] Du, G.; Fanwick, P. E.; Abu-Omar, M. M. J. Am. Chem. Soc. 2007, 129, 5180
[14] Sharma, H. K.; Pannell, K. H. Angew. Chem. Int. Ed. 2009, 48, 7052
[15] Mao, Z.; Gregg, B. T.; Cutler, A. R. J. Am. Chem. Soc. 1995, 117, 10139
[16] Ojima, Y.; Yamaguchi, K.; Mizuno, N. Adv. Synth. Catal. 2009, 351, 1405
[17]
Gevorgyan, V.; Rubin, M.; Liu, J.; Yamamoto, Y. J. Org. Chem. 2001, 66,
1672
[18] Hua, R.; Jiang, J. Cur. Org. Synth. 2007, 4, 151
[19] Wrighton, M S.; Morse, D. L.; Gray, H. B.; Ottessen, D. K. J. Am Chem. Soc.
1976, 98, 1111
[20] Hoyano, J. K.; Graham, W. A. G. Inorg. Chem. 1972, 11, 1265
[21]
Frisch, M. J.; et al. Gaussian 03, ReVisions B. and B.05; Gaussian Inc.:
Wallingford, CT, 2004.
[22]
Braterman, P. S.; Harrill, R. W.; Kaesz, H. D. J. Am. Chem. Soc. 1967, 89,
2851
[23] Byers, B. H.; Brown, T. L. J. Am. Chem. Soc. 1977, 99, 2527
[24] Li, N.; Xie, Y.; King, R. B.; Schaefer III, H. F. J. Phys. Chem. A. 2010, 114,
11670
45
Chapter 3
Ruthenium carbonyl-catalyzed Si-X coupling
(X = S, O, N)
3.1 Introduction
Organosilanes containing Si-H bonds have been used extensively in synthetic
organic chemistry for many years. Hydrosilylation of unsaturated bonds such as
alkenes1, alkynes2 and carbonyls3 is one of the more important processes involving
organosilanes. Transition metal complexes containing iron,4-5(re
paper)
tungsten,6
molybdenum,7 manganese,8-9 rhenium,10-11 rhodium,12 silver,13 nickel,14 and most
recently ruthenium.15-16, 2 have been used to catalyze such processes.
Silane hydrolysis and alcoholysis forming silanol, silyl ether and siloxane17
demonstrate further the versatility of organosilanes as valuable substrates. Ruthenium
carbonyl complexes have also been reported to generate hydrogen from a
silane/water mixture17-18 and from formic acid.19 The hydrogen gas produced can be
used in a variety of reactions, of which one of the most well-studied reaction involves
the hydrogenation of unsaturated compounds catalyzed by palladium (II)20 or iridium
(I)21 complexes. These transformations are made possible because of the relatively
weaker Si-H bonds available for activation as compared to C-H bonds.
Silicon coupling with other heteroatoms such as sulfur and nitrogen are, however,
less widely studied. Besides organosilanes, organosulfur derivatives of silanes can
also be used to protect carbonyl groups selectively.22 However, methods of their
46
preparation require air-sensitive catalysts such as B(C6H5)3,23 and in some cases,
sulfur poisoning is observed leading to much lower efficiency achieved in the catalytic
system.24
The various couplings of Si-N bonds from silanes and amines resulting in the
production of iminosilanes16a, silazanes16b and diaminosilanes16c have also been
reported. However, these reactions were not extensively explored although they
provide much insight into the role of amine ligands in processes such as alkene
polymerisation25 and alkene metathesis26. Of recent interest are the formation of
direct Si-N bonds via the addition of silanes to metal-imido27-30 (M=NR) complexes,
which provides novel pathways to synthesize potential catalysts for metathesis
reactions.
Ruthenium dodecacarbonyl Ru3(CO)12 has been reported to activate Si-H bonds31 to
form dimeric ruthenium silyl complex of formula (R3Si)Ru-Ru(SiR3). Chatani et. al.32
reported the activation of benzylic C-H bonds by Ru3(CO)12 in the presence of silanes
to afford benzylsilanes, while Madalska et. al.33 was able to utilise Ru3(CO)12 to
catalyse the silylation of silanols with vinylsilanes to afford alkenes and siloxanes.
In this work, we have found that Ru3(CO)12 acts as a versatile and efficient catalyst
for Si-heteroatom (S, N and O) coupling with high turnover numbers. The scope of the
catalysis as well as the kinetics and detection of reactive intermediates have been
investigated
using
FTIR
and
NMR
spectroscopy.
A
ruthenium
dimer,
[Ru(CO)4(SiEt3)] 2, has been identified as the likely resting state of the catalyst and a
mechanism for silane-thiol coupling based on this dimer has been proposed.
47
Computational chemistry using the Gaussian suite of programs has also been
employed for elucidating the reaction mechanism.
3.2 Results and Discussion
A preliminary study using a 1:1 mixture of Et3SiH and dodecanethiol as the
substrates has been carried out in order to optimise the experimental conditions for
efficient coupling (see Table 1). A temperature of 80oC over 4 hours using 1% mol
loading of Ru3(CO)12 was sufficient for a full conversion (turnover number TON =
100, turnover frequency = 25 hr-1) of the substrates into the silylthioether, C12H25S–
SiEt3 (Table 1, Entry 1).
Complex
Experimental conditions
TON
TOF/hr-1[a]
Ru3(CO)12
80oC, 4 hrs
100
25
80oC, 2 hrs
65
33
80oC, 8 hrs
100
25
80oC, 4 hrs, vacuum
100
25
25oC, 4 hrs
32
8
80oC, 8 hrs, 0.5% mol loading
180
23
25oC, 4 hrs, 300-800 nm photolysis
22
6
80oC, 4 hrs, silane/thiol = 1.5
100
25
80oC, 4 hrs, thiol/silane = 1.5
50
13
Ru2(CO)4(PPh3)2Br4
80oC, 4 hrs
55
15
Ru2(CO)6Br4
80oC, 4 hrs
52
13
52
13
Ru2(CO)6I4
o
80 C, 4 hrs
Table 1: Optimization of the experimental conditions for the coupling of triethylsilane with dodecanethiol
using [Ru] catalyst. Unless stated otherwise, the reactions were carried out in air with 1% catalyst mol
loading using 1:1 silane/thiol mol ratio in toluene solvent.
48
No coupling was observed in the absence of Ru3(CO)12 while heating the reaction
mixture to higher than 80°C did not increase the coupling rate significantly. Full
conversion can only be achieved after 24 hrs under lower temperature conditions
( 35 hr-1 compared to
the reverse case of using more thiol to silane. As depicted in Scheme 1, hydrogen gas
is expected to be generated and indeed, on-line mass spectrometric detection of the
headspace above the solution mixture has recorded its presence at m/z = 2 (Figure 1).
By calibration with a known H2 pressure, the yield of the gaseous product has been
determined to be 85%±20% compared to the silylthioether.
49
Figure 1. Gaseous mass spectra taken of the headspace content of the coupling
reaction between Et3SiH and dodecanethiol catalysed by Ru3(CO)12. The peak
labelled 'After reaction' corresponds to m/z = 2.
As haloruthenium carbonyls such as Ru2(CO)4(PPh3)2Br4, Ru2(CO)6Br4 and
Ru2(CO)6I4 are known to activate Si-H bonds18, we have prepared these complexes
using reported methods34 for the Si-S coupling reaction. However the coupling is
found to proceed less effectively for each of them compared to Ru3(CO)12. Based on
the optimization studies, the catalytic conditions have been fixed at 80oC, 4 hours and
1% mol loading of Ru3(CO)12 for further studies with different substrates.
50
Entry
Substrate
[a]
Product
TON[b]
TOF/ hr1[c]
C12H25SH
C12H25S – SiEt3
100
25
2
C8H17SH
C8H17S – SiEt3
100
25
3
PhCH2SH
PhCH2S – SiEt3
100
25
4
p-Me(C6H4)SH
p-Me(C6H4)S – SiEt3
100
25
5
C2H5OH
C2H5O – SiEt3
100
25
6
iPrOH
iPrO – SiEt3
100
25
7
C5H11OH
C5H11O – SiEt3
100
25
8
C6H5OH
C6H5O – SiEt3
100
25
9
C8H17NH2
C8H17NH – SiEt3
100
25
10
PhNH2
PhNH – SiEt3
100
25
11
PhMeNH
PhMeN – SiEt3
100
25
C12H25SH
C12H25S–SiHPh2
78
13
13
PhCH2SH
PhCH2S–SiHPh2
69
12
14
p-Me(C6H4)SH
15
C2H5OH
C2H5O–SiHPh2
70
12
16
C5H11OH
C5H11O–SiHPh2
65
11
17
PhNH2
PhNH–SiHPh2
58
10
C12H25SH
C12H25S–SiPh3
65
9
19
PhCH2SH
PhCH2S– SiPh3
57
8
20
p-Me(C6H4)SH
p-Me(C6H4)S– SiPh3
62
9
21
C2H5OH
C2H5O– SiPh3
55
8
22
C5H11OH
C5H11O– SiPh3
51
7
23
PhNH2
PhNH– SiPh3
35
5
1
12
18
Et3SiH
Ph2SiH2
Ph3SiH
p-Me(C6H4)S–
SiHPh2
73
12
Table 2. Coupling reactions between different silanes and various substrates catalysed by 1 mol%
Ru3(CO)12 at 80oC, 4 hrs.
[a]
Products identified using 1H NMR.
[b]
TON based on NMR integration.
[c]
TOF
based on reaction time: Et3SiH (4 hrs), Ph2SiH2(6 hrs) and Ph3SiH (7 hrs).
51
Table 2 shows the various coupling reactions of triethylsilane that have been
carried out with thiol, amine and alcohol substrates. In general, all three types of
coupling Si-S, Si-N and Si-O take place with similar TOF efficiency under the same
conditions. Based on the TOF, it was found that thiols undergo silane coupling
slightly more effectively, followed by alcohols and amines, and with the aliphatic
substrates being more reactive than aromatic substrates. When bulkier silanes were
used, a 2-3 times decrease in the rate of coupling with all three types of substrates
was observed. Full conversion can still be achieved for all substrates upon prolonging
the reaction time to 6 hrs for Ph2SiH2 and a minimum of 7 hours for Ph3SiH.
3.2.1 Reactive intermediates and Mechanism
The detection and identification of reactive intermediates using FTIR, NMR and
mass spectroscopy have been carried out in order to provide some insights into the
mechanism of the coupling reactions. We have chosen the Si-S coupling with
dodecanethiol and triethylsilane substrates as the model reaction. For this purpose,
the concentration of the ruthenium catalyst has been scaled up in order to generate
more intermediates for characterization. The FTIR spectrum taken of the reaction
mixture containing 5%mol Ru3(CO)12 and a 1:1 silane:thiol mixture heated for 3
hours at 80oC is shown in Figure 2.
0.9
Abs
0.8
0.7
0.6
52
53
Two new sets of peaks have appeared at the expense of the initial Ru3(CO)12 signals
at 2061, 2029 and 2009cm-1. The first set consisting of a strong CO band at 2017cm-1
and 2 weaker peaks at 2041cm-1 and 2006cm-1 has been assigned to a dimeric
ruthenium silyl carbonyl complex, [Ru(CO)4(SiEt3)] 2 based on previous literature.31
Each ruthenium atom carries a silyl ligand which is oriented trans to each other
(Scheme 2). The reduced number of carbonyl bands observed in the spectrum is
indicative of the highly-symmetrical nature of the complex. An ESI mass spectrum
recorded of the mixture shows a cluster of peaks centred at m/z=659 which matches
the molecular ion mass of the ruthenium dimer.
OC
OC
Ru 3(CO) 12
+ Et3SiH
Et3Si
OC
Ru
OC
OC
CO
Ru
CO
CO
SiEt3
CO
+
OC
OC
Ru
CO
CO
Ru
CO CO
Ru
OC H
CO
H
Scheme 2. Ru2(CO)8(SiEt3)2 and H2Ru3(CO)10 complexes
The second set of CO peaks (2082cm-1, 2067cm-1, 2031cm-1, 2013cm-1) are due to a
ruthenium carbonyl hydride cluster H2Ru3(CO)10.35 A 1H NMR spectrum recorded of
the mixture showed a hydride signal at -14.3ppm, which is attributed to the equivalent
bridging hydrides of the cluster. Similarly the mass spectrum showed the expected
molecular ion cluster of peaks at m/z=559 together with the fragmentation patterns
consistent with sequential losses of the CO ligands.
Indeed, both the [Ru(CO)4(SiEt3)] 2 and H2Ru3(CO)10 complexes can also be
generated in the reaction between Ru3(CO)12 with silane only, independent of any
thiols in the solution. This result also implied that the hydrogen atoms of the hydride
54
CO
cluster originate from the silane substrate. The detailed mechanism of the formation
of both of these complexes from is probably very complicated but most likely involves
the initial dissociation of Ru3(CO)12 into reactive Ru(CO)4 monomeric units. Both
complexes could represent the eventual thermodynamically-stable ruthenium silylated
and hydride products upon reactions of Ru(CO)4 with the silane.
While the [Ru(CO)4(SiEt3)] 2 can be purified in small yields ([...]... HRe(CO)5 1 .2 Abs 1.05 0.9 0.75 0.6 0.45 0.3 0.15 -0 21 15 21 00 20 85 20 70 20 55 20 40 20 25 20 10 1995 1980 Cy 1965 1950 1/cm 21 00 20 80 20 60 20 40 20 20 20 00 1980 Figure 3 IR and 1H Hydride NMR Spectrum of (A) obtained from photolysis of Re(CO)5Cl with Et3SiH IR peaks (cm-1): 21 02, 20 94, 20 85, 20 30, 20 26, 20 20, 20 07, 20 00, 1978 Lit Values13 for HRe2(CO)9SiCl3(cm-1): 21 50, 20 95, 20 85, 20 47, 20 19, 20 12, 1999,... 1908 1779, 17 92, 1 820 , 1916 1767, 1784, 1808, 1 921 1789, 1799, 1817, 1900 i = 28 1 1797, 1 820 , 18 32, 1 921 i = 798 -763.804788 -763.868593 - 727 .96 120 7 - 728 . 024 76 -879.689834 -879.758904 TS II Transition states for different substrates TS I for HCOOCH3 TS I for (CH3)2CO TS I for PhCHO 1781, 1793, 1 821 , 1907 i = 29 1 1777, 1793, 1805, 1900 i = 27 9 1780, 17 92, 1809, 19 02 i = 26 5 Ea 9.6 67.7 24 .0 2. 0 Table 4... in vacuo 24 Entry Substrate Silane Time Product TOF[b](hr1 [a] ) 25 1 OSiPh3 O Ph3SiH 10hrs (2a) 1 H OSiPh3 O 2 H Ph3SiH 10hrs (2b) 1.5 OSiPh2H O 3 Ph2SiH2 5hrs (2c) 8 (2d) 10 (2e) 8 (2f) 14 H OSiPh2H O 4 H Ph2SiH2 5hrs OSiMe2Ph O Me2PhSiH 5 5hrs H O 6 OSiMe2Ph H Me2PhSiH 5hrs Table 2 The reactions of Acetone and Benzaldehyde with various silanes in ratio of silane : carbonyl = 3:1 [a] TON calculated... Organomet Chem 20 05, 690, 3774 [22 ] Hua, R.; Tian, X J J Org Chem 20 04, 69, 57 82 [23 ] Ojima, Y.; Yamaguchi, K.; Mizuno, N Adv Synth Catal 20 09, 351, 1405 [24 ] Hua, R.; Jiang, J Cur Org Synth 20 07, 4, 151 [25 ] Murahashi, S -I Ruthenium in Organic Synthesis, Wiley-VCH:Weinheim 20 04 [26 ] Renaud, J -L.; Demerseman, B.; Mbaye, M D.; Bruneau, C Cur Org Chem 20 06, 10, 115 [27 ] Bruneau, C.; Dixneuf, P.H Ruthenium. .. Organometallics 20 08, 27 , 24 90 [9] Hong, S H.; Sander, D P.; Lee, C W.; Grubbs, R H J Am Chem Soc 20 05, 127 , 17160 [10] Kuninobu, Y.; Takai, K Chem Rev 20 11, 111, 1938 [11] Hua, R.; Jiang, J -L Maitlis, P M Curr Org Synth 20 07, 4, 151 [ 12] Kühn, F E.; Santos, A M.; Herrmann, W A Dalton Trans 20 05, 24 83 [13] Espenson, J H Adv Inorg.Chem 20 03, 54, 157 [14] Kusama, H.; Narasaka, K Bull Chem Soc Jpn 1995, 68, 23 79... intermediates and explore the mechanism of hydrosilylation using a combination of IR and NMR spectroscopy as well as computational studies 20 O O + R3SiH Me 2N OSiR3 [CpFe(CO)2Me] H r.t., h Me 2N R = H, Alkyl, Aryl O PhSiH3 + O R 3% (L)(CO)4MnC(O)CH3 C6D6 OEt O Et3SiH OH H B(C6F5)3, r.t R R3SiH Me 2N OSiEt 3 PhSiH3 SiR3 + H CH3 O(SiH 2Ph )2 OEt H + OEt OSiEt3 Et3SiH OSiEt3 R3Si 2 Et3SiH R CH3 R H OSiEt3 - 2O(SiEt3 )2. .. Catalysts and Fine Chemistry, Springer: Berlin, 20 04 [28 ] Tan, S T.; Kee, J W.; Fan, W Y.; Organometallics 20 11, 30, 4008 [29 ] Naota, T.; Takaya, H.; Murahashi, S -I Chem Rev 1998, 98, 25 99 [30] Nishibayashi, Y.; Uemura, S Cur Org Chem 20 06, 10, 135 [31] Trost, B M.; Toste, D.; Pinkerton, A B Chem Rev 20 01, 101, 20 67 [ 32] Zhao, X.; Yu, Z.; Yan, S.; Wu, S.; Liu, R.; He, W.; Wang, L J Org Chem 20 05, 70,... iron ,2- 3 tungsten,4 molybdenum,5 manganese,6-7 rhenium, 1,8 rhodium,9 silver10 and most recently ruthenium. 11 Recent works were also done on the mechanistic studies of hydrosilylation using oxo -rhenium complexes. 12- 13 While Berke et al.8 has reported the use of different rhenium (I) catalysts for the hydrosilylation of aldehydes and ketones, one of the more interesting processes is the CpFe(CO)2Me-catalysed... cross-coupling reactions involving organosilanes also provide many useful applications in organic and bioorganic chemistry Transition metal -catalyzed silane hydrolysis and alcoholysis involving the reaction between organosilanes and water/alcohol produces silanols, silyl ethers and siloxanes,50 which provide useful starting materials for the manufacture of pharmaceuticals51 and also as excellent research... K J Am Chem Soc 20 09, 131, 9914 15 [16] Zuo, W -X.; Hua, R.; Qiu, X Synth Commun 20 04, 34, 321 9 [17] Kuninobu, Y.; Kawata, A.; Yudha, S S.; Takata, H.; Nishi, M.; Takai, K Pure Appl Chem 20 10, 82, 1491 [18] Chen, H.; Hartwig, J F Angew Chem Int Ed 1999, 38, 3391 [19] Kuninobu, Y.; Kawata, A.; Takai, K J Am Chem Soc 20 05, 127 , 13498 [20 ] Zhao, W -G.; Hua, R Eur J Org Chem 20 06, 5495 [21 ] Ouh, L L.; Müller, ... 38 2. 61A Re(CO)4(SiH3) 2. 80A 2. 05A 2. 24A TS 2. 31A 2. 63A 1.66A 2. 39A TS Figure continued 39 2. 05 A 2. 80 A 2. 24 A 2. 83 A TS (Formaldehyde) 2. 01 A 2. 26 A TS (Acetone) 2. 82 A 2. 04 A 2. 23 A 2. 84 A 2. 00... Re(CO)5Cl with Et3SiH IR peaks (cm-1): 21 02, 20 94, 20 85, 20 30, 20 26, 20 20, 20 07, 20 00, 1978 Lit Values13 for HRe2(CO)9SiCl3(cm-1): 21 50, 20 95, 20 85, 20 47, 20 19, 20 12, 1999, 1978 H NMR peaks (ppm):... attributed to HRe(CO)5 1 .2 Abs 1.05 0.9 0.75 0.6 0.45 0.3 0.15 -0 21 15 21 00 20 85 20 70 20 55 20 40 20 25 20 10 1995 1980 Cy 1965 1950 1/cm 21 00 20 80 20 60 20 40 20 20 20 00 1980 Figure IR and 1H Hydride NMR