ArticlePDF Available

Synthesis of Pyridine-Fused Siloles by Palladium-Catalyzed Intramolecular Bis-Silylation

Authors:

Abstract

Silole derivatives are attracting significant attention as new functional materials with excellent electronic and photophysical properties. Thus, the development of synthesis methods to afford such derivatives is highly desirable. Herein, the synthesis of pyridine-fused siloles under the conditions of the Sonogashira coupling reaction is described. The reactions of 2-bromo-3-(pentamethyldisilanyl)pyridine (1) with ethynylbenzene derivatives in the presence of PdCl2(PPh3)2-CuI as a catalyst afforded the corresponding pyridine-fused siloles (2a-2c) through intramolecular trans-bis-silylation. DFT calculations were also performed to understand the reaction mechanism. This paper is the first to report on the successful use of palladium catalysts in the trans-bis-silylation of alkynes with disilanes.
Synthesis of Pyridine-Fused Siloles by Palladium-Catalyzed
Intramolecular Bis-Silylation
Akinobu Naka,*Natsumi Shimomura, and Hisayoshi Kobayashi*
Cite This: ACS Omega 2022, 7, 30369−30375
Read Online
ACCESS Metrics & More Article Recommendations *
Supporting Information
ABSTRACT: Silole derivatives are attracting significant attention as
new functional materials with excellent electronic and photophysical
properties. Thus, the development of synthesis methods to aord such
derivatives is highly desirable. Herein, the synthesis of pyridine-fused
siloles under the conditions of the Sonogashira coupling reaction is
described. The reactions of 2-bromo-3-(pentamethyldisilanyl)pyridine
(1) with ethynylbenzene derivatives in the presence of PdCl2(PPh3)2-CuI as a catalyst aorded the corresponding pyridine-fused
siloles (2a2c) through intramolecular trans-bis-silylation. DFT calculations were also performed to understand the reaction
mechanism. This paper is the first to report on the successful use of palladium catalysts in the trans-bis-silylation of alkynes with
disilanes.
INTRODUCTION
Silicon-containing compounds are used in various fields
ranging from synthetic organic chemistry to functional
materials and pharmaceutical synthesis. Thus, new synthetic
methods for such compounds have been developed.
15
Many
articles have reported on the transition metal-catalyzed bis-
silylation of alkynes with disilane.
613
Some examples of cis-
bis-silylation of internal alkynes with acyclic disilanes have also
been reported (Scheme 1). For example, Ozawa et al. reported
the bis-silylation of Me3SiSiF2Ph with internal alkynes using a
Pd catalyst.
14
Spencer et al. investigated the intermolecular bis-
silylation of internal alkynes with hexamethyldisilane using the
[(NHC)2Pd(SiMe3)2] (NHC: N-heterocyclic carbene) com-
plex as a precatalyst.
15
Zhao et al. reported nickel-catalyzed bis-
silylation of internal alkynes with an unsymmetrical coordinat-
ing disilane, 8-(2-substituted-1,1,2,2-tetramethyldisilanyl)-
quinoline.
6
Some examples of intramolecular versions have also been
described, which allow access to heterocycles with dierent
ring sizes. In 2012, Matsuda and Ichioka developed a selective
synthesis of four or five silylated rings.
16
They investigated
whether regioselectivity depends on the catalyst used and
found that when (2-alkynylphenyl)disilane was catalyzed by a
Pd/isocyanide system, cis-bis-silylation occurred to form four-
membered ring compounds. In contrast, rhodium-catalyzed
reactions of (2-alkynylphenyl)disilane aorded five-membered
ring compounds, that is, 3-silyl-1-benzosiloles, via trans-bis-
silylation (Scheme 2).
Recently, we reported the synthesis of pyridine-fused siloles
using ruthenium-catalyzed hydrosilylation and their optical
properties.
17
Silole derivatives are attracting attention as new
functional materials with excellent electronic and photo-
physical properties for various applications, such as organic
light-emitting diodes (OLEDs), photovoltaic devices, and
semiconductors.
1823
Herein, we report the synthesis of
pyridine-fused siloles under the conditions of the Sonogashira
coupling reaction.
24
These reactions are expected to proceed
through the trans-bis-silylation of 3-(1,1,2,2,2-pentamethyldi-
silanyl)-2-(arylethynyl)pyridine. Notably, palladium catalysts
have never been utilized in the trans-bis-silylation of alkynes
with disilanes before.
Received: June 10, 2022
Accepted: August 8, 2022
Published: August 16, 2022
Scheme 1. Bis-Silylation of Internal Alkynes
Article
http://pubs.acs.org/journal/acsodf
© 2022 The Authors. Published by
American Chemical Society 30369
https://doi.org/10.1021/acsomega.2c03637
ACS Omega 2022, 7, 3036930375
RESULTS AND DISCUSSION
The reaction of 2,3-dibromopyridine with i-PrMgCl at room
temperature followed by quenching with chloropentamethyl-
disilane aorded 2-bromo-3-(pentamethyldisilanyl)pyridine
(1) in 70% yield.
25
To obtain 2-phenylethynyl-3-
(pentamethyldisilanyl)pyridine, 1was reacted with phenyl-
acetylene under the Sonogashira coupling reaction conditions.
When a mixture of 1and ethynylbenzene in the presence of a
PdCl2(PPh3)2-CuI catalyst was heated to reflux in triethyl-
amine, an unexpected product, 1,1-dimethyl-2-phenyl-3-
(trimethylsilyl)-1H-silolo(3,2-b)pyridine (2a), was obtained
in 15% isolated yield, along with intermolecular bis-silylation
product 3a (8% yield) from the reaction of 2-phenylethynyl-3-
(pentamethyldisilanyl)pyridine with phenylacetylene (Scheme
3). Compound 2a was produced via intramolecular trans-bis-
silylation. Many unidentified products were also detected in
the reaction mixture by gas liquid chromatography (GLC) and
gel permeation chromatography (GPC). Song et al. reported
an unsymmetrical disilane by attaching a coordinating group to
one of the two silicon atoms, which was successfully applied to
the Pd-catalyzed bis-silylation of terminal alkynes.
26
Compound 2a was isolated by column chromatography, and
its structure was verified by spectrometric analyses. The mass
spectrum for 2a shows a parent ion at m/z 309 corresponding
to the calculated molecular weight of C18H23Si2N. The 1H
NMR spectrum for 2a shows singlets at 0.04 and 0.29 ppm,
representing methyl protons on the silicon atoms, three
doublets of doublet signals at 7.00, 7.76, and 8.49 ppm,
ascribed to the pyridyl protons, and signals corresponding to
the phenyl protons. The 29Si NMR spectrum for 2a shows
signals at 6.2 and 3.7 ppm.
We carried out the Pd(PPh3)2Cl2-catalyzed reaction of 1
with ethynylbenzene in the absence of CuI. Although many
unidentified products were detected in the reaction mixture by
GLC and GPC, compound 2a was not detected in the reaction
mixture. We believe that a copper catalyst is necessary for
coupling reactions.
A similar reaction of 1with 4-ethynyltoluene under the same
conditions aorded 1,1-dimethyl-2-(p-tolyl)-3-(trimethylsilyl)-
1H-silolo(3,2-b)pyridine (2b) in 21% yield, in addition to
compound 3b (7% yield), which is analogous to 3a. Many
unidentified products were also detected in the reaction
mixture by GLC and GPC. The structures of 2b and 3b were
verified by NMR spectroscopy and mass spectrometry.
3-Ethynyltoluene was also reacted with 1in the presence of
the PdCl2(PPh3)2-CuI catalyst to obtain 1,1-dimethyl-2-(m-
tolyl)-3-(trimethylsilyl)-1H-silolo(3,2-b)pyridine (2c) in 16%
isolated yield. NMR and MS analyses showed the existence of
bis-silylation product 3c (3% yield) obtained from the reaction
of 3-pentamethyldisilanyl-2-(m-tolylethynyl)pyridine with 3-
ethynyltoluene.
Scheme 2. Regioselective Intramolecular Bis-Silylation of
Aryldisilane and Alkynes
Scheme 3. Palladium-Catalyzed Reaction of 1 with Ethynylbenzene Derivatives
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.2c03637
ACS Omega 2022, 7, 3036930375
30370
When 1was heated with 1-chloro-4-ethynylbenzene and 2-
ethynyl-1,3,5-trimethylbenzene under the same conditions,
neither 1H-silolo(3,2-b)pyridine derivatives such as 2a2c nor
the products analogous to 3a3c were detected by NMR
spectroscopy. However, many unidentified products were
detected in the reaction mixture by GLC and GPC.
In the proposed reaction mechanism to obtain 2, cyclic
products were formed with the catalyst, as shown in Scheme 4.
First, compound Ais produced by the Sonogashira coupling
reaction of 1with ethynylbenzene derivatives, followed by
oxidative addition of SiSi bonds to produce Pd complex B.
Complex Bfurther undergoes SiC bond formation to deliver
intermediate C. Complex Cthen undergoes ring expansion to
the five-membered cyclic complex D, followed by reductive
elimination to aord bis-silylated product 2and regenerate the
Pd catalyst.
Next, we carried out a similar reaction of 1with
ethynyltrimethylsilane in the presence of a catalytic amount
of PdCl2(PPh3)2-CuI. When compound 1was stirred under
reflux for 12 h in triethylamine in the presence of a catalytic
amount of the Pd complex, 3-(1,1,2,2,2-pentamethyldisilanyl)-
2-(trimethylsilylethynyl)pyridine (4) was obtained in 17%
isolated yield (Scheme 5). No 1,1-dimethyl-2,3-bis-
(trimethylsilyl)-1H-silolo(3,2-b)pyridine (2d) analogous to
2a2c was detected in the reaction mixture. A similar reaction
of 1with 3,3-dimethyl-1-butyne aorded 2-(3,3-dimethylbut-1-
yn-1-yl)-3-(1,1,2,2,2-pentamethyldisilanyl)pyridine (5) in 5%
yield. Many unidentified products were detected in the
reaction mixture by GLC and GPC. Bulky substituents,
which prevented the formation of 4and 5, explain their low
yields.
27
THEORETICAL STUDY
DFT calculations were performed to investigate the energy and
structural changes in the synthesis route from 3-(1,1,2,2,2,-
pentamethyldisilanyl)-2-(phenylethynyl)pyridine (6) to trans-
adduct 2a through cis-adduct (Z)-7,7-dimethyl-8-(phenyl-
(trimethylsilyl)methylene)-2-aza-7-silabicyclo(4.2.0)octa-1,3,5-
triene (7,Scheme 6). The Gaussian 09 program package
28
was
employed along with Becke’s three-parameter LeeYangParr
hybrid functional.
29
The Los Alamos eective core potentials
30
and the Dunning/Huzinaga full double zeta basis sets
31
were
used for the Pd atom, while the 6-311G(d) basis sets were
used for the H, C, N, Si, and P atoms.
The Pd catalyst was modeled using the Pd(PPh3) complex,
which formed when three PPh3ligands detach from Pd(PPh3)4
owing to steric hindrance. Two reaction mechanisms were
expected: (1) a parallel mechanism from 6to cis-adduct 7and
trans-adduct 2a and (2) a sequential mechanism in which cis-
adduct 7is first formed from 6, and then 7is converted to 2a.
However, only the latter reaction route could be determined
by the DFT calculations. Figure 1 shows the energy change
along the reaction coordinate which consists of seven
transition states (TSs) and eight local minima (LMs). LM0
is 6, LM4 is cis-adduct 7, and LM7 is trans-adduct 2a.
Individual structures are shown in the Supporting Information.
The formation of TS6 is the rate-determining step with an
activation energy of 161 kJ mol1. Although the activation
energy was higher for cis-to-trans conversion than that for the
formation of cis-adduct 7,trans-adduct 2a was more stable by
91.4 kJ mol1. If the activation energy of 161 kJ mol1was not
a significant barrier under the present reaction conditions, the
dierence between the stability would be the driving force for
the formation of 2a.
We also carried out DFT calculations for the production of
trans-adducts 2a in the absence of a ligand on the palladium
atom. The results revealed that the TS was the rate-
determining step with an activation energy of 215.3 kJ mol1.
Scheme 4. Proposed Reaction Mechanism for the Production of 2
Scheme 5. Palladium-Catalyzed Reactions of 1 with
Ethynyltrimethylsilane and 3,3-Dimethyl-1-butyne
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.2c03637
ACS Omega 2022, 7, 3036930375
30371
CONCLUSIONS
We investigated the reactions of 2-bromo-3-
(pentamethyldisilanyl)pyridine (1) with terminal alkynes in
the presence of PdCl2(PPh3)2-CuI as a catalyst. The reactions
of 1with ethynylbenzene derivatives provided the correspond-
ing ring compounds (2a2c) via intramolecular bis-silylation.
Similar reactions of 1with ethynyltrimethylsilane and 3,3-
dimethyl-1-butyne aorded 2-ethynyl-3-
(pentamethyldisilanyl)pyridine derivatives 4and 5, respec-
tively. DFT calculations were performed to rationalize the
formation of 1,1-dimethyl-2-phenyl-3-(trimethylsilyl)-1H-
silolo(3,2-b)pyridine 2a via cis-bis-silylation adduct 7. These
reactions comprise the palladium-catalyzed trans-bis-silylation
of internal alkynes. To the best of our knowledge, this is the
first report on the use of palladium-containing catalysts in such
reactions.
METHODS
General Procedure. All reactions of 1were carried out
under an inert atmosphere using dry nitrogen. NMR spectra
were recorded on a JMN-ECS400 spectrometer. Low-
resolution mass spectrometry was performed on a JEOL
JMS-700 mass spectrometer. High-resolution mass spectrom-
etry (HR-MS) was performed on a JEOL JMS-700 mass
spectrometer and a Thermo Scientific LTQ Orbitrap XL
hybrid Fourier-transform mass spectrometer using electrospray
ionization. Column chromatography was performed using a
silica gel column (Wakogel C-300; Wako Pure Chemical
Industries).
Preparation of 2-Bromo-3-(1,1,2,2,2-
Pentamethyldisilanyl)Pyridine (1). In a 300 mL three-
necked flask fitted with a stirrer, reflux condenser, and
dropping funnel, 10.008 g (42.2 mmol) of 2,3-dibromopyr-
idine was added to 100 mL of dry tetrahydrofuran (THF).
Next, a THF solution comprising 21.1 mL (42.2 mmol) of 2.0
M isopropyl magnesium chloride was added dropwise at room
temperature. The mixture was then stirred for 1 h at room
temperature, and 7.047 g (42.2 mmol) of chloropentamethyl-
disilane was added. The resulting mixture was stirred for 6 h. It
was then hydrolyzed, after which the organic layer was
separated, washed with water, and dried over anhydrous
magnesium sulfate. The solvent was evaporated, and the
residue was chromatographed on a silica gel column eluting
with hexane to obtain 8.474 g (70% yield) of 2-bromo-3-
(1,1,2,2,2-pentamethyldisilanyl)pyridine 1: HR-MS: calcd for
C10H19NSi2Br: (M + H+): 288.02339, found: 288.02411. MS
m/z 287 (M+); 1H NMR δ(CDCl3) 0.13 (s, 9H, Me3Si), 0.43
(s, 6H, Me2Si), 7.22 (dd, 1H, pyridyl-ring proton, J= 7.2, 5.2
Hz), 7.65 (dd, 1H, pyridyl-ring proton, J= 7.2, 2.4 Hz), 8.29
(dd, 1H, pyridyl-ring proton, J= 5.2, 2.4 Hz); 13C NMR
δ(CDCl3)3.0 (Me2Si), 1.4 (Me3Si), 122.3, 139.4, 144.9,
149.6, 149.9 (pyridyl-ring carbons); 29Si NMR δ(CDCl3)
18.1, 17.8.
Palladium-Catalyzed Reaction of 1 with Ethynylben-
zene. In a 100 mL two-necked flask fitted with a reflux
condenser, 1(2.014 g, 6.99 mmol), bis(triphenylphosphine)-
dichloropalladium (0.244 g, 0.348 mmol), and copper(I)
iodide (0.067 g, 0.352 mmol) were added to 25 mL of dry
triethylamine. To this mixture, ethynylbenzene (1.491 g, 13.9
Scheme 6. Synthesis Route of Compound 6 via 7 to Form Compound 2a
Figure 1. Energy diagrams for the transition of compound 6via 7to form compound 2a at the B3LYP/6-311G(d) level of theory.
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.2c03637
ACS Omega 2022, 7, 3036930375
30372
mmol) was added dropwise at room temperature, after which
the mixture was heated to reflux for 12 h. The solvent was then
evaporated, and the residue was chromatographed on a silica
gel column eluting with hexane-ethyl acetate (5:1) to obtain
0.303 g (15% yield) of 2a and 0.241 g (8% yield) of 3a. For 2a:
HR-MS: calcd for C18H24NSi2(M + H+): 310.14418, found:
310.14465. MS m/z 309 (M+); 1H NMR δ(CDCl3) 0.04 (s,
9H, Me3Si), 0.29 (s, 6H, Me2Si), 7.00 (dd, 1H, pyridyl-ring
proton, J= 6.8 Hz, 5.2 Hz), 7.05 (dd, 2H, phenyl-ring protons,
J= 7.2 Hz, 1.6 Hz), 7.23 (tt, 1H, phenyl-ring proton, J= 7.2
Hz, 1.6 Hz), 7.32 (t, 2H, phenyl-ring protons, J= 7.2 Hz), 7.76
(dd, 1H, pyridyl-ring proton, J= 6.8 Hz, 2.0 Hz), 8.49 (dd, 1H,
pyridyl-ring proton, J= 5.2 Hz, 2.0 Hz); 13C NMR δ(CDCl3)
4.9 (Me2Si), 0.9 (Me3Si), 120.2, 125.9, 126.6, 127.9, 131.3,
138.7, 142.8, 149.5, 158.9, 165.9, 171.7 (phenyl, pyridyl-ring
and olefinic carbons); 29Si NMR δ(CDCl3)6.2, 3.7. For 3a:
HR-MS: calcd for C26H30NSi2(M + H+), 412.19113; found,
412.19119. MS m/z 411 (M+); 1H NMR δ(CDCl3)0.11 (s,
3H, MeSi), 0.19 (s, 9H, Me3Si), 0.28 (s, 3H, Me2Si), 6.79
6.83 (m, 2H, phenyl-ring protons), 6.89 (s, 1H, HCC),
7.067.08 (m, 3H, phenyl-ring protons), 7.17 (dd, 1H,
pyridyl-ring proton, J= 7.4, 4.8 Hz), 7.297.31 (m, 3H,
phenyl-ring protons), 7.417.44 (m, 2H, phenyl-ring protons),
8.12 (dd, 1H, pyridyl-ring protons, J= 7.4 Hz, 2.0 Hz), 8.70
(dd, 1H, pyridyl-ring proton, J= 4.8 Hz, 2.0 Hz); 13C NMR
δ(CDCl3)1.1 (Me3Si), 0.1, 0.5 (MeSi), 93.7, 106.7 (sp
carbons), 120.2, 123.0, 127.5, 128.0, 128.2, 128.5, 128.6, 129.5,
131.9, 137.2, 138.7, 144.8, 148.2, 150.2, 166.5 (phenyl, pyridyl-
ring, and olefinic carbons); 29Si NMR δ(CDCl3)20.6, 1.2.
Palladium-Catalyzed Reaction of 1 with 4-Ethynylto-
luene. In a 100 mL two-necked flask fitted with a reflex
condenser, 1(2.145 g, 7.44 mmol), bis(triphenylphosphine)-
dichloropalladium (0.263 g, 0.375 mmol), and copper(I)
iodide (0.066 g, 0.347 mmol) were added to 25 mL of dry
triethylamine. To this mixture, 4-ethynyltoluene (1.728 g, 14.9
mmol) was added dropwise at room temperature. The mixture
was heated to reflux for 12 h. The solvent was then evaporated,
and the residue was chromatographed on a silica gel column
eluting with hexane-ethyl acetate (10:1) to obtain 0.585 g
(21% yield) of 2b and 0.244 g (7% yield) of 3b. For 2b: HR-
MS: calcd for C19H26NSi2(M + H+), 324.15983; found,
324.16043. MS m/z 323 (M+); 1H NMR δ(CDCl3) 0.05 (s,
9H, Me3Si), 0.28 (s, 6H, Me2Si), 2.32 (s, 3H, CH3), 6.94 (d,
2H, phenylene-ring protons, J= 8.0 Hz), 6.99 (dd, 1H, pyridyl-
ring proton, J= 7.4 Hz, 4.8 Hz), 7.12 (d, 2H, phenylene-ring
protons, J= 8.0 Hz), 7.75 (dd, 1H, pyridyl-ring proton, J= 7.4
Hz, 2.0 Hz), 8.48 (dd, 1H, pyridyl-ring proton, J= 4.8 Hz, 2.0
Hz); 13C NMR δ(CDCl3)4.9 (Me2Si), 1.0 (Me3Si), 21.2
(CH3), 120.1, 126.5, 128.6, 131.3, 135.5, 138.7, 139.7, 149.5,
158.6, 166.1, 171.7 (phenylene, pyridyl-ring, and olefinic
carbons); 29Si NMR δ(CDCl3)6.4, 3.3. For 3b: HR-MS:
calcd for C28H33NSi2(M+), 439.2152; found, 439.2156. MS
m/z 439 (M+); 1H NMR δ(CDCl3)0.10 (s, 3H, MeSi), 0.17
(s, 9H, Me3Si), 0.28 (s, 3H, Me2Si), 2.20 (s, 3H, Me), 2.34 (s,
3H, Me), 6.68 (d, 2H, phenylene-ring protons, J= 8.0 Hz),
6.84 (s, 1H, HCC), 6.87 (d, 2H, phenylene-ring protons, J=
8.0 Hz), 7.10 (d, 2H, phenylene-ring protons, J= 8.0 Hz), 7.16
(dd, 1H, pyridyl-ring proton, J= 7.6, 4.8 Hz), 7.31 (d, 2H,
phenylene-ring protons, J= 8.0 Hz), 8.12 (dd, 1H, pyridyl-ring
protons, J= 7.6 Hz, 2.0 Hz), 8.69 (dd, 1H, pyridyl-ring proton,
J= 4.8 Hz, 2.0 Hz); 13C NMR δ(CDCl3)1.1 (Me3Si), 0.2,
0.5 (MeSi), 21.0, 21.4 (Me), 92.2, 106.9 (sp carbons), 119.8,
120.1, 128.5, 128.6 (2C), 128.9, 129.3, 131.7, 134.4, 137.2,
138.7, 144.8, 146.7, 150.0, 166.6 (phenyl, pyridyl-ring, and
olefinic carbons); 29Si NMR δ(CDCl3)20.8, 1.4.
Palladium-Catalyzed Reaction of 1 with 3-Ethynylto-
luene. In a 100 mL two-necked flask fitted with a reflex
condenser, 1(2.020 g, 7.01 mmol), bis(triphenylphosphine)-
dichloropalladium (0.248 g, 0.353 mmol), and copper(I)
iodide (0.069 g, 0.362 mmol) were added to 25 mL of dry
triethylamine. To this mixture, 3-ethynyltoluene (1.677 g, 14.4
mmol) was added dropwise at room temperature. The mixture
was heated to reflux for 12 h. The solvent was then evaporated,
and the residue was chromatographed on a silica gel column
eluting with hexane-ethyl acetate (20:1) to obtain 0.420 g
(16% yield) of 2c and 0.107 g (3% yield) of 3c. For 2c: HR-
MS: calcd for C19H26NSi2(M + H+), 324.15983; found,
324.16028. MS m/z 323 (M+); 1H NMR δ(CDCl3) 0.04 (s,
9H, Me3Si), 0.29 (s, 6H, Me2Si), 2.36 (s, 3H, CH3), 6.84 (d,
1H, phenylene-ring proton, J= 7.2 Hz), 6.86 (s, 1H,
phenylene-ring proton), 7.00 (dd, 1H, pyridyl-ring proton, J
= 7.2, 4.8 Hz), 7.04 (d, 1H, phenylene-ring proton, J= 7.2
Hz), 7.20 (t, 1H, phenylene-ring proton, J= 7.2 Hz), 7.75 (dd,
1H, pyridyl-ring proton, J= 7.2, 1.6 Hz), 8.49 (dd, 1H, pyridyl-
ring proton, J= 4.8, 1.6 Hz); 13C NMR δ(CDCl3)4.8
(Me2Si), 1.0 (Me3Si), 21.5 (CH3), 120.2, 123.7, 126.7, 127.3,
127.8, 131.3, 137.3, 138.7, 142.7, 149.5, 158.6, 166.1, 171.7
(phenylene, pyridyl-ring, and olefinic carbons); 29Si NMR
δ(CDCl3)6.4, 3.5. For 3c: HR-MS: calcd for C28H33NSi2
(M+), 439.2152; found, 439.2155. MS m/z 439 (M+); 1H
NMR δ(CDCl3)0.14 (s, 3H, MeSi), 0.19 (s, 9H, Me3Si),
0.29 (s, 3H, Me2Si), 2.10 (s, 3H, Me), 2.32 (s, 3H, Me), 6.59
(s, 1H, phenylene-ring proton), 6.60 (d, 1H, phenylene-ring
proton, J= 8.0 Hz), 6.85 (s, 1H, HC = C), 6.89 (d, 1H,
phenylene-ring proton, J= 8.0 Hz), 6.96 (t, 1H, phenylene-
ring proton, J= 8.0 Hz), 7.13 (d, 1H, phenylene-ring proton, J
= 8.0 Hz), 7.17 (dd, 1H, pyridyl-ring proton, J= 7.4, 5.2 Hz),
7.18 (t, 1H, phenylene-ring proton), 7.24 (d, 1H, phenylene-
ring proton, J= 8.0 Hz), 7.25 (s, 1H, phenylene-ring proton),
8.13 (dd, 1H, pyridyl-ring protons, J= 7.4 Hz, 2.0 Hz), 8.70
(dd, 1H, pyridyl-ring proton, J= 5.2 Hz, 2.0 Hz); 13C NMR
δ(CDCl3)1.1 (Me3Si), 0.1, 0.5 (MeSi), 21.16, 21.23 (Me),
92.7, 106.8 (sp carbons), 120.1, 122.7, 126.5, 127.8, 128.1,
128.2, 128.6, 128.9, 129.5, 130.3, 132.4, 137.0, 137.3, 137.8,
138.9, 144.8, 147.8, 150.1, 166.6 (phenyl, pyridyl-ring, and
olefinic carbons); 29Si NMR δ(CDCl3)20.6, 1.3.
Preparation of 3-(1,1,2,2,2-Pentamethyldisilanyl)-2-
(Trimethylsilylethynyl)pyridine (4). In a 300 mL three-
necked flask fitted with a stirrer, reflux condenser, and
dropping funnel, 1(2.007 g, 6.96 mmol), bis-
(triphenylphosphine)dichloropalladium (0.246 g, 0.351
mmol), and copper(I) iodide (0.067 g, 0.352 mmol) were
added to 25 mL of dry triethylamine. To this mixture,
ethynyltrimethylsilane (0.784 g, 7.98 mmol) was added
dropwise at room temperature. The mixture was heated to
reflux for 12 h. The solution was then hydrolyzed, and the
organic layer was separated, washed with water, and dried over
anhydrous magnesium sulfate. The solvent was then
evaporated, and the residue was chromatographed on a silica
gel column eluting with hexane-ethyl acetate (10:1) to obtain
0.326 g (17% yield) of 4: HR-MS: calcd for C15H28NSi3: (M +
H+)m 306.15241; found, 306.15237. MS m/z 305 (M+); 1H
NMR δ(CDCl3) 0.09 (s, 9H, Me3Si), 0.27 (s, 9H, Me3Si), 0.45
(s, 6H, Me2Si), 7.17 (dd, 1H, pyridyl-ring proton, J= 7.6, 5.2
Hz), 7.70 (dd, 1H, pyridyl-ring proton, J= 7.6, 2.0 Hz), 8.51
(dd, 1H, pyridyl-ring proton, J= 5.2, 2.0 Hz); 13C NMR
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.2c03637
ACS Omega 2022, 7, 3036930375
30373
δ(CDCl3)3.7 (Me2Si), 1.5, 0.3 (Me3Si), 96.6, 106.1 (sp
carbons), 122.4, 137.3, 142.0, 147.4, 149.4 (pyridyl-ring
carbons); 29Si NMR δ(CDCl3)20.7, 17.2, 16.6.
Preparation of 2-(3,3-Dimethylbut-1-yn-1-yl)-3-
(1,1,2,2,2-Pentamethyldisilanyl)pyridine (5). In a 300
mL three-necked flask fitted with a stirrer, reflux condenser,
and dropping funnel, 1(2.061 g, 7.15 mmol), bis-
(triphenylphosphine)dichloropalladium (0.246 g, 0.351
mmol), and copper(I)iodide (0.068 g, 0.357 mmol) were
added to 50 mL of dry triethylamine. To this mixture, 3,3-
dimethyl-1-butyne (0.672 g, 8.18 mmol) was added dropwise
at room temperature. The mixture was heated to reflux for 12
h. The solution was then hydrolyzed, and the organic layer was
separated, washed with water, and dried over anhydrous
magnesium sulfate. The solvent was then evaporated, and the
residue was chromatographed on a silica gel column eluting
with hexane-ethyl acetate (10:1) to obtain 0.101 g (5% yield)
of 5: HR-MS: calcd for C16H28NSi2: (M + H+), 290.17548;
found, 290.17584. MS m/z 289 (M+); 1H NMR δ(CDCl3)
0.09 (s, 9H, Me3Si), 0.45 (s, 6H, Me2Si), 1.35 (s, 9H, t-Bu),
7.11 (dd, 1H, pyridyl-ring proton, J= 7.6, 5.2 Hz), 7.66 (dd,
1H, pyridyl-ring proton, J= 7.6, 2.0 Hz), 8.48 (dd, 1H, pyridyl-
ring proton, J= 5.2, 2.0 Hz); 13C NMR δ(CDCl3)3.6
(Me2Si), 1.6 (Me3Si), 28.0 (CMe3), 30.6 (Me3C), 81.5, 99.9
(sp carbons), 121.6, 136.1, 142.1, 148.2, 149.4 (pyridyl-ring
carbons); 29Si NMR δ(CDCl3)20.5, 17.5.
ASSOCIATED CONTENT
*
Supporting Information
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acsomega.2c03637.
NMR spectra, optimized structures for all LMs and TSs,
and high-resolution mass spectra (PDF)
AUTHOR INFORMATION
Corresponding Authors
Akinobu Naka Department of Life Science, Kurashiki
University of Science and the Arts, Kurashiki, Okayama 712-
8505, Japan; orcid.org/0000-0003-0019-9104;
Email: anaka@chem.kusa.ac.jp
Hisayoshi Kobayashi Professor Emeritus, Kyoto Institute of
Technology, Kyoto 606-8585, Japan; Email: hisabbit@
yahoo.co.jp
Author
Natsumi Shimomura Department of Life Science, Kurashiki
University of Science and the Arts, Kurashiki, Okayama 712-
8505, Japan
Complete contact information is available at:
https://pubs.acs.org/10.1021/acsomega.2c03637
Notes
The authors declare no competing financial interest.
ACKNOWLEDGMENTS
We are grateful for grants to A.N. from the Yakumo
Foundation for Environmental Science and WESCO Science
Promotion Foundation. We thank Amimoto, the Natural
Science Center for Basic Research and Development (N-
BARD), and Hiroshima University for the HR-MS measure-
ment.
REFERENCES
(1) Hiyama, T.; Oestreich, M. Organosilicon Chemistry: Novel
Approaches and Reactions; Wiley-VCH Press: Weinheim, 2019.
(2) Min, G. K.; Hernández, D.; Skrydstrup, T. Efficient Routes to
CarbonSilicon Bond Formation for the Synthesis of Silicon-
Containing Peptides and Azasilaheterocycles. Acc. Chem. Res. 2013,
46, 457470.
(3) Cheng, C.; Hartwig, J. F. Catalytic Silylation of Unactivated C
H Bonds. Chem. Rev. 2015,115, 89468975.
(4) Mu, Q.-C.; Chen, J.; Xia, C.-G.; Xu, L.-W. Synthesis of
Silacyclobutanes and Their Catalytic Transformations Enabled by
Transition-Metal Complexes. Coord. Chem. Rev. 2018,374, 93113.
(5) Santra, S. Synthesis and Application of Siloles: From the Past to
Present. ChemistrySelect 2020,5, 90349058.
(6) Zhang, Y.; Wang, X. C.; Ju, C. W.; Zhao, D. Bis-Silylation of
Internal Alkynes Enabled by Ni(0) Catalysis. Nat. Commun. 2021,12,
68.
(7) Suginome, M.; Ito, Y. Palladium-Catalyzed Reactions of Organic
Halides with Olefins. Activation of SiliconSilicon σBonds by
Transition-Metal Complexes: Synthesis and Catalysis of New
Organosilyl Transition-Metal Complexes. J. Chem. Soc., Dalton
Trans. 1998, 19251934.
(8) Suginome, M.; Ito, Y. Activation of SiSi Bonds by Transition-
Metal Complexes. Organomet. Chem. 1999,3, 131159.
(9) Beletskaya, I.; Moberg, C. Element-Element Addition to Alkynes
Catalyzed by the Group 10 Metals. Chem. Rev. 1999,99, 34353462.
(10) Suginome, M.; Ito, Y. Transition-Metal-Catalyzed Additions of
SiliconSilicon and Silicon-Heteroatom Bonds to Unsaturated
Organic Molecules. Chem. Rev. 2000,100, 32213256.
(11) Beletskaya, I.; Moberg, C. Element-Element Additions to
Unsaturated CarbonCarbon Bonds Catalyzed by Transition Metal
Complexes. Chem. Rev. 2006,106, 23202354.
(12) Suginome, M.; Matsuda, T.; Ohmura, T.; Seki, A.; Murakami,
M. Comprehensive Organometallic Chemistry, 10th ed.; Crabtree, R. H.,
Mingos, D. M. P., Eds.; Elsevier: Oxford, 2007; Vol. III, pp 725787.
(13) Ansell, M. B.; Navarro, O.; Spencer, J. Transition Metal
Catalyzed ElementElement’ Additions to Alkynes. Coord. Chem. Rev.
2017,336, 5477.
(14) Ozawa, F.; Sugawara, M.; Hayashi, T. A New Reactive System
for Catalytic Bis-Silylation of Acetylenes and Olefins. Organometallics
1994,13, 32373243.
(15) Ansell, M. B.; Roberts, D. E.; Cloke, F. G. N.; Navarro, O.;
Spencer, J. Synthesis of an [(NHC)2Pd(SiMe3)2] Complex and
Catalytic cis-Bis(Silyl)ations of Alkynes with Unactivated Disilanes.
Angew. Chem., Int. Ed. 2015,54, 55785582.
(16) Matsuda, T.; Ichioka, Y. Rhodium-Catalysed Intramolecular
Trans-Bis-Silylation of Alkynes to Synthesise 3-Silyl-1-Benzosiloles.
Org. Biomol. Chem. 2012,10, 31753177.
(17) Naka, A.; Shimomura, N.; Kobayashi, H. Synthesis of Pyridine-
Fused Siloles by Ruthenium-Catalyzed Intramolecular Hydrosilyla-
tion. J. Organomet. Chem. 2022,959, 122201.
(18) Yamaguchi, S.; Tamao, K. The Chemistry of Organic Silicon
Compounds; Rappoport, Z., Apeloig, Y., Eds.; Wiley: Chichester,
2001; Chapter 11, Vol. 3.
(19) Liu, J.; Lam, J. W. Y.; Tang, B. Z. Aggregation-Induced
Emission of Silole Molecules and Polymers: Fundamental and
Applications. J. Inorg. Organomet. Polym. Mater. 2009,19, 249285.
(20) Sołoducho, J.; Zając, D.; Spychalska, K.; Baluta, S.; Cabaj, J.
Conducting Silicone-Based Polymers and Their Application. Molecules
2021,26, 2012.
(21) Dang, T. T.; Nguyen, H. M. T.; Nguyen, H.; Dung, T. N.;
Nguyen, M. T.; Dehaen, W. Advances in Synthesis of π-Extended
Benzosilole Derivatives and Their Analogs. Molecules 2020,25, 548.
(22) Chen, J. W.; Cao, Y. Silole-Containing Polymers: Chemistry
and Optoelectronic Properties. Macromol. Rapid Commun. 2007,28,
17141742.
(23) Lu, G.; Usta, H.; Risko, C.; Wang, L.; Facchetti, A.; Ratner, M.
A.; Marks, T. J. Synthesis, Characterization, and Transistor Response
of Semiconducting Silole Polymers with Substantial Hole Mobility
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.2c03637
ACS Omega 2022, 7, 3036930375
30374
and Air Stability. Experiment and Theory. J. Am. Chem. Soc. 2008,
130, 76707685.
(24) Sonogashira, K. Development of PdCu Catalyzed Cross-
Coupling of Terminal Acetylenes with sp2-Carbon Halides. J.
Organomet. Chem. 2002,653, 4649.
(25) Naka, A.; Yoshima, S.; Ishikawa, M. Platinum-Catalyzed
Reactions of 2,3-Bis(dimethylsilyl)pyridine with Mono- and Di-
Substituted Alkynes. Inorg. Chim. Acta 2021,517, 120210.
(26) Xiao, P.; Cao, Y.; Gui, Y.; Gao, L.; Song, Z. Me3Si
SiMe2[oCON(iPr)2C6H4]: An Unsymmetrical Disilane Reagent for
Regio- and Stereoselective Bis-Silylation of Alkynes. Angew. Chem., Int.
Ed. 2018,57, 47694773.
(27) We carried out the reaction of 1 with 1-hexyne in the presence
of a PdCl2(PPh3)2-CuI catalyst. Many products were detected in the
reaction mixture by GLC and GPC, and all attempts to isolate 3-
(dimethylsilyl)-2-(hex-1-yn-1-yl)pyridine (8), analogous to 4 and 5,
were unsuccessful. NMR and MS analyses showed the existence of
compound 8.
(28) 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.; 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.; 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.
(29) Becke, A. D. Density-Functional Thermochemistry. III. The
Role of Exact Exchange. J. Chem. Phys. 1993,98, 56485652.
(30) 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.
(31) Dunning, T. H., Jr.; Hay, P. J. Modern Theoretical Chemistry;
Schaefer, H. F., III, Ed.; Plenum Press: New York, 1976; pp 128.
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.2c03637
ACS Omega 2022, 7, 3036930375
30375
... Recently, we reported that the reactions of 2-bromo-3-(pentamethyldisilanyl)pyridine with ethynylbenzene derivatives in the presence of PdCl 2 (PPh 3 ) 2 -CuI as catalysts afforded the corresponding pyridine-fused siloles through intramolecular trans-bis-silylation [24]. DFT calculations for the above reaction were performed to rationalize the formation of trans-bis-silylation adducts via cis-bis-silylation adducts. ...
... DFT calculations for the above reaction were performed to rationalize the formation of trans-bis-silylation adducts via cis-bis-silylation adducts. We also demonstrated that the similar reactions of 2-bromo-3-(pentamethyldisilanyl)pyridine with alkynes having bulky substituents, such as ethynyltrimethylsilane, produced 2-ethynyl-3-pentamethyldisilanylpyridine derivatives arising from Sonogashira-coupling reactions [24]. Pyridine-containing materials have been examined for their optical and physical properties, as well as their medical potential [25]. ...
... The starting compound, 3-(1,1,2,2,2-pentamethyldisilanyl)-2-(trimethylsilylethynyl) pyridine (1), was prepared by the Sonogashira coupling reaction of 2-bromo-3-(1,1,2,2,2pentamethyldisilanyl)pyridine with ethynyltrimethylsilane in triethylamine [24]. It was shown that base desilylation can be accomplished in potassium carbonate/methanol to produce 2-ethynyl-3-(1,1,2,2,2-pentamethyldisilanyl)pyridine (2) (Scheme 1). ...
Article
Full-text available
Rhodium-catalyzed reactions of 2-ethynyl-3-pentamethyldisilanylpyridine derivatives (1 and 2) are reported. The reactions of compounds 1 and 2 in the presence of catalytic amounts of rhodium complexes at 110 °C gave the corresponding pyridine-fused siloles (3) and (4) through intramolecular trans-bis-silylation cyclization. The reaction of 2-bromo-3-(1,1,2,2,2-pentamethyldisilanyl)pyridine with 3-phenyl-1-propyne in the presence of PdCl2(PPh3)2-CuI catalysts afforded 1:2 bis-silylation adduct 6. DFT calculations were also performed to understand the reaction mechanism for the production of compound 3 from compound 1.
... In 2012, Matsuda et al. demonstrated the possibility of trans-selective bis-silylation of a C-C triple bond via intramolecular cyclization using a Rh(I) catalyst with (2-alkynylphenyl)disilanes [21]. In 2022, we reported the first intramolecular trans-bis-silylation reaction in the presence of PdCl 2 (PPh 3 ) 2 -CuI as the catalyst [22]. In 2023, we discovered the synthesis of siloles with condensed pyridine rings through intramolecular trans-bis-silylation reactions in the presence of a rhodium complex catalyst [23]. ...
... In most bis-silylation reactions, cis-adducts are stere 20]; however, trans-adduct formation reactions have been rece Matsuda et al. demonstrated the possibility of trans-selective bis-s bond via intramolecular cyclization using a Rh(I) catalyst with (2-a [21]. In 2022, we reported the first intramolecular trans-bis-silylat ence of PdCl2(PPh3)2-CuI as the catalyst [22]. In 2023, we discovere with condensed pyridine rings through intramolecular trans-bis-s presence of a rhodium complex catalyst [23]. ...
Article
Full-text available
Rhodium-catalyzed reactions of 3-ethynyl-2-pentamethyldisilanylthiophene derivatives (1a–1c) have been reported. At 110 °C, compounds 1a–1c reacted in the presence of a rhodium complex catalyst, yielding thiophene-fused siloles (2a–2c) through intramolecular trans-bis-silylation. To understand the production of 2a from 1a, the mechanism was investigated using density functional theory (DFT) calculations.
Article
Full-text available
The trans‐addition of intermetallic σ‐bonds (E−E′; E E’=Si, B, Se, P, Sn, etc.) across alkynes has emerged as an exceptionally atom‐economical means of constructing multi‐substituted alkenes, which readily transform into densely functionalized and biologically active compounds. Compared to the widely studied cis‐E−E’ additions, trans‐E−E’ additions remain a more challenging and underdeveloped area. In this article, recent advances in trans‐E−E’ additions are summarized, encompassing various interelement compounds such as Si−Si, B−B, Si−B, P−B, Se−B, and B−Sn. We aim to provide a deeper understanding of trans‐E−E’ additions and facilitate the further development of synthetic methodologies and applications of vicinally trans‐bismetallated alkenes.
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
Transition metal‐catalyzed bissilylation reactions of alkynes with disilane reagents have become one of the most straightforward and efficient protocols to rapidly produce structurally diverse alkenyl silicon derivatives. In these reactions, the utilization of unsymmetrical disilane reagents provided the possibilities for reactivity enhancement as well as the synthetic merits in contrast to symmetrical disilane reagents. However, a major yet challenging objective is achieving precise control over the selectivity including the regioselectivity and the cis / trans ‐selectivity. Herein we realized the first divergent bissilylation of alkynoates with our developed air‐stable disilane reagent 8‐(2‐substituted‐1,1,2,2‐tetramethyldisilanyl)quinoline (TMDQ) by means of synergistic Pd/Lewis acid catalytic system. The catalytic system precisely dictates the selectivity, resulting in the divergent synthesis of 1,2‐bissilyl alkenes. The power of these 1,2‐bissilyl alkenes serving as the key synthetic intermediates has been clearly demonstrated by rapid construction of diverse motifs and densely functionalized biologically active compounds. In addition, the origins of the switchable selectivities were well elucidated by experimental and computational studies on the reaction mechanism and were mainly attributed to different ligand steric effects, the use of the specific disilane reagent TMDQ and the different coordination modes of different Lewis acid with alkynoates.
Article
Full-text available
Transition metal‐catalyzed bissilylation reactions of alkynes with disilane reagents have become one of the most straightforward and efficient protocols to rapidly produce structurally diverse alkenyl silicon derivatives. In these reactions, the utilization of unsymmetrical disilane reagents provided the possibilities for reactivity enhancement as well as the synthetic merits in contrast to symmetrical disilane reagents. However, a major yet challenging objective is achieving precise control over the selectivity including the regioselectivity and the cis/trans‐selectivity. Herein we realized the first divergent bissilylation of alkynoates with our developed air‐stable disilane reagent 8‐(2‐substituted‐1,1,2,2‐tetramethyldisilanyl)quinoline (TMDQ) by means of synergistic Pd/Lewis acid catalytic system. The catalytic system precisely dictates the selectivity, resulting in the divergent synthesis of 1,2‐bissilyl alkenes. The power of these 1,2‐bissilyl alkenes serving as the key synthetic intermediates has been clearly demonstrated by rapid construction of diverse motifs and densely functionalized biologically active compounds. In addition, the origins of the switchable selectivities were well elucidated by experimental and computational studies on the reaction mechanism and were mainly attributed to different ligand steric effects, the use of the specific disilane reagent TMDQ and the different coordination modes of different Lewis acid with alkynoates.
Article
A copper-catalyzed disilylative cyclization of silicon-containing internal alkynes with silylboronates has been developed for the synthesis of 3-silyl-1-silacyclopent-2-enes. The reaction proceeded regio- and anti-selectively under simple and mild conditions by employing a combination of nucleophilic silicon donors and electrophilic silicon acceptors. The reaction could also be extended to the synthesis of a 1-germacyclopent-2-ene and a silicon-centered spirocyclic compound by using appropriate alkyne substrates.
Article
Full-text available
Over the past two decades, both fundamental and applied research in conducting polymers have grown rapidly. Conducting polymers (CPs) are unique due to their ease of synthesis, environmental stability, and simple doping/dedoping chemistry. Electrically conductive silicone polymers are the current state-of-the-art for, e.g., optoelectronic materials. The combination of inorganic elements and organic polymers leads to a highly electrically conductive composite with improved thermal stability. Silicone-based materials have a set of extremely interesting properties, i.e., very low surface energy, excellent gas and moisture permeability, good heat stability, low-temperature flexibility, and biocompatibility. The most effective parameters constructing the physical properties of CPs are conjugation length, degree of crystallinity, and intra- and inter-chain interactions. Conducting polymers, owing to their ease of synthesis, remarkable environmental stability, and high conductivity in the doped form, have remained thoroughly studied due to their varied applications in fields like biological activity, drug release systems, rechargeable batteries, and sensors. For this reason, this review provides an overview of organosilicon polymers that have been reported over the past two decades.
Article
Full-text available
1,2-Bis-silyl alkenes have exciting synthetic potential for programmable sequential synthesis via manipulation of the two vicinal silyl groups. Transition metal-catalyzed bis-silylation of alkynes with disilanes is the most straightforward strategy to access such useful building blocks. However, this process has some limitations: (1) symmetric disilanes are frequently employed in most of the reactions to assemble two identical silyl groups, which makes chemoselective differentiation for stepwise downstream transformations difficult; (2) the main catalysts are low-valent platinum group transition metal complexes, which are expensive; and (3) internal alkynes remain challenging substrates with low inherent reactivity. Thus, the development of abundant metal-catalyzed bis-silylation of internal alkynes with unsymmetrical disilanes is of significance. Herein, we solve most of the aforementioned limitations in bis-silylation of unsaturated bonds by developing a strongly coordinating disilane reagent and a Ni(0) catalytic system. Importantly, we sufficiently realize the stepwise recognition of the two silyl groups, making this synthetic protocol of wide potential utility.
Article
Full-text available
Heterocyclic compounds in particular siloles in which a carbon atom has been replaced by an isovalent silicon atom, has gained wide attention recently. Due to their unique and interesting electronic and aggregation‐induced emmision properties, they found practical application in many areas. The properties of siloles depend upon the nature of subsituents present on the ring. Therefore, the synthetic methods and strategies continued to evolve since their first synthetic report. The advances in the synthesis of siloles, mechanistic details, substituents effect and applications have been discussed in this review.
Article
Full-text available
Benzosiloles and their π-extended derivatives are present in many important advanced materials due to their excellent physical properties. Especially, they have found many potential applications in the development of novel electronic materials such as OLEDs, semiconductors and solar cells. In this review, we have summarized several main approaches to construct (di)benzosilole derivatives and (benzo)siloles fused to aromatic five- and six-membered heterocycles.
Article
Full-text available
The efficient and stereoselective synthesis of, or precursors to multi-substituted alkenes has attracted substantial interest due to their existence in various industrially and biologically important compounds. One of the most atom economical routes to such alkenes is the transition metal catalyzed hetero element-element’π-insertion into alkynes. This article provides a thorough up-to-date review on this area of chemistry, including discussions on the mechanism, range of E-E’ bonds accessible and the stoichiometric/catalytic transition metal mediators employed.
Article
The synthesis and ruthenium-catalyzed hydrosilylation reactions of 3-dimethylsilyl-2-ethynylpyridine derivatives (1a–1d) are described. The reaction of 1a and 1b in the presence of a ruthenium catalyst gave the corresponding pyridine-fused siloles (2a and 2b) through intramolecular hydrosilylation. The UV-Vis absorption and fluorescence properties of 2a and 2b were investigated, and DFT calculations were performed to understand their photophysical properties.
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
Silacyclobutanes (SCB), a family of useful organosilicon compounds, have attracted much attention because the highly strained silicon-carbon bond can be activated, as evidenced by transition-metal catalyzed ring-opening and ring expansion reactions. Efforts of extending silicon-carbon bond activation will definitely trigger powerful and privileged catalyst systems capable of unveiling opportunities for highly efficient and valuable transformations of SCB. This review summarized and highlighted representative and important advances for the synthesis of silacyclobutanes and its analogues as well as their catalytic transformations enabled by various transition-metal complexes whose versatility will be demonstrated widely in organosilicon chemistry and transition-metal catalysis. The different contributions in this growing research area summarized in this review could provide inspiration and incentive to make SCB as a precious source of organosilicon compounds.
Article
The air‐stable unsymmetrical disilane Me3Si−SiMe2[oCON(iPr)2C6H4] has been developed for bis‐silylation of alkynes. This reagent tolerates a range of functional groups, providing Z‐vinyl disilanes in high yields. It is proposed that the phenyl‐ring‐tethered amide group directs oxidative addition of Pd0 into the Si−Si bond, which might facilitate formation of a six‐membered Pd cycle, generating products with good to excellent regioselectivity. Open Si‐Si‐Me: Air‐stable asymmetric disilane, Me3Si−SiMe2[oCON(iPr)2C6H4], was developed for bis‐silylation of alkynes. A range of functional groups are tolerated, providing Z‐vinyl disilanes in high yields. A phenyl‐ring‐tethered amide group is proposed that directs oxidative addition of Pd0 into the Si−Si bond, forming a six‐membered Pd cycle and generating products with good to excellent regioselectivity. Key: regioselectivity ratio (r.r.).