ArticlePDF Available

Platinum-Catalyzed Reactions of 2,3-Bis(diisopropylsilyl)thiophene with Alkynes

Authors:

Abstract and Figures

The reactions of 2,3-bis(diisopropylsilyl)thiophene (1) with diphenylacetylene, phenylacetylene, trimethylsilylacetylene, and mesitylacetylene have been reported. The reactions of 1 with diphenylacetylene and phenylacetylene in the presence of a catalytic amount of tetrakis(triphenylphosphine)platinum(0) at 80 °C gave [1,4]disilino[2,3-b]thiophene derivatives. With trimethylsilylacetylene, 1 afforded two types of products arising from sp-hybridized C–H bond activation of the acetylene, together with [1,3]disilolo[4,5-b]thiophene derivatives. A similar treatment of 1 with mesitylacetylene produced two regioisomers of products arising from the C–H bond activation of mesitylacetylene. Theoretical calculations for the intramolecular reactions of 10a and 10b are also discussed.
Content may be subject to copyright.
Platinum-Catalyzed Reactions of 2,3-Bis(diisopropylsilyl)thiophene
with Alkynes
Akinobu Naka,*
,
Takashi Mihara,
Hisayoshi Kobayashi,*
,
and Mitsuo Ishikawa
§
Department of Life Science, Kurashiki University of Science and the Arts, Nishinoura, Tsurajima-cho, Kurashiki, Okayama 712-8505,
Japan
Professor Emeritus Kyoto Institute of Technology, Matsugasaki, Kyoto 606-8585, Japan
§
Professor Emeritus Hiroshima University, Higashi-Hiroshima 739-8527, Japan
*
SSupporting Information
ABSTRACT: The reactions of 2,3-bis(diisopropylsilyl)-
thiophene (1) with diphenylacetylene, phenylacetylene,
trimethylsilylacetylene, and mesitylacetylene have been
reported. The reactions of 1with diphenylacetylene and
phenylacetylene in the presence of a catalytic amount of
tetrakis(triphenylphosphine)platinum(0) at 80 °C gave [1,4]-
disilino[2,3-b]thiophene derivatives. With trimethylsilylacety-
lene, 1aorded two types of products arising from sp-
hybridized CH bond activation of the acetylene, together
with [1,3]disilolo[4,5-b]thiophene derivatives. A similar treat-
ment of 1with mesitylacetylene produced two regioisomers of products arising from the CH bond activation of
mesitylacetylene. Theoretical calculations for the intramolecular reactions of 10a and 10b are also discussed.
1. INTRODUCTION
It is well-known that reactions of organosilanes with unsaturated
compounds in the presence of a catalytic amount of transition-
metal complexes can be employed for the synthesis of various
organosilicon compounds.
1
In these reactions, the silyl-
transition-metal complexes would be formed as reactive species.
For example, the benzodisilaplatinacyclopentene derivatives are
produced by reactions of 1,2-bis(hydrosilyl)benzenes with
platinum complexes, and their chemical properties have been
widely examined by Tanaka et al., Nagashima et al., and other
chemists.
28
We have reported that the platinum-catalyzed reactions of
tetraethyl-substituted benzodisilacyclobutene with alkynes such
as diphenylacetylene, 3-hexyne, and phenylacetylene readily
produce the respective 1:1 adducts, benzodisilacyclohexadiene
derivatives, in high yield.
911
On the other hand, platinum-
catalyzed reactions of tetraisopropyl-substituted benzodisilacy-
clobutene with the same alkynes aord two types of products,
benzodisilacyclohexadienes and benzodisilacyclopentenes.
12,13
Recently, we have demonstrated that the platinum-catalyzed
reactions of 2,3-bis(diethylsilyl)thiophene with alkynes such as
diphenylacetylene, 3-hexyne, and phenylacetylene aord the
respective [1,4]disilino[2,3-b]thiophenes.
14
In these reactions,
[1,2,5]platinadisilolo[3,4-b]thiophene derivatives would be
formed as the reactive intermediates.
It is of considerable interest to us to investigate the chemical
behavior of 2,3-bis(silyl)thiophene with bulky substituents on
the silicon atoms toward alkynes in the presence of a platinum
catalyst. In this article, we report the platinum-catalyzed reactions
of 2,3-bis(diisopropylsilyl)thiophene with mono- and disub-
stituted alkynes.
2. RESULTS AND DISCUSSION
The starting compound, 2,3-bis(diisopropylsilyl)thiophene (1),
was prepared by the reaction of 2-bromo-3-iodothiophene with
magnesium and chlorodiisopropylsilane in the presence of a
catalytic amount of copper(I) cyanide in tetrahydrofuran (THF)
(Scheme 1).
15
We rst examined the reactions of 1with diphenylacetylene.
Treatment of compound 1with diphenylacetylene in the
presence of Pt(PPh3)4in reuxing benzene for 96 h gave
1,1,4,4-tetraisopropyl-2,3-diphenyl-1,4-dihydro-[1,4]disilino-
[2,3-b]thiophene (2) in 80% yield (Scheme 2). The structure of
2was veried by spectroscopic analysis. The mass spectrum for 2
shows parent ions at m/z488, corresponding to the calculated
molecular weight of C30H40Si2S. The 1H NMR spectrum of 2
Received: October 24, 2017
Accepted: November 21, 2017
Published: November 30, 2017
Scheme 1. Synthesis of Compound 1
Article
Cite This: ACS Omega 2017, 2, 85178525
© 2017 American Chemical Society 8517 DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
This is an open access article published under an ACS AuthorChoice License, which permits
copying and redistribution of the article or any adaptations for non-commercial purposes.
shows septet signals at 1.33 and 1.34 ppm, attributable to the
methine protons in the isopropyl groups at silicon; four doublet
signals at 0.80, 0.81, 1.04, and 1.09 ppm, attributable to the
methyl protons of the isopropyl groups; two doublet signals at
7.39 and 7.76 ppm, attributable to the thienylene protons; and
signals attributed to the phenyl protons. The 29Si NMR spectrum
for 2shows signals at 10.6 and 9.2 ppm.
On the basis of the results obtained from the platinum-
catalyzed dehydrogenative double silylation
26,14
of various
unsaturated compounds with bis(hydrosilane)s, the formation of
adducts 2obtained from the platinum-catalyzed reaction of 1
with diphenylacetylene may be explained as follows: (1)
oxidative addition of the platinum species to one of the SiH
bonds in 1; (2) dehydrogenation to give the bis(silyl)platinum
complex (3); (3) insertion of a triple bond of alkynes into a
platinumsilicon bond to produce the platinum complex (4);
(4) reductive elimination of 2from the Pt complex (4)(Scheme
3).
Unfortunately, compound 1did not react with phenyl-
trimethylsilylacetylene and bis(trimethylsilyl)acetylene. The
starting compound 1was recovered quantitatively.
Next, we investigated the platinum-catalyzed reaction of 1with
monosubstituted alkynes under the same conditions as described
above. Treatment of 1with phenylacetylene at 80 °C indicated
that 1:1 cyclic adducts as a mixture of two regioisomers, namely,
1,1,4,4-tetraisoproyl-3-phenyl-1,4-dihydro-[1,4]disilino[2,3-b]-
thiophene (5a) and 1,1,4,4-tetraisopropyl-2-phenyl-1,4-dihydro-
[1,4]disilino[2,3-b]thiophene (5b), were produced in a ratio of
3:1 in 60% yield, along with unreacted starting compound 1
(40%) (Scheme 4). All attempts to separate 5a and 5b were
unsuccessful, but their structures were veried by spectroscopic
analysis of the mixture.
The regioselective formation of 5a and 5b in the platinum-
catalyzed reaction of 1with monosubstituted alkynes can be best
explained by a series of reactions similar to the reaction shown in
Scheme 3. Namely, the regioselective addition of a PtSi bond in
the intermediate similar to 3to the carboncarbon triple bond of
monosubstituted acetylene aords predominantly 1,1,5,5-
tetraisopropyl-3-phenyl-1,5-dihydro-[1,2,5]platinadisilepino-
[3,4-b]thiophene as the intermediate. It has been reported that
the introduction of electron-withdrawing substituents on the
silicon atom in the silylplatinum complexes enhances the PtSi
bond.
14,16,17
In the present reactions, the PtSi bond of the Pt
SiCC moiety is probably weaker than that of the PtSiCS
moiety (sulfur has higher electronegativity than carbon) and the
former migrates to the less hindered carbon atom in the
coordinated monosubstituted alkyne to produce insertion
product 5a.
When a mixture of 1and trimethylsilylacetylene in the
presence of Pt(PPh3)4catalyst was heated to reux in benzene for
20 h, four products were obtained. We separated these products
Scheme 2. Reactions of 1 with Diphenylacetylene
Scheme 3. Proposed Mechanism for the Production of
Compound 2
Scheme 4. Reactions of 1 with Phenylacetylene and Trimethylsilylacetylene
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8518
from the reaction mixture by column chromatography and
identied them as (E)- and (Z)-1,1,3,3-tetraisopropyl-2-
((trimethylsilyl)methylene)-2,3-dihydro-1H-[1,3]disilolo[4,5-
b]thiophene (6a,b), (2-((diisopropyl)silyl)thiophen-3-yl)-
diisopropyl((trimethylsilyl)ethynyl)silane (7), and (3-
((diisopropyl)silyl)thiophen-2-yl)diisopropyl((trimethylsilyl)-
ethynyl)silane (8) by mass spectrometry (MS) and 1H, 13C, and
29Si NMR spectroscopy (Scheme 4). Unfortunately, all attempts
Scheme 5. Reaction Mechanism for the Production of 7 and 8
Scheme 6. Production of 6a and 6b from 7 and 8
Scheme 7. Reaction Mechanism of 6a and 6b from 7 and 8
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8519
to separate 6a and 6b were unsuccessful. The ratio of 6a and 6b
was determined to be 1.6:1 by 1H NMR spectrometric analysis.
Compounds 7and 8were isolated by column chromatography.
The 1H NMR spectrum of 7shows a singlet signal at 0.22 ppm
attributed to trimethylsilyl protons and a triplet signal at 4.28
ppm attributed to a proton on the silicon atom along with
isopropyl and thienyl protons. The 13C NMR spectrum of 7
reveals a resonance at 0.16 ppm, attributed to trimethylsilyl
carbons, and two resonances at 108.70 and 118.17 ppm,
attributed to sp-hybridized carbons as well as isopropyl and
thienyl carbons. The 1H NMR spectrum of 8shows a singlet
signal at 0.21 ppm, attributed to trimethylsilyl protons, and a
double of triplet signal at 4.46 ppm (4J= 3.6 Hz, 5J= 0.8 Hz),
attributed to an SiH proton along with isopropyl and thienyl
protons. The 13C NMR spectrum of 8reveals a resonance at
0.10 ppm, attributed to trimethylsilyl carbons, and two
resonances at 109.90 and 119.38 ppm, attributed to the sp-
hybridized carbons as well as isopropyl and thienyl carbons.
Scheme 5 illustrates a possible mechanistic interpretation of
the observed reaction course for products 7and 8. The formation
of 7and 8can be understood in terms of insertion of the Pt atom
into the sp-hybridized CH bond of the acetylene leading to 9,
followed by the reductive elimination of the CSi bond to give
complexes 10a and 10b.
To conrm whether or not compounds 6a and 6b were
produced from compounds 7and 8by intramolecular hydro-
silation, we carried out reactions of 7and 8with a catalytic
amount of the platinum complex. Thus, the reaction of 7with 6
mol % of tetrakis(triphenylphosphine)platinum(0) at 80 °C for
130 h produced 6a and 6b in the ratio of 1.2:1 in 47% combined
yield, in addition to the starting compound 7and its isomer 8in
the ratio of 3:1 (Scheme 6). The presence of 8in this reaction
obviously indicates the production of intermediates 11a and 11b
(Scheme 7). A similar reaction of 8in the presence of the
platinum catalyst produced 6a and 6b in the ratio of 1.3:1 in 49%
combined yield, along with the starting compound 8and its
isomer 7in the ratio of 5:1. These reactions indicate that
compounds 6a and 6b must come from the intramolecular
hydrosilation of 7and 8.
Next, we investigated the reaction of 1with mesitylacetylene.
Thus, when a mixture of 1and mesitylacetylene in the presence
of the same catalyst was heated to reux in benzene for 4 h, a
mixture of the regioisomers was obtained in an almost
quantitative yield (Scheme 8). The regioisomers were separated
using column choromatography and assigned as [2-
(diisopropylsilyl)thiophen-3-yl]diisopropyl(mesitylethynyl)-
silane (12) and [3-(diisopropylsilyl)thiophen-2-yl]diisopropyl-
(mesitylethynyl)silane (13) by MS and 1H, 13C, and 29Si NMR
spectroscopy.
2.1. Theoretical Calculations. We carried out density
functional theory (DFT) calculations to investigate the energy
and structural changes from 7and 8to 6a and 6b. Gaussian09
program package
18
was employed together with the Becke-three-
parameter-LeeYangParr hybrid functional.
19
The Los Alamos
eective core potentials were used for Si, S, and Pt atoms along
with the corresponding valence basis sets.
20
For the H and C
atoms, the DunningHuzinaga full double-basis set was
employed.
21
Scheme 8. Reaction of 1 with Mesitylacetylene
Scheme 9. Species Evaluated by DFT Calculation
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8520
The calculated species are Pt complexes, that is, 10a,10b,11a,
and 11b, shown in Scheme 7, and their transition states (TSs)
are labeled as TS(1011a) and TS(1011b). Furthermore, 10a
and 10b mutually convert through TS(1014a), 14a, TS(14a
14b), 14b, and TS(1410b), as shown in Scheme 9, and these
structures were also investigated. In the calculations, TS is
searched rst. Then, the intrinsic reaction coordinate (IRC)
analysis is carried out at each TS for both directions, that is, the
reactant and product sides. Because the IRC analysis for the
whole reaction path is very time consuming, it is limited to the
Figure 1. Optimized structures for 10a, TS(1011a), and 11a (left) and 10b, TS(1011b), and 11b (right). Total energy of 10a is taken as the zero of
relative energy.
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8521
neighborhood of TSs, and normal optimization runs are
followed at the end points of IRC analysis. Finally, we conrmed
that the reaction path is seamlessly connected from the reactant,
via TS, to the product.
Figure 2. Optimized structures for TS(1014a) and 14a (left), TS(1014b) and 14b (right), and TS(14a14b) (bottom center). Total energy of 10a
is taken as the zero of relative energy.
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8522
The optimized structures for 10a, TS(1011a), and 11a are
shown in the left column and 10b, TS(1011b), and 11b are
shown in the right column of Figure 1.Figure 2 shows the
optimized structures for TS(1014a) and 14a in the left column
and TS(1014b) and 14b in the right column. TS(14a14b)is
a junction between the a- and b-series and shown at the center.
For all of the intermediate species, the optimized structures are
well-described by the line drawing structures shown in Schemes 7
and 9. A noticeable point is that a weak interaction between the
CC group and the Pt atom is suggested in 10a and 10b.
The energy changes for all reaction paths are shown in Figure
3. The total energy of 10a is taken as zero. Among the TSs,
TS(1011a)andTS(1011b) are the highest, and the
conversions from 10a to 11a and 10b to 11b are the rate-
determining step. This means that compounds in the a- and b-
series can mix with each other because the energies for TS(10a
14a), TS(14a14b), and TS(1014b) are lower than those for
TS(1011a) and TS(1011b). The conversion from 14a to
14b proceeds by an internal rotation of the SiCPt group. The
energy dierence between the corresponding compounds in the
a- and b-series is relatively small (at maximum, 16.3 kJ mol1
between 10a and 10b). Thus, the energetics for the a- and b-
series reactions resemble each other, and the two sets of reactions
occur in parallel.
3. CONCLUSIONS
We describe here the preparation and platinum-catalyzed
reactions of 2,3-bis(diisopropylsilyl)thiophene. The reactions
of 2,3-bis(diisopropylsilyl)thiophene 1with diphenylacetylene in
the presence of a catalytic amount of Pt(PPh3)4proceeded to
give cyclic products such as 2. A similar treatment of 1with
monosubstituted alkynes bearing less bulky substituents such as
phenylacetylene aorded regioisomers of the cyclic products,
whereas with alkynes bearing bulky substituents, 1gives products
arising from sp-hybridized CH bond activation of the alkynes.
In the intramolecular reactions of 10a and 10b, the theoretical
calculations indicate that the conversion from 14a to 14b or 14b
to 14a proceeds with the internal rotation of the SiCPt group.
4. EXPERIMENTAL SECTION
4.1. General Procedure. All reactions of 2,3-bis-
(isopropylsilyl)thiophene 1were carried out under an
atmosphere of dry nitrogen. Yields of the products were
determined by analytical gasliquid chromatography (GLC)
with the use of tridecane or pentadecane as an internal standard
on the basis of the starting compound used. NMR spectra were
recorded on a JMN-ECS400 spectrometer. Low- and high-
resolution mass spectra were recorded on a JEOL model JMS-
700 instrument. Column chromatography was performed using
Wakogel C-300 (WAKO).
4.2. Preparation of 2,3-Bis(diisopropylsilyl)thiophene
1. In a 200 mL three-necked ask tted with a stirrer, a reux
condenser, and a dropping funnel were placed 1.75 g (72.0
mmol) of magnesium, 11.2 g (74.3 mmol) of chlorodiisopro-
pylsilane, and 0.411 g (4.59 mol) of copper(I) cyanide in 10 mL
of dry THF. To this mixture was added dropwise a solution of
5.94 g (20.6 mmol) of 2-bromo-3-iodothiophene in 10 mL of dry
THF. The mixture was heated to reux for 4 h. The resulting
magnesium salts were removed by ltration and washed with
ether. The solvent was evaporated, and the residue was distilled
under reduced pressure to give 2.35 g (37% yield) of 2,3-
bis(diisopropylsilyl)thiophene 1:bp7780 °C/2 torr; anal.
calcd for C16H32Si2S: C, 61.46; H, 10.32. Found: C, 61.16; H,
10.01. MS m/z312 (M+); 1H NMR δ(CDCl3) 0.96 (d, 6H,
MeCH, J= 7.6 Hz), 1.01 (d, 6H, MeCH, J= 7.6 Hz), 1.06 (d, 6H,
MeCH, J= 7.6 Hz), 1.07 (d, 6H, MeCH, J= 7.6 Hz), 1.151.28
(m, 4H, CHMe), 4.24 (t, 1H, HSi, J= 3.6 Hz), 4.36 (t, 1H, HSi, J
= 3.6 Hz), 7.29 (d, 1H, thienyl ring proton, J= 4.4 Hz), 7.67 (d,
1H, thienyl ring proton, J= 4.4 Hz); 13C NMR δ(CDCl3) 12.25,
11.72 (CHMe); 18.79 (2C), 18.84, 18.93 (MeCH); 130.05,
134.29, 142.04, 143.82 (thienyl ring carbons); 29Si NMR
δ(CDCl3)3.52, 3.09.
4.2.1. Platinum-Catalyzed Reaction of 1with Diphenyla-
cetylene. In a 30 mL two-necked ask tted with a reux
condenser were placed 0.310 g (0.991 mmol) of 1, 0.355 g (1.99
mmol) of diphenylacetylene, and 0.0658 g (0.0529 mmol) of
Pt(PPh3)4in 15 mL of dry benzene. The mixture was heated to
reux for 96 h. The mixture was analyzed by GLC as being 2
(80% yield), along with starting compound 1(20% yield). The
solvent benzene was evaporated, and the residue was chromato-
graphed on a silica gel column using hexane as the eluent: HR-
MS: calcd for C30H40Si2S 488.2389, found: 488.2378. MS m/z
488 (M+); 1H NMR δ(CDCl3) 0.80 (d, 6H, MeCH, J= 7.6 Hz),
0.81 (d, 6H, MeCH, J= 7.6 Hz), 1.04 (d, 6H, MeCH, J= 7.6 Hz),
1.09 (d, 6H, MeCH, J= 7.6 Hz), 1.33 (sep, 2H, CHMe, J= 7.6
Hz), 1.34 (sep, 2H, CHMe, J= 7.6 Hz), 6.856.90 (m, 4H,
phenyl ring protons), 6.916.96 (m, 2H, phenyl ring protons),
7.04 (dt, 4H, phenyl ring protons, J= 7.8, 0.9 Hz), 7.39 (d, 1H,
thienyl ring proton, J= 4.4 Hz), 7.76 (d, 1H, thienyl ring proton, J
= 4.4 Hz); 13C NMR δ(CDCl3) 11.98, 12.44 (CHMe), 17.62,
17.85, 17.87, 18.04 (MeCH), 125.14, 125.22, 127.28, 127.31,
128.47, 128.56, 129.87, 132.75, 141.44, 143.99, 144.05, 144.16,
156.36, 157.62 (phenyl, thienyl ring and olenic carbons); 29Si
NMR δ(CDCl3)10.6, 9.2.
Figure 3. Energies for all optimized structures along the a- and b-reaction paths. Total energy of 10a is taken as the zero of relative energy.
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8523
4.2.2. Platinum-Catalyzed Reaction of 1with Phenyl-
acetylene. In a 30 mL two-necked ask tted with a reux
condenser were placed 0.312 g (0.998 mmol) of 1, 0.205 g (2.01
mmol) of phenylacetylene, and 0.0649 g (0.0522 mmol) of
Pt(PPh3)4in 15 mL of dry benzene. The mixture was heated to
reux for 73 h. The mixture was analyzed by GLC as being 5a and
5b (60% combined yield), together with starting compound 1
(40% yield). The solvent benzene was evaporated, and the
residue was chromatographed on a silica gel column using hexane
as the eluent. The ratio of 5a and 5b was determined to be 3:1 by
1H NMR spectroscopic analysis. For 5a and 5b: HR-MS: calcd
for C24H36Si2S 412.2076, found: 412.2055. MS m/z412 (M+);
1H NMR δ(CDCl3) 0.79 (d, 3H, MeCH, J= 7.2 Hz (5b)), 0.80
(d, 3H, MeCH, J= 7.2 Hz (5a)), 0.986 (d, 3H, MeCH, J= 7.2 Hz
(5b)), 0.992 (d, 3H, MeCH, J= 7.2 Hz (5a)), 1.02 (d, 6H,
MeCH, J= 7.2 Hz (5b)), 1.03 (d, 6H, MeCH, J= 7.2 Hz (5a)),
1.24 (sep, 1H, CHMe, J= 7.2 Hz (5a), overlapped with 5b), 1.36
(sep, 1H, CHMe, J= 7.2 Hz (5a)), 1.37 (sep, 1H, CHMe, J= 7.2
Hz (5b)), 6.75 (s, 1H, HCC(5b)), 6.77 (s, 1H, HCC
(5a)), 7.197.27 (m, 3H, phenyl ring protons (5a), overlapped
with 5b), 7.307.34 (m, 2H, phenyl ring protons (5a),
overlapped with 5b), 7.35 (d, 1H, thienyl ring proton, J= 4.4
Hz (5a)), 7.36 (d, 1H, thienyl ring proton, J= 4.4 Hz (5b)), 7.74
(d, 1H, thienyl ring proton, J= 4.4 Hz (5b)), 7.75 (d, H, thienyl
ring proton, J= 4.4 Hz (5a)); 13C NMR δ(CDCl3) 12.38 (5a),
12.53 (5b), 12.70 (5b), 12.96 (5a) (CHSi), 17.68 (5a), 17.90
(5a), 17.98 (5b), 18.09 (3C) (5b), 18.33 (5a), 18.40 (5a)
(MeCH), 126.42 (2C) (5a and 5b), 126.68 (5b), 126.72 (5a),
128.15 (5b), 128.18 (5a), 129.97 (5b), 130.18 (5a), 132.34 (5a),
133.17 (5b), 141.23 (5b), 141.85 (5a), 143.82 (5a), 144.00 (5b),
144.15 (5b), 145.30 (5a), 149.25 (5a), 160.74 (5a), 162.08 (5b)
(phenyl and thienyl ring and olenic carbons); 29Si NMR
δ(CDCl3)10.4 (5a), 8.84 (5b), 8.76 (5b), 7.51 (5a).
4.2.3. Platinum-Catalyzed Reaction of 1with Trimethylsi-
lylacetylene. In a 30 mL two-necked ask tted with a reux
condenser were placed 0.309 g (0.988 mmol) of 1, 0.215 g (2.19
mmol) of trimethylsilylacetylene, and 0.0640 g (0.0514 mmol) of
Pt(PPh3)4in 15 mL of dry benzene. The mixture was heated to
reux for 20 h. The mixture was analyzed by GLC as being 6a and
6b (22% combined yield) and 7and 8(71% combined yield).
The ratio of 6a and 6b in the reaction mixture was determined to
be 1.6:1 by 1H NMR spectroscopic analysis. The ratio of 7and 8
in the reaction mixture was determined to be 1:1 by 1H NMR
spectroscopic analysis. The solvent benzene was evaporated, and
the residue was chromatographed on a silica gel column using
hexane as the eluent. For 6a and 6b: HR-MS: calcd for
C21H40Si3S 408.2159, found: 408.2141. MS m/z408 (M+); 1H
NMR δ(CDCl3) 0.19 (s, 9H, Me3Si (6a and 6b)), 0.82 (d, 6H,
MeC, J= 7.6 Hz (6a)), 0.83 (d, 6H, MeC, J= 7.6 Hz (6b)), 0.98
(d, 6H, MeC, J= 7.6 Hz (6b)), 1.00 (d, 6H, MeC, J= 7.6 Hz
(6a)), 1.05 (d, 6H, MeC, J= 7.6 Hz (6b)), 1.07 (d, 6H, MeC, J=
7.6 Hz (6a)), 1.17 (d, 6H, MeC, J= 7.6 Hz (6b)), 1.19 (d, 6H,
MeC, J= 7.6 Hz (6a)), 0.951.28 (m, 4H, HC (6a), overlapped
with 6b), 7.34 (d, 1H, thienyl ring proton, J= 4.0 Hz (6b)), 7.35
(d, 1H, thienyl ring proton, J= 4.0 Hz (6a)), 7.64 (s, 1H, HCC
(6a), overlapped with 6b), 7.75 (d, 1H, thienyl ring proton, J=
4.0 Hz (6a)), 7.77 (d, 1H, thienyl ring proton, J= 4.0 Hz (6b));
13C NMR δ(CDCl3)0.49 (6a and 6b) (Me3Si), 13.12 (6a),
13.19 (6b), 14.34 (6b), 14.58 (6a) (HCSi), 18.41 (6a), 18.57
(6b), 18.66 (6a), 18.82 (6b), 19.38 (6b), 19.50 (6a), 19.53 (6b),
19.75 (6a) (MeC), 130.93 (6b), 132.24 (6a), 133.19 (6a),
133.69 (6b), 147.52 (6b), 148.17 (6a), 151.61 (6a), 152.09 (6b),
163.92 (6b), 164.07 (6a), 167.91 (6a), 167.97 (6b) (thienyl ring
and olenic carbons); 29Si NMR δ(CDCl3)9.70 (6b), 9.64
(6a), 2.36 (6b), 2.08 (6b), 0.62 (6a), 0.46 (6a). For 7:
HR-MS: calcd for C18H33Si3S 365.1611, found: 365.1614. MS m/
z365 (M+-(i-Pr)); 1H NMR δ(CDCl3) 0.22 (s, 9H, Me3Si), 0.96
(d, 6H, MeC, J= 7.6 Hz), 0.98 (d, 6H, MeC, J= 7.6 Hz), 1.06 (d,
6H, MeC, J= 7.6 Hz), 1.12 (d, 6H, MeC, J= 7.6 Hz), 1.20 (dsep,
2H, HC, J= 7.6 Hz, 3.6 Hz), 1.33 (sep, 2H, HC, J= 7.6 Hz), 4.28
(t, 1H, SiH, J= 3.6 Hz), 7.29 (d, 1H, thienyl ring proton, J= 4.4
Hz), 7.65 (d, 1H, thienyl ring proton, J= 4.4 Hz) 13C NMR
δ(CDCl3)0.16 (Me3Si), 11.95, 13.88 (HC), 18.23, 18.60,
19.00, 19.14 (MeC), 108.70, 118.17 (CC), 130.48, 134.75,
142.60 (2C), (thienyl ring carbons); 29Si NMR δ(CDCl3)
18.19, 11.07, 3.71. For 8: HR-MS: calcd for C18H33Si3S
365.1611, found: 365.1592. MS m/z365 (M+-(i-Pr)); 1H NMR
δ(CDCl3) 0.21 (s, 9H, Me3Si), 0.92 (d, 6H, MeC, J= 7.6 Hz),
1.02 (d, 6H, MeC, J= 7.6 Hz), 1.08 (d, 6H, MeC, J= 7.6 Hz),
1.09 (d, 6H, MeC, J= 7.6 Hz), 1.25 (dsep, 2H, HC, J= 7.6 Hz, 3.6
Hz), 1.26 (sep, 2H, HC, J= 7.6 Hz), 4.46 (dt, 1H, SiH, J= 3.6 Hz,
0.8 Hz), 7.51 (d, 1H, thienyl ring proton, J= 4.4 Hz), 7.65 (dd,
1H, thienyl ring proton, J= 4.4 Hz, 0.8 Hz); 13C NMR δ(CDCl3)
0.10 (Me3Si), 12.58, 13.56 (HC), 18.21, 18.49, 18.79, 18.96
(MeC), 109.90, 117.38 (CC), 129.56, 136.54, 141.43, 143.69
(thienyl ring carbons); 29Si NMR δ(CDCl3)18.52, 10.61,
3.23.
4.2.4. Platinum-Catalyzed Reaction of 7.In a 30 mL two-
necked ask tted with a reux condenser were placed 0.1110 g
(0.271 mmol) of 7and 0.0190 g (0.0152 mmol) of Pt(PPh3)4in
9 mL of dry benzene. The mixture was heated to reux for 130 h.
The mixture was analyzed by GLC as being 6a and 6b (47%
combined yield) and 7and 8(53% combined yield). The ratio of
6a and 6b in the reaction mixture was determined to be 1.2:1 by
1H NMR spectroscopic analysis. The ratio of 7and 8in the
reaction mixture was determined to be 3:1 by 1H NMR
spectroscopic analysis. All spectral data for 6a,6b,7, and 8were
identical to those of the authentic samples.
4.2.5. Platinum-Catalyzed Reaction of 8.In a 30 mL two-
necked ask tted with a reux condenser were placed 0.0979 g
(0.239 mmol) of 8and 0.0168 g (0.0135 mmol) of Pt(PPh3)4in
6 mL of dry benzene. The mixture was heated to reux for 130 h.
The mixture was analyzed by GLC as being 6a and 6b (49%
combined yield) and 7and 8(51% combined yield). The ratio of
6a and 6b in the reaction mixture was determined to be 1.3:1 by
1H NMR spectroscopic analysis. The ratio of 7and 8in the
reaction mixture was determined to be 1:5 by 1H NMR
spectroscopic analysis. All spectral data for 6a,6b,7, and 8were
identical to those of the authentic samples.
4.2.6. Platinum-Catalyzed Reaction of 1with Mesitylace-
tylene. In a 30 mL two-necked ask tted with a reux condenser
were placed 0.306 g (0.979 mmol) of 1, 0.312 g (2.16 mmol) of
mesitylacetylene, and 0.0653 g (0.0525 mmol) of Pt(PPh3)4in
15 mL of dry benzene. The mixture was heated to reux for 4 h.
The mixture was analyzed by GLC as being 12 and 13 (97%
combined yield). The ratio of 12 and 13 in the reaction mixture
was determined to be 1:1 by 1H NMR spectroscopic analysis.
The solvent benzene was evaporated, and the residue was
chromatographed on a silica gel column using hexane as the
eluent. For 12: HR-MS: calcd for C24H35Si2S 411.1998, found:
411.2005. MS m/z411 (M+-(i-Pr)); 1H NMR δ(CDCl3) 0.99 (d,
6H, MeCH, J= 7.2 Hz), 1.00 (d, 6H, MeCH, J= 7.2 Hz), 1.07 (d,
6H, MeCH, J= 7.2 Hz), 1.151.25 (m, 2H, HC), 1.20 (d, 6H,
MeCH, J= 7.2 Hz), 1.43 (sep, 2H, HC, J= 7.2 Hz), 2.29 (s, 3H,
Mes), 2.51 (s, 6H, Mes), 4.29 (t, 1H, SiH, J= 3.2 Hz), 6.89 (s,
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8524
2H, Mes-H), 7.32 (d, 1H, thienyl ring proton, J= 4.4 Hz), 7.66
(d, 1H, thienyl ring proton, J= 4.4 Hz); 13C NMR δ(CDCl3)
11.98, 14.26, (HCMe), 18.53, 18.90, 18.98, 19.11 (MeCH),
21.31, 21.36 (Mes-Me), 96.45, 106.24 (CC), 120.17, 127.53,
130.59, 134.78, 138.05, 141.07, 142.05, 143.63 (mesityl and
thienyl ring carbons); 29Si NMR δ(CDCl3)9.20, 3.57. For
13: HR-MS: calcd for C24H35Si2S 411.1998, found: 411.2000.
MS m/z411 (M+-(i-Pr)); 1H NMR δ(CDCl3) 0.97 (d, 6H, MeC,
J= 7.2 Hz), 1.02 (d, 6H, MeC, J= 7.2 Hz), 1.06 (d, 6H, MeC, J=
7.2 Hz), 1.18 (d, 6H, MeC, J= 7.2 Hz), 1.601.25 (m, 2H, HC),
1.38 (sep, 2H, HC, J= 7.2 Hz), 2.29 (s, 3H, p-Mes), 2.48 (s, 6H,
o-Mes), 4.50 (brt, 1H, SiH, J= 3.2 Hz), 6.89 (s, 2H, Mes-H), 7.65
(s, 2H, thienyl ring protons); 13C NMR δ(CDCl3) 12.59, 13.97
(CHSi), 18.51, 18.72, 18.76, 18.90 (CH3), 21.30, 31.35 (Mes-
Me), 97.68, 105.65 (CC), 120.32, 127.55, 129.54, 137.01,
137.93, 140.82, 140.97, 144.60 (thienyl and mesityl ring
carbons); 29Si NMR δ(CDCl3)8.86, 3.32.
ASSOCIATED CONTENT
*
SSupporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acsomega.7b01628.
1H, 13C, and 29Si NMR spectra for all products (PDF)
AUTHOR INFORMATION
Corresponding Authors
*E-mail: anaka@chem.kusa.ac.jp (A.N.).
*E-mail: hisabbit@yahoo.co.jp (H.K.).
ORCID
Akinobu Naka: 0000-0003-0019-9104
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
We thank grants-in-aid or Scientic Research program (No.
26410061) from the Ministry of Education, Science, Sports, and
Culture of Japan, and Wesco Science Promotion Foundation.
REFERENCES
(1) Corey, J. Y. Reactions of hydrosilanes with transition metal
complexes. Chem. Rev. 2016,116, 1129111435.
(2) Tanaka, M.; Uchimaru, Y.; Lautenschlager, H.-J. Platinum-
complex-catalyzed dehydrogenative double silylation of acetylenes,
dienes, and olefins with bis(hydrosilanes). Organometallics 1991,10,
1618.
(3) Uchimaru, Y.; Lautenschlager, H.-J.; Wynd, A. J.; Tanaka, M.;
Goto, M. Platinum-catalyzed dehydrogenative double silylation of
carbonyl compounds with o-bis(dimethylsilyl)benzene. Organometallics
1992,11, 26392643.
(4) Tanaka, M.; Uchimaru, Y.; Lautenschlager, H.-J. Platinum
complex-catalyzed dehydrogenative double silylation of olefins and
dienes with o-bis(dimethylsilyl)benzene. J. Organomet. Chem. 1992,428,
112.
(5) Tanaka, M.; Uchimaru, Y. Dehydrogenative double silylation of
acetylenes with bis(hydrosilane) compounds. Bull. Soc. Chim. Fr. 1992,
129, 667675.
(6) Uchimaru, Y.; El Sayed, A. M. M.; Tanaka, M. Selective arylation of
a silicon-hydrogen bond in o-bis(dimethylsilyl)benzene via carbon-
hydrogen bond activation of arenes. Organometallics 1993,12, 2065
2069.
(7) Tsutsumi, H.; Sunada, Y.; Nagashima, H. Novel disilaplatinacy-
clopentenes bearing dialkylsulfide ligands: Preparation, characterization,
and mechanistic consideration of hydrosilane reduction of carboxamides
by bifunctional organohydrosilanes. Organometallics 2011,30,6876.
(8) Eaborn, C.; Metham, T. N.; Pidcock, A. Some platinum complexes
containing chelating bis(silyl) ligands. J. Organomet. Chem. 1973,63,
107117.
(9) Ishikawa, M.; Naka, A.; Ohshita, J. Platinum-catalyzed reactions of
3,4-benzo-1,1,2,2-tetraethyl-1,2-disilacyclobut-3-ene. Organometallics
1993,12, 49874992.
(10) Naka, A.; Yoshizawa, K.; Kang, S.-Y.; Yamabe, T.; Ishikawa, M.
Synthesis and platinum- and palladium-Catalyzed reactions of benzo-
[1,2:4,5]bis(1,1,2,2-tetraethyl-1,2-disilacyclobut-3-ene). Organometal-
lics 1998,17, 58305835.
(11) Naka, A.; Ishikawa, M. Thermolysis, photolysis, and transition-
metal-catalyzed reactions of 3,4-benzo-1,1,2,2-tetraethyl-1,2-disilacyclo-
but-3-ene. Synlett 1995,1995, 794802.
(12) Naka, A.; Okada, T.; Kunai, A.; Ishikawa, M. Platinum-catalyzed
reactions of 3,4-benzo-1,1,2,2-tetra(isopropyl)-1,2-disilacyclobut-3-ene
with mono- and di-substituted alkynes. J. Organomet. Chem. 1997,547,
149156.
(13) Ishikawa, M.; Naka, A.; Yoshizawa, K. The chemistry of
benzodisilacyclobutenes and benzobis(disilacyclobutene)s: new devel-
opment of transition-metal-catalyzed reactions, stereochemistry and
theoretical studies. Dalton Trans. 2016,45, 32103235.
(14) Naka, A.; Mihara, T.; Kobayashi, H.; Ishikawa, M. Platinum-
catalyzed reactions of 2,3-bis(diethylsilyl)thiophene with alkynes. J.
Organomet. Chem. 2016,822, 221227.
(15) Kyushin, S.; Matsuura, T.; Matsumoto, H. 2,3,4,5-Tetrakis-
(dimethylsilyl)thiophene: The First 2,3,4,5-tetrasilylthiophene. Organo-
metallics 2006,25, 27612765.
(16) Ozawa, F.; Kamite, J. Mechanistic study on the insertion of
phenylacetylene into cis-bis(silyl)platinum(II) complexes. Organo-
metallics 1998,17, 56305639.
(17) Sakaki, S.; Mizoe, N.; Sugimoto, M. Theoretical study of
platinum(0)-catalyzed hydrosilylation of ethylene. Chalk-Harrod
mechanism or modified Chalk-Harrod mechanism. Organometallics
1998,17, 25102523.
(18) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb,
M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.;
Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.;
Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima,
T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.;
Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin,
K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.;
Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.;
Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.;
Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.;
Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.;
Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador,
P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman,
J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision C.01;
Gaussian, Inc.: Wallingford, CT, 2010.
(19) Becke, A. D. Density-functional thermochemistry. III. The role of
exact exchange. J. Chem. Phys. 1993,98, 56485652.
(20) Hay, P. J.; Wadt, W. R. Ab initio effective core potentials for
molecular calculations. Potentials for the transition metal atoms Sc to
Hg. J. Chem. Phys. 1985,82, 270283.
(21) Dunning, T. H., Jr.; Hay, P. J. Methods of Electronic Structure
Theory. In Modern Theoretical Chemistry; Schaefer, H. F., III, Ed.;
Plenum Press: New York, 1976; pp 128.
ACS Omega Article
DOI: 10.1021/acsomega.7b01628
ACS Omega 2017, 2, 85178525
8525
... [29][30][31][32] A key step in the mechanism of transition metal-catalysed hydrosilylation is the oxidative addition of the metal to the Si-H bond. [21][22][23]25,26,[33][34][35] While this fundamental step of organometallic chemistry, to form (silyl)M(H) complexes, has been well-studied for Pd's neighbouring elements Rh, [36][37][38][39][40][41][42][43][44][45][46][47][48] Pt, [49][50][51][52][53][54][55][56][57][58][59][60][61][62][63][64] and most other transition metals, [33][34][35] it is much less understood for Pd. The oxidative addition of Rh and Pt to Si-H is facile, and complexes have been reported with a variety of ligand scaffolds and silanes. ...
Article
Full-text available
The oxidative addition of Pd to Si-H bonds is a crucial step in a variety of catalytic applications, and many aspects of this reaction are poorly understood. One important yet underexplored aspect is the electronic effect of silane substituents on reactivity. Herein we describe a systematic investigation of the formation of silyl palladium hydride complexes as a function of silane identity, focusing on electronic influence of the silanes. Using [(μ-dcpe)Pd]2 (dcpe = dicyclohexyl(phosphino)ethane) and tertiary silanes, data show that equilibrium strongly favours products formed from electron-deficient silanes, and is fully dynamic with respect to both temperature and product distribution. A notable kinetic isotope effect (KIE) of 1.21 is observed with H/DSiPhMe2 at 233 K, and the reaction is shown to be 0.5th order in [(μ-dcpe)Pd]2 and 1st order in silane. Formed complexes exhibit temperature-dependent intramolecular H/Si ligand exchange on the NMR timescale, allowing determination of the energetic barrier to reversible oxidative addition. Taken together, these results give unique insight into the individual steps of oxidative addition and suggest the initial formation of a σ-complex intermediate to be rate-limiting. The insight gained from these mechanistic studies was applied to hydrosilylation of alkynes, which shows parallel trends in the effect of the silanes' substituents. Importantly, this work highlights the relevance of in-depth mechanistic studies of fundamental steps to catalysis.
Article
Organosilane compounds are widely used in both organic synthesis and materials science. Particularly, 1,2-disilylated and gem-disilylated alkenes, characterized by a carbon-carbon double bond and multiple silyl groups, exhibit significant potential for subsequently diverse transformations. The versatility of these compounds renders them highly promising for applications in materials, enabling them to be valuable and versatile building blocks in organic synthesis. This review provides a comprehensive summary of methods for the preparation of cis/trans-1,2-disilylated and gem-disilylated alkenes. Despite notable advancements in this field, certain limitations persist, including challenges related to regioselectivity in the incorporation and chemoselectivity in the transformation of two nearly identical silyl groups. The primary objective of this review is to outline synthetic methodologies for the generation of these alkenes through disilylation reactions, employing silicon reagents, specifically disilanes, hydrosilanes, and silylborane reagents. The review places particular emphasis on investigating the practical applications of the C-Si bond of disilylalkenes and delves into an in-depth discussion of reaction mechanisms, particularly those reactions involving the activation of Si-Si, Si-H, and Si-B bonds, as well as the C-Si bond formation.
Article
The synthesis and platinum-catalyzed reactions of 2,3-bis(dimethylsilyl)pyridine (1) are described. The reaction of 1 with diphenylacetylene in the presence of a platinum catalyst gave the corresponding six-membered ring compound arising from dehydrogenative double silylation of the alkyne. Similar reactions of 1 with mono-substituted acetylenes such as tert-butylacetylene, 1-hexyne, and phenylacetylene proceeded to give regioisomers of cyclic products.
Article
Full-text available
When mixed with two different Lewis acid catalysts of zinc and indium, terminal alkynes were found to react with bis(hydrosilane)s to selectively provide 1,1‐disilylalkenes from among several possible products, by way of a sequential dehydrogenative silylation/intramolecular hydrosilylation reaction. Adding a pyridine base is crucial in this reaction; a switch as a catalyst of the zinc Lewis acid is turned on by forming a zinc−pyridine‐base complex. A range of the 1,1‐disilylalkenes can be obtained by a combination of aryl and aliphatic terminal alkynes plus aryl‐, heteroaryl‐, and naphthyl‐tethered bis(hydrosilane)s. The 1,1‐disilylalkene prepared here is available as a reagent for further transformations by utilizing its C−Si or C=C bond. The former includes Hiyama cross‐coupling, bismuth‐catalyzed ether formation, and iododesilylation; the latter includes double alkylation and epoxidation. Mechanistic studies clarified the role of the two Lewis acids: the zinc–pyridine‐base complex catalyzes the dehydrogenative silylation as a first stage, and, following on this, the indium Lewis acid catalyzes the ring‐closing hydrosilylation as a second stage, thus leading to the 1,1‐disilylalkene. image
Article
New types of polymers composed of bithiophene condensed with disilacyclohexadiene rings and benzothiadiazole were prepared by the platinum-catalyzed dehydrogenative bissilylation. UV–Vis absorption and fluorescence properties of these polymers have been investigated. The fluorescence spectra of these polymers showed only the bands ascribed to the benzothiadiazole units even when the bithiophene units were irradiated. For these polymers, photoenergy transfer occurs from donor to acceptor. DFT calculations were performed to understand the photophysical properties of the polymers.
Article
The reactions of 3,4-bis(dimethylsilyl)thiophene (1) and 2,3,4,5-tetrakis(dimethylsilyl)thiophene (17) with alkynes and alkenes have been reported. The reactions of 1 with alkynes such as diphenylacetylene, 3-hexyne, phenyltrimethylsilylacetylene, and trimethylsilylacetylene, in the presence of a catalytic amount of tetrakis(triphenylphosphine)platinum(0) at 80 °C gave the [1,4]disilino[2,3-c]thiophene derivatives. With alkenes under the same conditions, 1 afforded the five-membered ring products, in addition to hydrosilylation products. Treatment of 2,3,4,5-tetrakis(dimethylsilyl)thiophene (17) with diphenylacetylene and styrene gave the respective dehydrogenative double silylation products, arising from 2 equivalents of the alkyne and alkene, and compound 17.
Article
The reactions of 2,3-bis(diethylsilyl)thiophene 1 with alkynes such as diphenylacetylene, 3-hexyne and phenylacetylene in the presence of a catalytic amount of Pt(PPh3)4 in benzene at 80 °C proceeded to give products 2, 3, and 6a,b, respectively. Treatment of a mixture consisting of 1 and 1,4-diethynylbenzene with a Pt(PPh3)4 catalyst in refluxing benzene produced three types of isomers consisting of two di(silyl)thiophene units and a diethynylbenzene unit. Computational analyses for the reaction mechanism of the platinum-catalyzed reaction have also been reported.
Article
This third review in a series involving reactions of hydrosilanes and transition metal complexes and characterization of the products covers the period 2009 thru 2013. After a brief discussion of other synthetic methods used for the formation of Si-TM complexes, Section 3 provides an extended discussion of the types of ligands and metal complexes used as reactants with hydrosilanes. The increase in use of pincer ligands in forming stable, isolable complexes is featured. Three tables list the complexes reported for primary, secondary, and tertiary silanes. Many reactions leading to SiTM complexes are initiated by loss of a ligand prior to oxidative addition of the hydrosilane or by a metathesis reaction. Structural data are tabulated for the isolated complexes and provide the Si-TM bond ranges for the elements in the "d" block. A major section on nonclassical (σ or agostic) complexes includes two general groupings with Si-H···TM and M-H···Si interactions. A section on solution processes identified by NMR spectroscopy is dominated by hydride exchange examples. A section on bonding outlines a unifying bonding description that has been proposed as well as calculations reported for Si-TM complexes, both real and hypothetical examples.
Article
The synthesis and reactions of 3,4-benzo-1,2-disilacyclobut-3-enes and benzo[1,2:4,5]bis(1,2-disilacyclobut-3-ene)s developed in our group are reported in this review. The palladium-, platinum- and nickel-catalyzed reactions of benzodisilacyclobutenes and benzobis(disilacyclobutene)s with unsaturated compounds afford various types of products. The structures of the products depend highly on the nature of transition metal used as a catalyst. The reactions of cis- and trans-benzodisilacyclobutene with alkenes and alkynes in the presence of a transition-metal catalyst proceed with high stereospecificity to give the respective adducts. The thermal reactions of cis- and trans-benzodisilacyclobutene with various substrates also proceed stereospecifically to give adducts. Results of theoretical calculations for the platinum-catalyzed reaction of disilacyclobutene with acetylene, the nickel-catalyzed reaction of benzodisilacyclobutene and thermal reaction of benzobis(disilacyclobutene) are discussed in this review.
Article
Three new disilaplatinacyclopentane complexes, (Me2S)Pt[(SiMe2)2C6H4]2 (1), (MeSCH2CH2CH2SMe)Pt[(SiMe2)2C6H4]2 (2), and (MeSCH2CH2CH2SMe)Pt[(SiMe2)2C6H4] (3), were prepared and characterized by the reaction of (PtMe2)2(μ-SMe2)2 with 1,2-bis(dimethylsilyl)benzene in the presence of dialkylsulfide ligands. Catalytic hydrosilane reduction of carboxamides with 1,2-bis(dimethylsilyl)benzene was investigated in the presence of these new complexes, and the possible involvement of disilaplatinacycle dihydride species, “Pt(Si)2(H)2(SR2)”, stabilized by dialkylsulfides in catalysis was discussed.
Article
Olefins underwent dehydrogenative 1,2- and/or 1,1-double silylation with o-bis(dimethylsilyl)benzene (1) in the presence of a catalytic amount of Pt(CH2CH2)PPh3)2 to afford benzo-1,4-disilacyclohexene (3) and/or benzo-1,3-disilacyclopentene (4) derivatives, respectively. In reactions of aliphatic olefins, single hydrosilylation with one of the SiH bonds took place as a side reaction to give HMe2Si(o-C6H4)SiMe2R compounds. Depending on the conditions of the latter reactions, HMeRSi(o-C6H4)SiMe3 compounds arising from 1,4-rearrangement of a methyl group were also formed. The cyclic bis(silyl)platinum complex Me2t(PPh3)2 (11) formed upon treatment of 1 with the platinum-ethylene complex readily reacted with olefins in the presence of 1 to afford 3 and/or 4. Compound 4 coming from the catalytic reaction of styrene with o-bis(deuteriodimethylsilyl)benzene did not have the deuterium label incorporated. Based on these results, the dehydrogenative double silylation is proposed to take place via 11 as a key intermediate.
Article
Acetylenes, dienes, and olefins underwent dehydrogenative double silylation with bis(hydrosilane) species in the presence of platinum complex catalysts to give disilacyclic compounds in good yields.
Article
Treatment of 3,4-benzo-1,1,2,2-tetraethyl-1,2-disilacyclobutene (1) with a catalytic amount of (eta2-ethylene)bis(triphenylphosphine)platinum(0) in refluxing benzene gave 1-(diethylphe-nylsilyl)-2-(diethylsilyl)benzene (2) and cis-4,5-benzo-1,1,3-triethyl-2-methyl-1,3-disilacyclopent-4-ene (3) in 10 % and 67 % yields. The platinum-catalyzed reaction of 1 with ethylene produced 5,6-benzo-1,1,4,4-tetraethyl-1,4-disilacyclohex-5-ene, while with styrene and 1-hexene, compound 1 afforded 4,5-benzo-1,3-disilacyclopent-4-ene derivatives. A similar reaction of 1 with a-methylstyrene produced an adduct arising from hydrosilylation of 3 to alpha-methylstyrene. The reaction of 1 with phenylacetylene, diphenylacetylene, phenyl(trimethylsilyl)acetylene, and 3-hexyne in the presence of a platinum catalyst yielded the respective 5,6-benzo-1,4-. disilacyclohexa-2,5-dienes. With benzaldehyde, 1 gave a 5,6-benzo-2-oxa-1,4-disilacyclohex-5-ene derivative.
Article
The reactions of 3,4-benzo-1,1,2,2-tetra(isopropyl)-1,2-disilacyclobut-3-ene (1) with phenylacetylene and 1-hexyne in the presence of a catalytic amount of (η2-ethylene)bis(triphenylphosphine)platinum(0) at 200°C for 24 h gave two types of 1:1 adducts, 2-phenyl- and 2-butyl-substituted 5,6-benzo-1,1,4,4-tetra(isopropyl)-1,4-disilacyclohexa-2,5-diene and 2-benzylidene- and 2-pentylidene-substituted 4,5-benzo-1,1,3,3-tetra(isopropyl)-1,3-disilacyclopent-4-ene, respectively. With mesityl- and dimethylphenylsilylacetylene. 1 afforded 2-(mesityl)methylene- and 2-(dimethylphenylsilyl)methylene-4,5-benzo-1,1,3,3-tetra(isopropyl)-1,3-disilacylclopent-4-ene. A similar reaction of 1 with trimethylsilylacetylene produced an adduct arising from sp-hybridized CH bond activation of the acetylene, together with a benzodisilacyclopentene derivative. The reaction of 1 with diphenylacetylene yielded a 2,3-diphenyl-5,6-benzodisilacyclohexa-2,5-diene derivative, while with methylphenylacetylene and 2-hexyne, 1 gave 2-benzylidene- and 2-butylidene-5,6-benzo-1,4-disilacyclohex-5-ene, along with the benzodisilacyclohexa-2,5-diene derivatives. Similar treatment of 1 with methyl(trimethylsilyl)acetylene produced a 5,6-benzo-2-(trimethylsilyl)methylenedisilacyclohexene derivative.
Article
The dihydrides o-(HMe2Si)2C6H4 and o-(HMe2Si)C6H4CH2SiMe2H react with [PtL2(C2H4)] (L=PPh3) at room temperature to give the 5- and 6-membered cyclic bis(silyl) complexes [Pt(SiMe2C6H4SiMe2-o)L2] (1), and [Pt{SiMe2C6H4-(CH2SiMe2)-o} L2] (II), respectively. The disiloxane (HPh2Si)2O reacts with [Pt2(C2H4)] to give a 4-membered cyclic species [Pt(SiPh2OSiPh2)L2] at 45° ; (HMe2-Si)2O gives the hydrido(silyl) complex cis-[PtH(SiMe2OSiMe2H)L2] at 45°, but the latter cyclises at 75° to give [Pt(SiMe2OSiMe2)L2]. The dihydrides o-(HMe2SiCH2)2C6H4 and HMe2Si(CH2)4SiMe2H, which could, in principle, give complexes with 7-membered rings, in fact give only cis-[PtH {SiMe2CH2C6H4(CH2-SiMe2H)-o} L2] and cis-[PtH{SiMe2(CH2)4SiMe2H}L2] on reaction with [PtL2-(C2H4)]. Complexes (I) and (11) react with 1,2-bis(diphenylphosphino) ethane (Diphos) to give the doubly-chelated species [Pt(SiMe2C6H4SiMe2-o) (Diphos)] and [Pt-{SiMe2C6H4(CH2SiMe2)-o} (Diphos)] respectively. Both silicon atoms in complexes (I) and (II) are displaced from platinum by bromine, methyl iodide, or phenylacetylene. The complexes cis-[PtH(SiPh2CH2CH=CH2)L2] and cis-[PtH(SiMe2C≡CSiMe2H)L2] react with methyldiphenylsilane to give cis-[PtH(SiPh2Me)L2].