Você está na página 1de 13

FULL PAPER

DOI: 10.1002/asia.201300872

Effect of the Substitution Pattern on the Intramolecular Charge-Transfer


Emissions in Organoboron-Based Biphenyls, Diphenylacetylenes, and
Stilbenes
Yi-Qiao Yan, Yan-Bang Li, Jian-Wu Wang,* and Cui-Hua Zhao*[a]

Abstract: Three series of organoboronbased molecules, including biphenyls


1 ac, diphenylacetylenes 2 ac, and stilbenes 3 ac, in which the electron-accepting boryl and the electron-donating
amino groups are introduced at different positions, have been comprehensively investigated to explore the effect
of the substitution pattern on the intramolecular charge-transfer emissions. In
cyclohexane solution, the change of
substitution pattern from p,p to o,p by
introduction of boryl at the lateral oposition rather than the terminal p-position leads to bathochromism in the
absorption and emission spectra. With
further variation of the amino position
from the terminal p-position in o,psubstitution to the lateral o-position in
an o,o-substitution pattern, a blueshift
was observed in the absorption owing

to the less-efficient conjugation extension of the amino group as the result of


sp3 hybridization. It is notable that the
emission of the three series of molecules changes with completely different
trends. Only the emission of the biphenyl is redshifted further from o,psubstituted 1 b to o,o-substituted 1 a,
whereas o,o-substituted diphenylacetylene 2 a maintains almost the same
spectrum as that of o,p-substituted diphenylacetylene 2 b and the fluorescence of o,o-substituted stilbene 3 a is
even blueshifted compared with o,psubstituted stilbene 3 b. As a result, the
Keywords: charge transfer fluorescence spectroscopy organoboron compounds photophysics
substituent effects

cimer formation and energy migration in the solid state.[4b, 5]


As a consequence, in contrast to the large number of molecules that are known to be highly emissive in solution, the
fluorescent molecules with excellent fluorescence efficiency
close to unity in the solid state are quite limited. Therefore,
the rational design of such a kind of emissive material is still
quite a challenging issue. In this context, it is of great interest to have a comprehensive investigation on the known
emissive systems to elucidate the structureproperty relationships and thus provide some basis for further rational
designs.
To obtain solid-state intensely fluorescent materials, several effective approaches have been developed, such as bulky
or dendritic substituent protection,[6] cross-dipole stacking,[7]
taking advantage of aggregation-induced emission,[810] J-aggregation formation,[11] the spiro concept,[12] and enhanced
intramolecular charge-transfer (CT) emission.[1315] Among
them, we and others have recently demonstrated an efficient
molecular design by introduction of bulky electron-accepting dimesitylboryl groups[1627] at the side positions of the
electron-donating p-conjugated framework.[13, 15b] This effec-

Introduction
With the growing advances in organic optoelectronic devices, such as organic light-emitting diodes (OLEDs),[1] organic solid-state lasers,[2] and organic fluorescent sensors,[3] it
has become one of the urgent research topics to obtain organic emissive materials that exhibit high fluorescence efficiency even in the solid state.[4] However, most organic fluorophores are highly emissive only in dilute solution and
weakly emissive or even nonemissive in the solid state
owing to the severe fluorescence quenching as a result of
certain intermolecular interactions, such as aggregate or ex[a] Y.-Q. Yan, Y.-B. Li, Prof. Dr. J.-W. Wang, Prof. Dr. C.-H. Zhao
School of Chemistry and Chemical Engineering
Shandong University
Shanda Nanlu 27, Jinan, 250100 (P.R. China)
Fax: (+ 86) 531-8856-4464
E-mail: jwwang@sdu.edu.cn
chzhao@sdu.edu.cn
Supporting information for this article is available on the WWW
under http://dx.doi.org/10.1002/asia.201300872.

Chem. Asian J. 2013, 8, 3164 3176

o,o-substituted biphenyl 1 a shows the


longest emission wavelength despite
the limited conjugation of the parent
biphenyl skeleton. The long emission
wavelength of 1 a may arise from its extremely twisted structure, which would
cause a significant structural relaxation
in the exited state. In the solid state, 1 a
still keeps almost the longest emission
wavelength. In addition, its quantum
yield is also among the highest. The unusual properties, intense solid-state
emission together with long emission
wavelength, and particularly large
Stokes shift, which are difficult to
attain by structural modification of
other parent p-conjugated frameworks,
have been achieved by the introduction
of boryl and amino groups at the o,opositions of the biphenyl skeleton.

3164

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

tive molecular design is ascribed to two kinds of effect of


the lateral boryl groups. One is the steric bulkiness that can
prevent the intermolecular interactions. The other is the
electron-accepting character that induces an intramolecular
CT transition with a large Stokes shift and thus efficiently
suppresses fluorescence self-quenching in the condensed
state. Based on this molecular design concept, we have
more recently disclosed another new class of CT-emitting organoboron compounds, in which the electron-accepting
boryl and the electron-donating amino groups are introduced at the o,o-positions of the biphenyl framework
(1 a).[28] In this boron-containing p system, the intramolecular CT transition takes place most likely through space
rather than through bonds owing to the particularly twisted
structure of the biphenyl skeleton and a very close contact
between the boryl and amino groups.[22eg] In addition to the
high fluorescence efficiency in the solid state, another two
features are notable for the photophysical properties of this
through-space CT-emitting organoboron p system. One is
the particularly large Stokes shift (Dl > 200 nm). The other
is the much longer emission wavelength compared with its
regioisomers (1 b and 1 c), which are modified with different
substitution patterns. The emissions of 1 a in cyclohexane
and in the spin-coated film are 112 and 45 nm longer than
the normally linear regioisomer 1 c, respectively, in which
the electron-donating amino and electron-accepting boryl
groups are attached at the terminal positions of the biphenyl
framework. Promoted by the intriguing photophysical properties of this through-space CT-emitting organoboron p-conjugated system, we were interested in the possibility to tune
the emission to a longer wavelength by choosing a parent
framework with more extended conjugation. The emission is
generally shifted to a longer wavelength with the conjugation extension of the parent framework for the normally
linear p-conjugated system. For example, the emission in cyclohexane shows a bathochromism from 409 nm of biphenyl
1 c[19c] to 418 nm of diphenylacetylene 2 c,[18i] and 450 nm of
stilbene 3 c.[18d, 19b, 20b] Unexpectedly, we found that the emission of biphenyl 1 a in cyclohexane (521 nm) is even much
longer than those of o,o-substituted diphenylacetylene 2 a
(462 nm) and stilbene 3 a (467 nm). These unusual phenomena have motivated us to have a comprehensive investigation on the regioisomers of the three series of p-conjugated
systems, biphenyls (1 ac), diphenylacetylenes (2 ac) and

Chem. Asian J. 2013, 8, 3164 3176

Jian-Wu Wang, Cui-Hua Zhao et al.

stilbenes (3 ac), in which boryl and amino groups are introduced at different positions. Herein, their single-crystal Xray structures, photophysical properties in solution and in
the solid state, and theoretical calculations were fully characterized to explore the substitution-pattern effect on the
photophysical properties and thus to elucidate the structureproperty relationships.

Results and Discussion


Synthesis
The synthetic routes to the target molecules are shown in
Scheme 1. Diphenylacetylene 2 a and stilbenes 3 a, b were
facilely prepared from their corresponding bromated precursors 2 a and 3 a, b, respectively. The lithiation of the bro-

Scheme 1. Synthesis of diphenylacetylene and stilbene derivatives. Reagents and conditions: a) [PdACHTUNGRE(PPh3)4], CuI, 1:3 Et3N/THF; b) i) nBuLi,
THF, 78 8C, 1 h; ii) Mes2BF, THF, 78 8C to RT; c) tBuONa, THF, 0 8C
to RT.

mated precursors with nBuLi at 78 8C followed by treatment with dimesitylfluoroborane provided the corresponding borylated compounds in moderate yields. The bromated
precursor 2 a was synthesized through the Pd0-catalyzed Sonogashira cross-coupling reaction of o-bromoiodobenzene
with 2-(N,N-dimethylamino)phenylethyne.[29] The Pd0-catalyzed Sonogashira coupling reaction of 2-(dimesitylboryl)phenylethyne[30] with 4-iodo-N,N-dimethylaniline under similar conditions afforded diphenylacetylene derivative 2 b.
The preparation of precursor 3 a involved a HornerWadsworth reaction of o-(dimethylamino)benzaldehyde with 2bromo-1-diethylphosphonomethylbenzene. Precursor 3 b[31]
and other boron-containing compounds 1 ac,[28, 19c] 2 c,[18i]
and 3 c[18d, 19b, 20b] were prepared according to the literature.
All the obtained organoboron compounds are stable in air

3165

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

and water and can be purified by silica-gel column chromatography. All the newly prepared compounds were fully
characterized using 1H and 13C NMR spectroscopy and highresolution mass spectrometry.

Jian-Wu Wang, Cui-Hua Zhao et al.

1 c). It is out of expectation that the two benzene rings P1


and P2 are almost perpendicular to each other with a torsion
angle up to 74.18 for o,p-substituted diphenylacetylene 2 b,
which is probably ascribed to the two benzene rings P1 and
P2 preferring to conjugate with two perpendicular p bonds
rather than the same p bond of triple bonds. Notably, o,osubstituted diphenylacetylene 2 a is even more coplanar
than 2 b although it contains one more amino substituent at
the lateral position (torsion angle between P1 and P2: 40.18
for 2 a). It is interesting to find that the main chain maintains basically coplanar irrespective of the substitution positions of boryl and amino groups for stilbene derivatives (torsion angle between P1 and P2: 13.18 for 3 a; 6.58 for 3 b;
20.38 for 3 c). 2) The relative positions between the boryl
and amino groups are quite different when they are introduced at o,o-positions. The boryl and amino groups lie on
the opposite sides of the main-chain axis in diphenylacetylene 2 a and stilbene 3 a, which is supposed to be beneficial
for minimizing the steric repulsion between them and thus
stabilizing the molecular structure. It is probably the repulsion between boryl and amino groups that also accounts for
the smaller dihedral angle between P2 and P1 in 2 a than
that in 2 b, since the smaller the dihedral angle between P1
and P2, the longer the distance between boryl and amino
groups and thus the weaker repulsion between them. On the
contrary, the boryl and amino groups are arranged at the
same side of the biphenyl axis in biphenyl 1 a with a very
close BN distance (3.59 ) in spite of the remarkable
steric congestion, which denotes the possible electronic attractions between boryl and amino groups. 3) The intramolecular pp stacking is observed in biphenyls 1 a, b and diphenylacetylene 2 b. In 1 a and 1 b, the dihedral angles between the benzene ring P3 of one mesityl and P1 ring are

X-ray Crystal Structure Analysis

The structures of the previously prepared biphenyl derivative 1 b and all the newly prepared diphenylacetylenes 2 a, b,
stilbenes 3 a, b were characterized by X-ray crystallography.
Unfortunately, all the efforts to prepare single crystals of 2 c
were unsuccessful. For a detailed comparison, the X-ray
crystal structures of all other compounds 1 a,[28] 1 c,[19c] and
3 c[20b] were also analyzed. The X-ray crystal structures are
shown in Figure 1 and the two benzene rings of the main
chain attached to boron and nitrogen atoms are labeled as
P1 and P2, respectively. In all these compounds, the central
boron and its three bonded carbon atoms are almost completely coplanar, which is indicative of the sp2-hybrid format
of the central boron atom. In addition, the three benzenering planes around the boron center are arranged in a propeller-like fashion, which accounts for the protection of the
trivalent boron atom and thus the high stabilities of these
compounds. Despite the planar structure of the trivalent
boron for all these compounds, the geometries of the nitrogen atom are quite different. When amino groups are attached at the terminal p-positions (1 b, c, 2 b, 3 b, c), the central nitrogen and its three bonded carbon atoms are basically coplanar and form a trigonal NC3 plane, which implies sp2
hybridization. In contrast, when amino groups are incorporated at the lateral o-positions (1 a3 a), the central nitrogen
displays a tetrahedral conformation as a result of the sp3 hybridization, which presumably would lead to the less-efficient delocalization of the lone
pair on nitrogen into the attached p systems compared
with sp2 hybridization.
In addition to the above
structural features of boryl and
amino groups, the following
three points are also noted.
1) With the variation of substitution pattern, the coplanarity
of the parent main chain
changes with different trends
for these three series of molecules. The biphenyl skeleton is
significantly twisted in 1 a with
a torsion angle of 70.78 due to
the great steric hindrance of lateral boryl and amino groups.
The biphenyl moiety becomes
increasingly coplanar with the
decreased steric congestion of
the lateral groups from 1 a to
1 c (torsion angle between P1 Figure 1. Crystal structures of a) biphenyls 1 ac; b) diphenylacetylenes 2 a, b, and c) stilbenes 3 ac. Hydrogen
and P2: 56.48 for 1 b; 21.18 for atoms are omitted for clarity.

Chem. Asian J. 2013, 8, 3164 3176

3166

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

Jian-Wu Wang, Cui-Hua Zhao et al.

27.48 for 1 a and 18.88 for 1 b, respectively. And the centroidcentroid distances are 3.87 and 3.64 , which suggests
possible p-orbital interactions between P3 and P1. In o,psubstituted diphenylacetylene 2 b, the orbital interactions between P3 and the triple bond most likely take place since
the triple bond is almost ideally parallel to the P3 ring with
a trivial angle of 14.58 and the centroidcentroid distance is
3.45 .
To gain insight into the molecular structure of 2 c, optimization of the molecular geometry was carried out using density functional theory (DFT) calculations at the B3LYP/631G(d) level of theory. In the obtained optimized structure
of 2 c, the parent main chain of diphenylacetylene is perfectly coplanar and both boryl and amino atoms display sp2trigonal conformation, as shown in the Supporting Information. The results derived from the structure analysis clearly
demonstrate that the substitution pattern has quite a different effect on the molecular structure of biphenyls, diphenylacetylenes, and stilbenes, which would lead to the different
effect on the electronic structure and thus the photophysical
properties.
Photophysical Properties in Cyclohexane
The UV/Vis absorption and emission spectra were first measured in cyclohexane and are shown in Figure 2. The corresponding data are summarized in Table 1. In the absorption
spectra, the normal linear p,p-substituted biphenyl 1 c has
an intense absorption band at 369 nm (log e = 4.50), which
mainly consists of an intramolecular CT transition from the
HOMO delocalized over the aminobiphenyl framework to
the LUMO located on the borylbiphenyl moiety.[28] When
the boryl site is changed from a p- (1 c) to o-position (1 b),
the CT absorption is redshifted by 19 nm, which is accompanied by a significant decrease in absorption intensity (log e =
3.48). It is noteworthy that the variation of the amino position from the p- (1 b) to the o-position (1 a) leads to a remarkable hypochromism in absorption and o,o-substituted
biphenyl 1 a only displays an intense absorption band at
306 nm (log e = 4.74), which is in fact assignable to the transition from HOMO1 located on the filled p orbitals of the
mesityl and phenyl rings to the LUMO localized on the dimesitylborylphenyl moiety. The intensity of the lowest excited-state transition, corresponding to an intramolecular CT
transition from the HOMO localized on the dimethylaminophenyl unit to the LUMO (oscillator strength f = 0.0048), is
too weak to be distinguished.[28] Similarly, the substitution
position has a magnificent effect on the absorption of the
other two series of compounds, diphenylacetylenes and stilbenes. In addition, the absorption changes in a similar way
to that of biphenyls. And thus the normal linear p,p-substituted derivatives show the most intense intramolecular CT
absorption band (2 c: labs = 377 nm, log e = 4.48; 3 c: labs =
393 nm, log e = 4.29). The o,p-substituted compounds (2 b,
3 b), in which a boryl group is introduced at the lateral o-position, display redshifted absorptions with lower intensity
than those of p,p-substituted ones (2 b: labs = 384 nm, log e =

Chem. Asian J. 2013, 8, 3164 3176

Figure 2. Absorption and fluorescence spectra of a) biphenyls 1 ac, b) diphenylacetylenes 2 ac, and c) stilbenes 3 ac in cyclohexane.

3.91; 3 b: labs = 402 nm, log e = 3.92). With the change of


amino group from the p- (2 b, 3 b) to o-position (2 a, 3 a),
the absorption maxima exhibit a blueshift while the intensities remain almost the same (2 a: labs = 368 nm, log e = 3.77;
3 a: labs = 371 nm, log e = 3.90). As a result, the intramolecular CT absorption bands for o,o-substituted compounds 2 a
and 3 a were observed at the shortest wavelength among
their three regioisomers of the same series. It is worth
noting that the intramolecular CT bands of 2 a and 3 a are
readily distinguished, in sharp contrast to the invisible CT
transition absorption of 1 a.

3167

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

Table 1. UV/Vis absorption and fluorescence data for biphenyls 1 ac, diphenylacetylenes 2 ac, and stilbenes 3 ac in cyclohexane.
Absorption
log e
labs [nm][a]
1a
1b
1c
2a
2b
2c
3a
3b
3c

306
388
369
368[c]
384
377
371
402
393

4.74
3.48
4.50
3.77
3.91
4.48
3.90
3.92
4.29

Fluorescence
lem [nm]
FF[b]
521
477
409
462
462
418
467
483
450

0.47
0.72
0.95
0.34
0.35
0.99
0.79
0.87
0.60

Stokes shift
Dl [nm]
Dn [cm1]
215
89
40
94
78
41
96
81
57

13 485
4808
2650
5528
4396
2601
5540
4171
3223

Jian-Wu Wang, Cui-Hua Zhao et al.

pounds are striking for the biphenyl parent skeleton


with such a limited conjugation length, which illustrates the special effect of this molecular modification.
Theoretical Studies

To gain deeper insight into the effect of the substitution pattern on the photophysical properties,
which are highly relevant to the electronic structures, we conducted theoretical calculations. The
molecular geometry of 2 c was optimized using
[a] Only the longest absorption maximum wavelengths are given. [b] Calculated using
DFT calculations at the B3LYP/6-31G(d) level of
quinine sulfate as standard. [c] Observed as a shoulder.
theory owing to the absence of its single X-ray
structure. For all the other compounds, we perIn the fluorescence spectra, the normally linear p,p-subformed single-point calculations at the B3LYP/6-31G(d)
stituted biphenyl 1 c exhibits intense purple-blue fluoreslevel of theory using the geometries derived from their cryscence at 409 nm (FF = 0.95). The emission is gradually shifttal structures. We also carried out time-dependent densityed to longer wavelength as the substitution pattern changes
functional theory (TD-DFT) calculations of all the moleto o,p- and o,o-substitution (lem = 477 nm for 1 b; 521 nm
cules at the B3LYP/6-31G(d) level of theory to understand
for 1 a). It is intriguing to find that the emission wavelength
their intramolecular CT transitions. The pictorial drawings
of 1 a is the longest among the three regioisomers of biphenof their molecular orbitals and the KohnSham HOMO and
yls, 112 nm longer than that of 1 c. Although the p,p-substiLUMO energy levels are shown in Figure 3.
tuted regioisomers 2 c and 3 c also display the shortest emisIn the p,p-substituted molecules 1 c3 c, the HOMO is desion wavelength for the series of diphenylacetylenes and stillocalized over the entire parent main chain, including the dibenes, the extent of the bathochromism is much less signifimethylamino group. And the LUMO can still spread over
cant than that of the biphenyls (Dl = 68 nm from 1 c to 1 b;
the whole main chain with remarkable contribution from
44 nm from 2 c to 2 b; 33 nm from 3 c to 3 b) with the change
the dimesitylboryl group and little contribution from benof boryl position from the terminal p-position to the lateral
zene P2 attached to the amino atom. All the o,p-substituted
o-position. Moreover, from the o,p- and o,o-substitution
compounds 1 b3 b show greatly elevated HOMO levels
pattern, no further obvious change was detected for the dicompared with their corresponding p,p-substituted rephenylacetylenes and the emission spectra of 2 a and 2 b are
gioisomers. Notably, the benzene ring P3 in 1 b2 b contribvery close to each other. A blueshift of 16 nm in the fluoresutes to the HOMO to some extent owing to the intramoleccence of stilbenes was even observed from o,p-substituted
ular pp stacking, as evidenced by the single-crystal struc3 b to o,o-substituted 3 a. It is most noteworthy that o,otures. Although the HOMO of 3 b maintains spreading over
substituted biphenyl 1 a displays fluorescence at the longest
the whole main chain, the contribution of the benzene ring
wavelength among o,o-substituted molecules, even among
P1 bonded to the boron center becomes much less signifithese nine boron-containing compounds, whereas the
cant in 1 b2 b owing to their twisted main-chain structure,
normal linear p,p-substituted compounds exhibit a gradual
which probably also accounts for their greatly elevated
bathochromism in fluorescence from 1 c to 3 c with the exLUMO energy levels compared with corresponding p,p-substituted regioisomers despite that the LUMO level of 3 b is
tended conjugation of the parent main chain from biphenyl
only slightly higher than that of 3 c. The change of substituto diphenylacetylene and stilbene (see the Supporting Infortion pattern from o,p- to o,o-substitution leads to signifimation). In view of the fluorescence efficiency, the quantum
cantly low-lying HOMOs in 1 a3 a as the result of less-effiyield is highly dependent on the planarity of the main-chain
cient electron-donating ability of the sp3-hybridized amino
framework. And thus the linear compounds 1 c and 2 c have
the highest fluorescence quantum yields among their regroup at the lateral position than that of the sp2-hybridized
gioisomers and the fluorescence efficiencies of stilbene deamino group at the terminal position. In addition, the great
rivatives are very high irrespective of the substitution patdecrease in the LUMO energy levels was observed in 2 a3 a
tern. This is easily understandable considering that the
relative to those of 2 b3 b, whereas the LUMO of 1 a is very
planar structure of the main chain generally tends to supclose to that of 1 b. These results clearly demonstrate that
press the nonradiative decay of the excited state in solution.
the change in the substitution position of either the boryl or
Another notable common feature for these three series of p
amino substituent has a great effect on the HOMO and the
systems is that the Stoke shift is gradually enlarged from
influence on the LUMO is highly dependent on the molecua p,p- to o,p- and o,o-substitution pattern. In particular,
lar structure. The TD-DFT calculations indicate that the
the Stokes shift of 1 a is rather large (Dl = 215 nm, Dn =
first excited state for all these molecules are assignable to
13 485 cm1). The properties of the longest emission wavethe intramolecular CT transition from the HOMO to the
LUMO. The charge transfer takes place most likely through
length and largest Stokes shift of 1 a among these nine com-

Chem. Asian J. 2013, 8, 3164 3176

3168

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Jian-Wu Wang, Cui-Hua Zhao et al.

www.chemasianj.org

is noted that the oscillator strength of the first excited state


for 1 a is particularly low, which presumably makes the corresponding intramolecular CT transition band too weak to
be distinguished in the absorption and thus 1 a displays the
rather large Stokes shift. Whereas the accuracy of the calculated frontier energy levels and the calculated excitation
energy is not sufficiently high by this level of calculations,
these calculated results apparently indicate that the modification of the same parent skeleton with different substitution patterns has a remarkable effect on the electronic structure and thus the photophysical properties. The first excitedstate transition occurs most efficiently for the linear p,psubstituted regioisomers, which is in good agreement with
the order of the experimentally obtained molecular absorption coefficients within the same series.
Fluorescence Solvatochromism
The most notable feature of the regioisomers for these three
series of organoboron compounds is their intramolecular CT
character in the transitions. To gain a deeper insight into
their excited states, we investigated the solvent effects on
their absorption and emission spectra, the data for which
are summarized in Table 2.
The fluorescence of all these molecules displays a significant redshift as the solvent polarity increases, whereas the
absorption spectra display negligible solvent dependence
(see the Supporting Information). These phenomena suggest
that these compounds have more polar structures in the excited state than those in the ground state, which is characteristic of organoboron molecules exhibiting intramolecular CT
transitions. To elucidate the degree of the polar excited
state, we employed the LippertMataga equation
[Eq. (1)],[32] in which h (= 6.6256  1027 erg) is Planks constant, c (= 2.9979  1010 cm s1) is the speed of light, and a is
the radius of the cavity, me and mg are the dipole moments in
the excited and ground state, respectively, and Dn is the
Stokes shift. Df is the solvent polarity and is given by Equation (2), in which e is the dielectric constant and n is the optical refractive constant.
Figure 3. Plot of KohnSham HOMO and LUMO energy levels and pictorial drawings of the HOMOs and LUMOs for a) biphenyls 1 ac, b) diphenylacetylenes 2 ac, and c) stilbenes 3 ac. The transition energies and
oscillator strengths were calculated at the B3LYP/6-31G(d) level of
theory.

Dn nA  nF

Df

e1
n2  1
 2
2e 1 2n 1

The LippertMataga equation accounts for the general solvent effect and does not account for specific solventfluorophore interactions, for example, through hydrogen bonding.
We indeed obtained linear relationships for the plots of Dn
as a function of Df for all of these organoboron compounds,
as shown in Figure 4. From the slope of these plots, the
change in the dipole moment (memg) of the fluorophore
upon electronic excitation was estimated assuming the molecular radius as the cavity radius.[33] The molecules under
consideration are nonspherical in nature and so the above

space rather than through bonds in 1 a as the result of no


overlap between the HOMO and LUMO orbitals and
a close contact between boryl and amino groups. On the
contrary, the intramolecular CT transition happens by
means of the corresponding aromatic linker for all the other
compounds. In addition, the oscillator strengths of linear
p,p-substituted 1 c3 c are the highest among their regioisomers and the o,o-substituted 2 a3 a have an oscillator
strength of the intramolecular CT transition at the same
level as their corresponding o,p-substituted regioisomers. It

Chem. Asian J. 2013, 8, 3164 3176

2Df
m  mg 2 C
hca3 e

3169

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

Table 2. UV/Vis absorption and fluorescence data of biphenyls 1 ac, diphenylacetylenes 2 ac, and stilbenes 3 ac in various solvents.

1a

1b

1c

Solvents

labs [nm][a]

cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN
cyclohexane
CHCl3
THF
MeCN

306
307
307
303
388
385
395
386
369
371
375
375
368[e]
375[e]
375[e]
375[e]
384
384
388
384
377
386
386
383
371
368
362
370
402
399
406
399
393
398
405
403

lem [nm]
521
547
551
580
477
520
541
570
409
459
488
528
462
507
532
563
462
505
534
566
418
482
518
561
467
505
526
559
483
526
552
572
450
508
534
567

Dn [cm1][b]
13 485
14 291
14 424
15 761
4808
6743
6832
8362
2650
5167
6174
7727
5528
6942
7869
8904
4396
6239
7046
8373
2601
5159
6601
8284
5540
7371
8612
9137
4171
6051
6514
7580
3223
5440
5964
7177

a [][c]

Dm [D][d]

7.73

17.3

7.42

21.1

7.83

27.9

Jian-Wu Wang, Cui-Hua Zhao et al.

substitution. It is interesting to note that 1 a displays


the smallest dipole change from the ground state to
the excited state, whereas its emission wavelength is
the longest among these nine compounds. The long
emission wavelength of 1 a may arise from its extremely twisted structure, which would cause a significant structural relaxation in the exited state.
Photophysical Properties in the Solid State

Another intriguing property of o,o-substituted biphenyl 1 a is its intense fluorescence in the solid
state. So we next investigated the effect of the sub7.59
21.8
2a
stitution pattern on the solid-state fluorescence
properties for the three series of compounds. Thin
films of all the molecules were prepared from their
solutions in dichloromethane with approximately
2b
7.72
24.1
3 mg mL1 concentration on the quartz plates and
their absorption and emission spectra were directly
measured. The absolute fluorescence quantum
2c
7.83
29.7
yields were determined by a calibrated integrating
sphere system. The corresponding spectra are
shown in Figure 5, and the data are summarized in
3a
7.63
23.1
Table 3.
In the absorption spectra, the o,o-substituted
1 a, 2 a and o,p-substituted 1 b, 2 b maintain almost
3b
7.95
23.3
the same spectra as those in solution (redshift from
cyclohexane to spin-coated film of 2 nm for 1 a;
11 nm for 2 a; 8 nm for 1 b; and 12 nm for 2 b).
These results suggest no formation of aggregates in
3c
8.13
26.0
the ground state, which is rational considering that
the steric effect of the bulky boryl group and the
nonplanar structure of the main-chain skeleton
[a] Only the longest absorption maximum wavelengths are given. [b] Stokes shift.
would prevent the intermolecular interactions in
[c] Calculated radius of cavity using the X-ray single-crystal structure. [d] Calculated
the solid state. On the contrary, the p,p-substituted
dipole moment change (mgme) from the ground state to the excited state. [e] Ob1
c3 c display obvious bathochromism from soluserved as a shoulder.
tion to the spin-coated film (redshift from cyclohexane to spin-coated film of 20 nm for 1 c; 24 nm for
2 c; and 24 nm for 3 c), which is indicative of stronger intermolecular interactions. Notably, the absorptions of
assumption of the molecular radius for the cavity radius is
o,o-substituted 3 a and o,p-substituted 3 b exhibit much less
only approximate. The molecular radii for all the comof a redshift from the cyclohexane solution to the spinpounds were estimated from the DFT calculations and are
coated film compared with that of 3 c, even though their
summarized in Table 2. The changes in the dipole moments
parent diphenylacetylene skeletons are almost perfectly cofrom the ground state to the excited state were calculated
planar (redshift from cyclohexane to spin-coated film of
and are also listed in Table 2. The remarkable changes of
2 nm for 3 a; and 15 nm for 3 b), which illustrates that the indipole moments upon photonic excitation were observed for
troduction of a boryl group at the lateral o-position instead
all these molecules, thereby confirming their emissions from
of terminal p-position is more efficient for suppressing interthe intramolecular CT excited state and the higher polarity
molecular interactions in the solid state even when it has
in the excited state than in the ground state. In addition, the
a trivial effect on the planarity of the main chain structure.
p,p-substituted regioisomers show the largest dipoleIn the fluorescence spectra, it is interesting to find that
moment changes, whereas the lowest dipole-moment
from o,o-substituted molecules to the corresponding p,pchanges were found for o,o-substituted regioisomers. This is
substituted regioisomers, the emissions are closer to those in
reasonable considering that an electron is transferred from
the more polar solvents (see the Supporting Information).
the amino nitrogen to the boryl boron during the intramoThus compared with the fluorescence in the nonpolar cyclolecular CT transition and thus the distance between the poshexane solutions, o,o-substituted regioisomers 1 a3 a disitive and negative center becomes gradually longer when
the substitution mode changed from o,o- to o,p- and p,pplay the least bathochromism in the solid state, whereas the

Chem. Asian J. 2013, 8, 3164 3176

3170

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

Jian-Wu Wang, Cui-Hua Zhao et al.

Figure 4. The LippertMataga plots of a) biphenyls: 1 a (&), 1 b (*), 1 c


(~); b) diphenylacetylenes: 2 a (&), 2 b (*), 2 c (~); and c) stilbenes 3 a
(&), 3 b (*), 3 c (~).

redshift of p,p-substituted regioisomers are most significant


(redshift of emission from solution in cyclohexane to spincoated film of 2 nm for 1 a; 30 nm for 1 b; 61 nm for 1 c;
37 nm for 2 a; 42 nm for 2 b; 78 nm for 2 c; 34 nm for 3 a;
43 nm for 3 b; and 74 nm for 3 c). Another interesting finding in the emission is that the fluorescence quantum yield
gradually increases with the introduction of more substituents at the lateral positions, which is probably owing to
gradually enlarged Stokes shifts. As a result, all the o,o-substituted molecules display intense fluorescence with excellent quantum yields in the spin-coated film (FF = 0.86 for
1 a; 0.78 for 2 a; 0.87 for 3 a). Regarding the emission spectra in the spin-coated film, the fluorescence maxima of 3 b
and 3 c are very close to that of 1 a. However, their emission
efficiencies are quite a bit lower than that of 1 a. So the photophysical properties of 1 a are very unique among all these

Chem. Asian J. 2013, 8, 3164 3176

Figure 5. Absorption and fluorescence spectra of a) biphenyls 1 ac, b) diphenylacetylenes 2 ac, and c) stilbenes 3 ac in spin-coated films.

molecules owing to its very long emission wavelength, high


solid-state fluorescence quantum yield, and particularly
large Stokes shift.

Conclusion
We have conducted a comprehensive investigation on three
series of organoboron molecules, biphenyls 1 ac, diphenylacetylenes 2 ac and stilbenes 3 ac, in which the electron-accepting dimesitylboryl group and electron-donating dimethylamino group are introduced at different positions. Their

3171

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

Table 3. UV/Vis absorption and fluorescence data for biphenyls 1 ac, diphenylacetylenes 2 ac, and stilbenes 3 ac in the spin-coated films.
labs [nm][a]
1a
1b
1c
2a
2b
2c
3a
3b
3c

308
396
389
379[c]
396
401
373
417
417

Fluorescence
FF[b]
lem [nm]
523
507
470
499
504
496
501
526
524

0.86
0.66
0.65
0.78
0.64
0.53
0.87
0.57
0.44

Experimental Section
General procedure

Stokes shift
Dl [nm]
Dn [cm1]
215
111
81
120
108
95
128
109
107

Jian-Wu Wang, Cui-Hua Zhao et al.

Melting points were measured on a Tektronix XT-4 instrument. 1H and


13
C NMR spectra were recorded with a Bruker 300 spectrometer or
a Bruker 400 spectrometer for samples in CDCl3. UV-visible absorption
spectra and fluorescence spectra measurements were performed with
a Hitachi UV-4100 spectrometer and a Hitachi F-4500 spectrometer, respectively. Spin-coating was carried out using JYL YJ-150 at room temperature. The spin-coated films were prepared by spinning the solutions
of the samples in dichloromethane (3 mg mL1) onto quartz plates at
1000 rpm for 30 s. The solid-state quantum yields were measured from
the freshly spin-coated film using an integrating sphere with an excitation
wavelength of 350 nm. All reactions were carried out under a nitrogen atmosphere. Compounds 1 a, b and 1 c3 c were prepared according to the
literature. All calculations were conducted by using the Gaussian 09 program.[34]

13 347
5528
4430
6345
5411
4776
6849
4969
4896

[a] Only the longest absorption maximum wavelengths are given. [b] Absolute quantum yields determined by a calibrated integrating sphere
system. [c] Observed as a shoulder.

2-Bromo-2-(N,N-dimethylamino)diphenylacetylene (2 a)
A degassed mixed solvent of Et3N/THF 1:3 (80 mL) was added to a solution of 2-bromoiodobenzene (7.55 g, 26.7 mmol), 2-(N,N-dimethylamino)phenylethyne (2.57 g, 17.8 mmol), [PdACHTUNGRE(PPh3)4] (0.12 g, 0. 1 mmol), and
CuI (0.08 g, 0.4 mmol). The reaction mixture was stirred at 50 8C overnight. The solvents were removed under reduced pressure. After addition
of CH2Cl2, the mixture was washed successively with a saturated NH4Cl
aqueous solution, and brine. The organic layer was dried over anhydrous
Na2SO4, filtered, and concentrated under reduced pressure. The resulting
mixture was subjected to silica gel column chromatography (4:1 petroleum ether/CH2Cl2, Rf = 0.26) to afford 2 a (2.37 g, 7.90 mmol) in 47 %
yield as a pale yellow liquid. 1H NMR (CDCl3, 300 MHz): d = 7.537.61
(m, 3 H), 7.247.30 (m, 2 H), 7.15 (td, J = 7.5, 1.8 Hz, 1 H), 6.876.94 (m,
2 H), 3.02 ppm (s, 6 H); 13C NMR (CDCl3, 75 MHz): d = 154.6, 134.8,
133.1, 132.4, 129.7, 129.0, 126.9, 126.0, 125.2, 120.3, 116.9, 114.5, 93.4,
93.2, 43.7 ppm; HRMS (ESI): m/z calcd for C16H15BrN: 300.0388; found:
300.0370 [M+H] + .

single-crystal X-ray structures, photophysical properties in


solution and the solid state, and theoretical calculations
were fully characterized to explore the substitution-pattern
effect on the photophysical properties. The change of the
substitution pattern from normal linear p,p- to o,p-substitution by the introduction of boryl at the lateral o- rather than
terminal o-position leads to the bathochromism in the absorption and emission spectra. With the further introduction
of an amino group at the lateral position in the o,o-substitution pattern, a blueshift was observed in the absorption
owing to the less-efficient conjugation extension of the
amino group as the result of sp3 hybridization. Notably, the
emission of the three series of molecules changes with completely different trends. The emission of the biphenyl is further redshifted from o,p-substituted 1 b to o,o-substituted
1 a, whereas o,o-substituted diphenylacetylene 2 a maintains
almost same spectrum as that of o,o-substituted diphenylacetylene 2 a, and o,o-substituted stilbene 3 a is even blueshifted compared with o,p-substituted stilbene 3 b. So the
o,o-substituted biphenyl 1 a shows the longest emission
wavelength in the solution. The long emission wavelength of
1 a may arise from its extremely twisted structure, which
would cause a significant structural relaxation in the excited
state. In the solid state, 1 a still keeps almost the longest
emission wavelength among these compounds. It is notable
that its quantum yield is also among the highest. As a result,
introduction of boryl and amino groups at the o,o-positions
of the biphenyl skeleton leads to unusual and interesting
properties, long emission wavelength, together with high
solid-state fluorescence efficiency, and particularly large
Stokes shift, which are difficult to attain by structural modification of other parent p-conjugated frameworks. The further functionalization of the o,o-substituted biphenyl skeleton to explore its application in optoelectronic fields by utilizing its unique properties is under way in our group.

Chem. Asian J. 2013, 8, 3164 3176

2-(Dimesitylboryl)-2-(N,N-dimethylamino)diphenylacetylene (2 a)
A solution of nBuLi in hexane (0.48 mL, 2.5 m, 1.2 mmol) was added
dropwise by syringe at 78 8C to a solution of 2 a (0.30 g, 1.0 mmol) in
anhydrous THF (20 mL). The mixture was stirred at the same temperature for 1 h. A solution of dimesitylboron fluoride (1.07 g, 4.0 mmol) in
anhydrous THF (5 mL) was added to the reaction mixture using a syringe.
The reaction mixture was warmed to room temperature and stirred overnight. The reaction was quenched with a saturated solution of NaCl and
the aqueous layer was extracted with CH2Cl2. The combined organic
layer was dried over anhydrous Na2SO4, filtered, and concentrated under
reduced pressure. The resulting mixture was subjected to silica-gel
column chromatography (5:1 petroleum ether/CH2Cl2, Rf = 0.16) to
afford 2 a (170 mg, 0.36 mmol) in 36 % yield as a green solid. M.p. 138
140 8C; 1H NMR: d = 7.56 (d, J = 7.5 Hz, 1 H), 7.36 (td, J = 7.5, 2.1 Hz,
1 H), 7.237.28 (m, 2 H), 7.13 (t, J = 7.2 Hz, 1 H), 6.78 (d, J = 8.1 Hz, 1 H),
6.75 (s, 4 H), 6.68 (t, J = 7.5 Hz, 1 H), 6.59 (d, J = 7.2 Hz, 1 H), 2.87 (s,
6 H), 2.24 (s, 6 H), 2.01 ppm (s, 12 H); 13C NMR (CDCl3, 75 MHz): d =
154.3, 149.8, 142.8, 140.8, 138.9, 134.6, 134.4, 132.4, 130.1, 128.7, 128.3,
127.9, 127.5, 119.7, 116.3, 115.2, 95.6, 91.9, 43.3, 23.2, 21.2 ppm; HRMS
(ESI): m/z calcd for C34H3711BN: 470.3019; found: 470.3009 [M+H] + .
2-(Dimesitylboryl)-4-(N,N-dimethylamino)diphenylacetylene (2 b)
This compound was prepared essentially in the same manner as described
for 2 a by using 4-iodo-N,N-dimethylaniline (0.19 g, 0.75 mmol), 2-dimesitylborylphenylacetylene (0.18 g, 0.50 mmol), [PdACHTUNGRE(PPh3)4] (0.012 g, 0.
01 mmol), and CuI (0.008 g, 0.04 mmol) in Et3N/THF 1:3 (20 mL) at
room temperature. Purification by silica-gel column chromatography (5:1
petroleum ether/CH2Cl2, Rf = 0.20) afforded 2 b (0.15 g, 0.32 mmol) in
64 % yield as a green solid. M.p. 145147 8C; 1H NMR (CDCl3,
400 MHz): d = 7.50 (d, J = 7.7 Hz, 1 H), 7.34 (td, J = 6.9, 1.8 Hz, 1 H),
7.187.26 (m, 2 H), 6.92 (d, J = 8.4 Hz, 2 H), 6.76 (s, 4 H), 6.59 (br, 2 H),
2.97 (s, 6 H), 2.27 (s, 6 H), 2.01 ppm (s, 12 H); 13C NMR (CDCl3,

3172

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

X-ray Crystal Structure Analysis of 2 a[35]

100 MHz): d = 149.7, 143.0, 140.9, 138.8, 132.8, 132.5, 134.3, 130.1, 128.2,
128.1, 127.0, 111.6, 93.7, 88.3, 40.3, 23.2, 21.3 ppm; HRMS (ESI): m/z
calcd for C34H3711BN: 470.3019; found: 470.3027 [M+H] + .

Single crystals of 2 a suitable for X-ray crystal analysis were obtained by


recrystallization from a MeOH/CH2Cl2 mixed solvent. Intensity data
were collected at 293 K on a Bruker SMART CCD X-ray diffractometer
(APEX II) with MoKa radiation (l = 0.71073 ) and graphite monochromator. A total of 6882 reflections were measured at a maximum 2q angle
of 50.08, of which 4779 were independent reflections (Rint = 0.0278). The
structure was solved by direct methods (SHELXS-97)[36] and refined by
full-matrix least-squares cycles on F2 (SHELEXL-97).[36] All non-hydrogen atoms were refined anisotropically and all hydrogen atoms except for
those of the disordered solvent molecules were placed using AFIX instructions. The crystal data are as follows: C32H36BN; Mr = 445.43; crystal
size 0.20  0.18  0.10 mm3, triclinic, P1, a = 8.3328(11) , b =
Z = 2,
1calcd =
12.0347(17) ,
c = 14.628(2) ,
V = 1375.9(3) 3,
3
1.133 g cm . The refinement converged to R1 = 0.0607, wR2 = 0.1561 (I >
2s(I)), GOF = 1.027.

2-Bromo-2-(N,N-dimethylamino)stilbene (3 a)
A solution of tBuONa (0.96 g, 10 mmol) in anhydrous THF (15 mL) was
added to a solution of 2-bromo-1-diethylphosphonomethylbenzene
(1.35 g, 4.4 mmol) and 2-dimethylaminobenzaldehyde (0.30 g, 2 mmol) in
anhydrous THF (20 mL) by syringe at 0 8C. The reaction mixture was
warmed to room temperature and stirred overnight. The reaction was
quenched with a saturated solution of NaCl and the aqueous layer was
extracted with CH2Cl2. The combined organic layer was dried over anhydrous Na2SO4, filtered, and concentrated under reduced pressure. The resulting mixture was subjected to silica-gel column chromatography (4:1
petroleum ether/CH2Cl2, Rf = 0.33) to afford 3 a (0.47 g, 1.56 mmol) in
77 % yield as a white solid. M.p. 5961 8C; 1H NMR (CDCl3, 300 MHz):
d = 7.76 (d, J = 6.6 Hz, 1 H), 7.64 (d, J = 8.1 Hz, 1 H), 7.58 (dd, J = 7.8,
0.9 Hz, 1 H), 7.41 (s, 2 H), 7.287.34 (m, 2 H), 7.067.14 (m, 3 H),
2.78 ppm (s, 6 H); 13C NMR (CDCl3, 75 MHz): d = 152.3, 137.8, 133.0,
131.1, 129.5, 128.7, 128.5, 127.5, 127.4, 126.8, 126.7, 124.1, 122.5, 118.1,
44.8 ppm; HRMS (ESI): m/z calcd for C16H17BrN: 302.0544; found:
302.0530 [M+H] + .

X-ray Crystal Structure Analysis of 2 b[35]


Single crystals of 2 b suitable for X-ray crystal analysis were obtained by
recrystallization from a MeOH/CH2Cl2 mixed solvent. Intensity data
were collected at 298 K on a Bruker SMART CCD X-ray diffractometer
(APEX II) with MoKa radiation (l = 0.71073 ) and graphite monochromator. A total of 6822 reflections were measured at a maximum 2q angle
of 50.08, of which 4812 were independent reflections (Rint = 0.0237). The
structure was solved by direct methods (SHELXS-97)[36] and refined by
full-matrix least-squares cycles on F2 (SHELEXL-97).[36] All non-hydrogen atoms were refined anisotropically and all hydrogen atoms except for
those of the disordered solvent molecules were placed using AFIX instructions. The crystal data are as follows: C34H36BN; Mr = 469.45; crystal
size 0.20  0.18  0.05 mm3, triclinic, P1, a = 8.202(7) , b = 8.393(8) , c =
22.389(19) , V = 1389(2) 3, Z = 2, 1calcd = 1.122 g cm3. The refinement
converged to R1 = 0.0584, wR2 = 0.1642 (I > 2s(I)), GOF = 1.004.

2-(Dimesitylboryl)-2-(N,N-dimethylamino)stilbene (3 a)
This compound was prepared essentially in the same manner as described
for 2 a using compound 3 a (0.30 g, 1.0 mmol), anhydrous THF (20 mL),
nBuLi (0.48 mL, 2.5 m, 1.2 mmol), and dimesitylboron fluoride (1.07 g,
4.0 mmol). The purification by silica-gel column chromatography (5:1 petroleum ether/CH2Cl2, Rf = 0.21) afforded 3 a (0.25 g, 0.53 mmol) in 53 %
yield as a green solid. M.p. 150152 8C; 1H NMR (CDCl3, 300 MHz): d =
7.75 (d, J = 7.8 Hz, 1 H), 7.43 (t, J = 7.2 Hz, 1 H), 7.037.30 (m, 5 H), 6.93
(d, J = 7.8 Hz, 1 H), 6.746.79 (m, 5 H), 6.44 (d, J = 7.5 Hz, 1 H), 2.72 (s,
6 H), 2.28 (s, 6 H), 2.00 ppm (s, 12 H); 13C NMR (CDCl3, 75 MHz): d =
151.7, 147.1, 143.6, 143.1, 140.5, 139.0, 135.4, 131.7, 131.2, 129.3, 128.5,
128.3, 127.7, 127.2, 127.0, 124.8, 121.9, 117.5, 44.6, 23.1, 21.2 ppm; HRMS
(ESI): m/z calcd for C34H3911BN: 472.3176; found: 472.3203 [M+H] + .

X-ray Crystal Structure Analysis of 3 a[35]


Single crystals of 3 a suitable for X-ray crystal analysis were obtained by
recrystallization from a MeOH/CH2Cl2 mixed solvent. Intensity data
were collected at 298 K on a Bruker SMART CCD X-ray diffractometer
(APEX II) with MoKa radiation (l = 0.71073 ) and graphite monochromator. A total of 13 498 reflections were measured at a maximum 2q
angle of 50.08, of which 4979 were independent reflections (Rint = 0.0660).
The structure was solved by direct methods (SHELXS-97)[36] and refined
by full-matrix least-squares cycles on F2 (SHELEXL-97).[36] All non-hydrogen atoms were refined anisotropically and all hydrogen atoms except
for those of the disordered solvent molecules were placed using AFIX instructions. The crystal data are as follows: C34H38BN; Mr = 471.46; crystal
size 0.20  0.15  0.15 mm3, monoclinic, P21/n, a = 15.525(19) , b =
9.443(12) , c = 19.35(2 , V=2830(6) 3, Z = 4, 1calcd = 1.107 g cm3. The
refinement converged to R1 = 0.0635, wR2 = 0.1713 (I > 2s(I)), GOF =
0.950.

2-(Dimesitylboryl)-4-(N,N-dimethylamino)stilbene (3 b)
This compound was prepared essentially in the same manner as described
for 2 a using 2-bromo-4-(N,N-dimethylamino)stilbene (0.30 g, 1.0 mmol),
anhydrous THF (20 mL), nBuLi (0.48 mL, 2.5 m, 1.2 mmol), and dimesitylboron fluoride (1.07 g, 4.0 mmol). The purification by silica-gel column
chromatography (5:1 petroleum ether/CH2Cl2, Rf = 0.19) afforded 3 b
(0.15 g, 0.32 mmol) in 32 % yield as a green solid. M.p. 155156 8C;
1
H NMR (CDCl3, 400 MHz): d = 7.69 (d, J = 7.84 Hz, 1 H), 7.41 (t, J =
7.4 Hz, 1 H), 7.287.30 (m, 1 H), 7.17 (t, J = 7.3 Hz, 1 H), 7.04 (d, J =
16.1 Hz, 1 H), 6.89 (d, J = 8.5 Hz, 2 H), 6.816.85 (m, 5 H), 6.62 (br, 2 H),
2.98 (s, 6 H), 2.31 (s, 6 H), 2.04 ppm (s, 12 H); 13C NMR (CDCl3,
100 MHz): d = 149.9, 146.9, 143.7, 143.0, 140.4, 138.9, 135.4, 131.0, 128.9,
128.5, 127.6, 126.6, 126.3, 125.7, 123.9, 112.3, 40.5, 23.1, 21.3 ppm; HRMS
(ESI): m/z calcd for C34H3911BN: 472.3176; found: 472.3173 [M+H] + .

X-ray Crystal Structure Analysis of 3 b[35]

X-ray Crystal Structure Analysis of 1 b[35]

Single crystals of 3 b suitable for X-ray crystal analysis were obtained by


recrystallization from a MeOH/CH2Cl2 mixed solvent. Intensity data
were collected at 298 K on a Bruker SMART CCD X-ray diffractometer
(APEX II) with MoKa radiation (l = 0.71073 ) and graphite monochromator. A total of 13 875 reflections were measured at a maximum 2q
angle of 50.08, of which 5020 were independent reflections (Rint = 0.0275).
The structure was solved by direct methods (SHELXS-97)[36] and refined
by full-matrix least-squares cycles on F2 (SHELEXL-97).[36] All non-hydrogen atoms were refined anisotropically and all hydrogen atoms except
for those of the disordered solvent molecules were placed using AFIX instructions. The crystal data are as follows: C34H38BN; Mr = 471.46; crystal
size 0.20  0.15  0.10 mm3, monoclinic, C2/c, a = 23.585(3) , b =
11.6158(13) , c = 20.898(2) , V = 5700.5(11) 3, Z = 8, 1calcd =
1.099 g cm3. The refinement converged to R1 = 0.0468, wR2 = 0.1244 (I >
2s(I)), GOF = 1.044.

Single crystals of 1 b suitable for X-ray crystal analysis were obtained by


recrystallization from a MeOH/CH2Cl2 mixed solvent. Intensity data
were collected at 293 K on a Bruker SMART CCD X-ray diffractometer
(APEX II) with MoKa radiation (l = 0.71073 ) and graphite monochromator. A total of 13 054 reflections were measured at a maximum 2q
angle of 50.08, of which 9217 were independent reflections (Rint = 0.0511).
The structure was solved by direct methods (SHELXS-97)[36] and refined
by full-matrix least-squares cycles on F2 (SHELEXL-97).[36] All non-hydrogen atoms were refined anisotropically and all hydrogen atoms except
for those of the disordered solvent molecules were placed using AFIX instructions. The crystal data are as follows: C32H36BN; Mr = 445.43; crystal
size 0.20  0.15  0.10 mm3, triclinic, P1, a = 9.550(7) , b = 11.625(9) ,
c = 24.216(19) , V = 2672(4) 3, Z = 2, 1calcd = 1.107 g cm3. The refinement converged to R1 = 0.1146, wR2 = 0.2974 (I > 2s(I)), GOF = 0.998.

Chem. Asian J. 2013, 8, 3164 3176

Jian-Wu Wang, Cui-Hua Zhao et al.

3173

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

Acknowledgements

[8] a) Y. Liu, C. Deng, L. Tang, A. Qin, R. Hu, J. Z. Sun, B. Z. Tang, J.


Am. Chem. Soc. 2011, 133, 660 663; b) W. Z. Yuan, S. Chen,
J. W. Y. Lam, C. Deng, P. Lu, H. H. Y. Sung, I. D. Williams, H. S.
Kwok, Y. Zhang, B. Z. Tang, Chem. Commun. 2011, 47, 11216
11218; c) Z. Zhao, Z. Wang, P. Lu, C. Y. K. Chan, D. Liu, J. W. Y.
Lam, H. H. Y. Sung, I. D. Williams, Y. Ma, B. Z. Tang, Angew.
Chem. 2009, 121, 7744 7747; Angew. Chem. Int. Ed. 2009, 48, 7608
7611; d) Y. Dong, J. W. Y. Lam, A. Qin, Z. Li, J. Sun, H. H. Y. Sung,
I. D. Williams, B. Z. Tang, Chem. Commun. 2007, 40 42; e) H.
Tong, Y. Hong, Y. Dong, M. H uler, J. W. Y. Lam, Z. Li, Z. Guo,
Z. Guo, B. Z. Tang, Chem. Commun. 2006, 3705 3707; f) G. Yu, S.
Yin, Y. Liu, J. Chen, X. Xu, X. Sun, D. Ma, X. Zhan, Q. Peng, Z.
Shuai, B. Tang, D. Zhu, W. Fang, Y. Luo, J. Am. Chem. Soc. 2005,
127, 6335 6346; g) J. Luo, Z. Xie, J. W. Y. Lam, L. Cheng, H. Chen,
C. Qiu, H. S. Kwok, X. Zhan, Y. Liu, D. Zhu, B. Z. Tang, Chem.
Commun. 2001, 1740 1741; h) R. Hu, J. W. Y. Lam, Y. Liu, X.
Zhang, B. Z. Tang, Chem. Eur. J. 2013, 19, 5617 5624.
[9] a) Z. Zhang, B. Xu, J. Su, L. Shen, Y. Xie, H. Tian, Angew. Chem.
2011, 123, 11858 11861; Angew. Chem. Int. Ed. 2011, 50, 11654
11657; b) B. Wang, Y. Wang, J. Hua, Y. Jiang, J. Huang, S. Qian, H.
Tian, Chem. Eur. J. 2011, 17, 2647 2655; c) Z. Ning, Z. Chen, Q.
Zhang, Y. Yan, S. Qian, Y. Cao, H. Tian, Adv. Funct. Mater. 2007,
17, 3799 3807.
[10] a) S. Kim, Q. Zheng, G. S. He, D. J. Bharali, H. E. Pudavar, A. Baev,
P. N. Prasad, Adv. Funct. Mater. 2006, 16, 2317 2323; b) K. Itami, Y.
Ohashi, J. Yoshida, J. Org. Chem. 2005, 70, 2778 2792; c) B. K. An,
S. K. Kwon, S. D. Jung, S. Y. Park, J. Am. Chem. Soc. 2002, 124,
14410 14415; d) X. Gu, J. Yao, G. Zhang, C. Zhang, Y. Yan, Y.
Zhao, D. Zhang, Chem. Asian. J. 2013, DOI: 10.1002/asia.201300451.
[11] a) F. Wrthner, T. E. Kaiser, C. R. Saha-Mller, Angew. Chem. 2011,
123, 3436 3473; Angew. Chem. Int. Ed. 2011, 50, 3376 3410;
b) T. E. Kaiser, H. Wang, V. Stepanenko, F. Wrthner, Angew.
Chem. 2007, 119, 5637 5640; Angew. Chem. Int. Ed. 2007, 46, 5541
5544.
[12] a) T. P. I. Saragi, T. Spehr, A. Siebert, T. Fuhrmann-Lieker, J. Salbeck, Chem. Rev. 2007, 107, 1011 1065; b) C. Poriel, N. Cocherel, J.
Rault-Berthelot, L. Vignau, O. Jeannin, Chem. Eur. J. 2011, 17,
12631 12645; c) D. Thirion, J. Rault-Berthelot, L. Vignau, C. Poriel,
Org. Lett. 2011, 13, 4418 4421; d) J. Wang, W. Wan, H. Jiang, Y.
Gao, X. Jiang, H. Lin, W. Zhao, J. Hao, Org. Lett. 2010, 12, 3874
3877; e) C. Fan, Y. Chen, Z. Liu, Z. Jiang, C. Zhong, D. Ma, J. Qin,
C. Yang, J. Mater. Chem. C 2013, 1, 463 469; f) Z. Li, B. Jiao, Z.
Wu, P. Liu, L. Ma, X. Lei, D. Wang, G. Zhou, H. Hu, X. Hou, J.
Mater. Chem. C 2013, 1, 2183 2192; g) C. Poriel, J. Rault-Berthelot,
D. Thirion, F. Barri re, L. Vignau, Chem. Eur. J. 2011, 17, 14031
14046.
[13] a) C. H. Zhao, A. Wakamiya, Y. Inukai, S. Yamaguchi, J. Am.
Chem. Soc. 2006, 128, 15934 15935; b) A. Wakamiya, K. Mori, S.
Yamaguchi, Angew. Chem. 2007, 119, 4351 4354; Angew. Chem.
Int. Ed. 2007, 46, 4273 4276; c) C. H. Zhao, A. Wakamiya, S. Yamaguchi, Macromolecules 2007, 40, 3898 3900; d) C. H. Zhao, E.
Sakuda, A. Wakamiya, S. Yamaguchi, Chem. Eur. J. 2009, 15,
10603 10616.
[14] a) M. Shimizu, Y. Takeda, M. Higashi, T. Hiyama, Angew. Chem.
2009, 121, 3707 3710; Angew. Chem. Int. Ed. 2009, 48, 3653 3656;
b) M. Shimizu, R. Kaki, Y. Takeda, T. Hiyama, N. Nagai, H. Yamagishi, H. Furutani, Angew. Chem. 2012, 124, 4171 4175; Angew.
Chem. Int. Ed. 2012, 51, 4095 4099.
[15] a) C. H. Zhao, Y. H. Zhao, H. Pan, G. L. Fu, Chem. Commun. 2011,
47, 5518 5520; b) G. L. Fu, H. Y. Zhang, Y. Q. Yan, C. H. Zhao, J.
Org. Chem. 2012, 77, 1983 1990.
[16] For reviews for boron-based p-electron materials, see: a) C. R.
Wade, A. E. J. Broomsgrove, S. Aldridge, F. P. Gabbai, Chem. Rev.
2010, 110, 3958 3984; b) F. J kle, Chem. Rev. 2010, 110, 3985 4022;
c) T. W. Hudnall, C.-W. Chiu, F. P. Gabbai, Acc. Chem. Res. 2009,
42, 388 397; d) Z. M. Hudson, S. Wang, Acc. Chem. Res. 2009, 42,
1584 1596; e) M. Elbing, G. C. Bazan, Angew. Chem. 2008, 120,
846 850; Angew. Chem. Int. Ed. 2008, 47, 834 838; f) N. Matsumi,
Y. Chujo, Polym. J. 2008, 40, 77 89; g) S. Yamaguchi, A. Wakamiya,

We sincerely acknowledge the financial support from the National


Nature Science Foundation of China (grant nos. 21072117, 21272141),
Promotive Research Fund for Excellent Young and Middle-Aged Scientists of Shandong Province (no. BS 2012L021).

[1] a) K. Meerholz, C. D. Mller, O. Nuyken, in Organic Light Emitting


Devices. Synthesis Properties and Applications (Eds.: K. Mllen, U.
Scherf), Wiley-VCH, Weinheim, 2006; b) R. H. Friend, R. W.
Gymer, A. B. Holmes, J. H. Burroughes, R. N. Marks, C. Taliani,
D. D. C. Bradley, D. A. D. Santos, J. L. Brdas, M. Lgdlund, W. R.
Salaneck, Nature 1999, 397, 121 128.
[2] a) I. D. W. Samuel, G. A. Turnbull, Chem. Rev. 2007, 107, 1272
1295; b) F. Gao, Q. Liao, Z. Z. Xu, Y. H. Yue, Q. Wang, H. L.
Zhang, H. B. Fu, Angew. Chem. 2010, 122, 744 747; Angew. Chem.
Int. Ed. 2010, 49, 732 735; c) U. Scherf, S. Riechel, U. Lemmer,
R. F. Mahrt, Curr. Opin. Solid State Mater. Sci. 2001, 5, 143 154;
d) M. D. McGehee, A. J. Heeger, Adv. Mater. 2000, 12, 1655 1668;
e) G. Kranzelbinder, G. Leising, Rep. Prog. Phys. 2000, 63, 729 762;
f) N. Tessler, Adv. Mater. 1999, 11, 363 370; g) V. G. Kozlov, S. R.
Forrest, Curr. Opin. Solid State Mater. Sci. 1999, 4, 203 208;
h) E. A. Chandross, M. Berggren, R. E. Slusher, Science 1997, 277,
1787 1788.
[3] a) D. Yan, J. Lu, J. Ma, M. Wei, D. G. Evans, X. Duan, Angew.
Chem. 2011, 123, 746 749; Angew. Chem. Int. Ed. 2011, 50, 720
723; b) S. Sreejith, K. P. Divya, A. Ajayaghosh, Chem. Commun.
2008, 2903 2905; c) M. S. Meaney, V. L. McGuffin, Anal. Bioanal.
Chem. 2008, 391, 2557 2576; d) S. W. Thomas, III. , G. D. Joly,
T. M. Swager, Chem. Rev. 2007, 107, 1339 1386; e) L. Basabe-Desmonts, D. N. Reinhoudt, M. Crego-Calama, Chem. Soc. Rev. 2007,
36, 993 1017; f) T. J. Dale, J. Rebek, J. Am. Chem. Soc. 2006, 128,
4500 4501; g) O. S. Wolfbeis, J. Mater. Chem. 2005, 15, 2657 2669;
h) J. F. Callan, A. P. D. Silva, D. C. Magri, Tetrahedron 2005, 61,
8551 8588; i) S. W. Zhang, T. M. Swager, J. Am. Chem. Soc. 2003,
125, 3420 3421; j) R. Martnez-M ez, F. Sancen
n, Chem. Rev.
2003, 103, 4419 4476; k) S. Chen, Y. Hong, Y. Liu, J. Liu, C. W. T.
Leung, M. Li, R. T. K. Kwok, E. Zhao, J. W. Y. Lam, Y. Yu, B. Z.
Tang, J. Am. Chem. Soc. 2013, 135, 4926 4929.
[4] a) Y. Hong, J. W. Y. Lam, B. Z. Tang, Chem. Soc. Rev. 2011, 40,
5361 5388; b) M. Shimizu, T. Hiyama, Chem. Asian J. 2010, 5,
1516 1531; c) S. S. Babu, K. K. Kartha, A. Ajayaghosh, J. Phys.
Chem. Lett. 2010, 1, 3413 3424; d) S. P. Anthony, ChemPlusChem
2012, 77, 518 531.
[5] a) M. Grell, D. D. C. Bradley, G. Ungar, J. Hill, K. S. Whitehead,
Macromolecules 1999, 32, 5810 5817; b) U. Lemmer, S. Heun, R. F.
Mahrt, U. Scherf, M. Hopmeier, U. Siegner, E. O. Gbel, K. Mllen,
H. B ssler, Chem. Phys. Lett. 1995, 240, 373 378.
[6] a) T. Qin, G. Zhou, H. Scheiber, R. E. Bauer, M. Baumgarten, C. E.
Anson, E. J. W. List, K. Mllen, Angew. Chem. 2008, 120, 8416
8420; Angew. Chem. Int. Ed. 2008, 47, 8292 8296; b) J. Wang, Y.
Zhao, C. Dou, H. Sun, P. Xu, K. Ye, J. Zhang, S. Jiang, F. Li, Y.
Wang, J. Phys. Chem. B 2007, 111, 5082 5089; c) T. Sanji, T. Kanzawa, M. Tanaka, J. Organomet. Chem. 2007, 692, 5053 5059; d) H.
Langhals, O. Krotz, K. Polborn, P. Mayer, Angew. Chem. 2005, 117,
2479 2480; Angew. Chem. Int. Ed. 2005, 44, 2427 2428; e) T. Sato,
D. L. Jiang, T. Aida, J. Am. Chem. Soc. 1999, 121, 10658 10659;
f) T. Ozdemir, S. Atilgan, I. Kutuk, L. T. Yildirim, A. Tulek, M.
Bayindir, E. U. Akkaya, Org. Lett. 2009, 11, 2105 2107; g) H. Lu,
Q. Wang, L. Gai, Z. Li, Y. Deng, X. Xiao, G. Lai, Z. Shen, Chem.
Eur. J. 2012, 18, 7852 7861; h) G. L. Fu, H. Pan, Y. H. Zhao, C. H.
Zhao, Org. Biomol. Chem. 2011, 9, 8141 8146.
[7] a) Z. Xie, B. Yang, F. Li, G. Cheng, Li. Liu, G. Yang, H. Xu, L. Ye,
M. Hanif, S. Liu, D. Ma, Y. Ma, J. Am. Chem. Soc. 2005, 127,
14152 14153; b) F. He, H. Xu, B. Yang, Y. Duan, L. Tian, K.
Huang, Y. Ma, S. Liu, S. Feng, J. Shen, Adv. Mater. 2005, 17, 2710
2714.

Chem. Asian J. 2013, 8, 3164 3176

Jian-Wu Wang, Cui-Hua Zhao et al.

3174

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

[17]

[18]

[19]

[20]

[21]

[22]

Pure Appl. Chem. 2006, 78, 1413 1424; h) C. D. Entwistle, T. B.


Marder, Chem. Mater. 2004, 16, 4574 4585; i) C. D. Entwistle, T. B.
Marder, Angew. Chem. 2002, 114, 3051 3056; Angew. Chem. Int.
Ed. 2002, 41, 2927 2931; j) Y. Shirota, J. Mater. Chem. 2000, 10, 1
25; k) F. J kle, Coord. Chem. Rev. 2006, 250, 1107 1121.
a) M. E. Glogowski, J. L. R. Williams, J. Organomet. Chem. 1981,
218, 137 146; b) A. Schulz, W. Kaim, Chem. Ber. 1989, 122, 1863
1868.
a) Z. Yuan, N. J. Taylor, T. B. Marder, I. D. Williams, S. K. Kurtz, L.T. Cheng, J. Chem. Soc. Chem. Commun. 1990, 1489 1492; b) Z.
Yuan, N. J. Taylor, T. B. Marder, I. D. Williams, S. K. Kurtz, L.-T.
Cheng, Organic Materials for Non-linear Optics II (Eds.: R. A.
Hann, D. Bloor), RSC, Cambridge, 1991, pp. 190 196; c) Z. Yuan,
N. J. Taylor, Y. Sun, T. B. Marder, I. D. Williams, L.-T. Cheng, J. Organomet. Chem. 1993, 449, 27 37; d) Z. Yuan, N. J. Taylor, R. Ramachandran, T. B. Marder, Appl. Organomet. Chem. 1996, 10, 305
316; e) Z. Yuan, J. C. Collings, N. J. Taylor, T. B. Marder, C. Jardin,
J.-F. Halet, J. Solid State Chem. 2000, 154, 5 12; f) L. Porr s, M.
Charlot, C. D. Entwistle, A. Beeby, T. B. Marder, M. BlanchardDesce, Proc. SPIE-Int. Soc. Opt. Eng. 2005, 5934, 92 103; g) M.
Charlot, L. Porr s, C. D. Entwistle, A. Beeby, T. B. Marder, M. Blanchard-Desce, Phys. Chem. Chem. Phys. 2005, 7, 600 606; h) Z.
Yuan, C. D. Entwistle, J. C. Collings, D. Albesa-Jov, A. S. Batsanov,
J. A. K. Howard, H. M. Kaiser, D. E. Kaufmann, S.-Y. Poon, W.-Y.
Wong, C. Jardin, S. Fatallah, A. Boucekkine, J.-F. Halet, T. B.
Marder, Chem. Eur. J. 2006, 12, 2758 2771; i) J. C. Collings, S.-Y.
Poon, C. Le Droumaguet, M. Charlot, C. Katan, L.-O. Pl sson, A.
Beeby, J. A. Mosely, H. M. Kaiser, D. Kaufmann, W.-Y. Wong, M.
Blanchard-Desce, T. B. Marder, Chem. Eur. J. 2009, 15, 198 208;
j) C. D. Entwistle, J. C. Collings, A. Steffen, L.-O. P lsson, A. Beeby,
D. Albesa-Jov, J. M. Burke, A. S. Batsanov, J. A. K. Howard, J. A.
Mosely, S.-Y. Poon, W.-Y. Wong, F. Ibersiene, S. Fathallah, A. Boucekkine, J.-F. Halet, T. B. Marder, J. Mater. Chem. 2009, 19, 7532
7544; k) G.-J. Zhou, C.-L. Ho, W.-Y. Wong, Q. Wang, D.-G. Ma, L.X. Wang, Z.-Y. Lin, T. B. Marder, A. Beeby, Adv. Funct. Mater.
2008, 18, 499 511; l) L. Weber, D. Eickhoff, T. B. Marder, M. A.
Fox, P. J. Low, A. D. Dwyer, D. J. Tozer, S. Schwedler, A. Brockhinke, H. G. Stammler, B. Neumann, Chem. Eur. J. 2012, 18, 1369
1382.
a) M. Lequan, R. M. Lequan, K. Chance-Ching, M. Barzoukas, A.
Fort, H. Lahouche, G. Bravic, D. Chasseau, J. Gaultier, J. Mater.
Chem. 1992, 2, 719 725; b) M. Lequan, R. M. Lequan, K. ChaneChing, A.-C. Callier, M. Barzoukas, A. Fort, Adv. Mater. Opt. Electron. 1992, 1, 243 247; c) M. Lequan, R. M. Lequan, K. ChaneChing, J. Mater. Chem. 1991, 1, 997 999; d) C. Branger, M. Lequan,
R. M. Lequan, M. Barzoukas, A. Fort, J. Mater. Chem. 1996, 6, 555
558.
a) Z.-Q. Liu, Q. Fang, D. Wang, G. Xue, W.-T. Yu, Z.-S. Shao, M.-H.
Jiang, Chem. Commun. 2002, 2900 2901; b) Z.-Q. Liu, Q. Fang, D.
Wang, D.-X. Cao, G. Xue, W.-T. Yu, H. Lei, Chem. Eur. J. 2003, 9,
5074 5084; c) D.-X. Cao, Z.-Q. Liu, Q. Fang, G.-B. Xu, G. Xue, G.Q. Liu, W.-T. Yu, J. Organomet. Chem. 2004, 689, 2201 2206; d) Z.Q. Liu, Q. Fang, D.-X. Cao, D. Wang, G.-B. Xu, Org. Lett. 2004, 6,
2933 2936; e) Z.-Q. Liu, M. Shi, F.-Y. Li, Q. Fang, Z.-H. Chen, T.
Yi, C.-H. Huang, Org. Lett. 2005, 7, 5481 5484; f) D.-X. Cao, Z.-Q.
Liu, G.-Z. Li, G.-Q. Liu, G.-H. Zhang, J. Mol. Struct. 2008, 874, 46
50; g) M.-S. Yuan, Z.-Q. Liu, Q. Fang, J. Org. Chem. 2007, 72, 7915
7922; h) L. Ji, Q. Fang, M. S. Yuan, Z. Q. Liu, Y. X. Shen, H. F.
Chen, Org. Lett. 2010, 12, 5192 5195.
a) T. Noda, Y. Shirota, J. Am. Chem. Soc. 1998, 120, 9714 9715;
b) T. Noda, H. Ogawa, Y. Shirota, Adv. Mater. 1999, 11, 283 285;
c) T. Noda, Y. Shirota, J. Lumin. 2000, 87 89, 1168 1170; d) Y.
Shirota, M. Kinoshita, T. Noda, K. Okumuto, T. Ohara, J. Am.
Chem. Soc. 2000, 122, 11021 11022; e) M. Kinoshita, N. Fujii, T.
Tsukaki, Y. Shirota, Synth. Met. 2001, 121, 1571 1572; f) H. Doi, M.
Kinoshita, K. Okumoto, Y. Shirota, Chem. Mater. 2003, 15, 1080
1089.
a) W.-L. Jia, D.-R. Bai, T. McCormick, Q.-D. Liu, M. Motala, R.-Y.
Wang, C. Seward, Y. Tao, S. Wang, Chem. Eur. J. 2004, 10, 994

Chem. Asian J. 2013, 8, 3164 3176

[23]

[24]

[25]

[26]
[27]

[28]
[29]
[30]
[31]

[32]

[33]
[34]

[35]

3175

Jian-Wu Wang, Cui-Hua Zhao et al.

1006; b) W.-L. Jia, X. D. Feng, D.-R. Bai, Z. H. Lu, S. Wang, G.


Vamvounis, Chem. Mater. 2005, 17, 164 170; c) W.-L. Jia, M. J.
Moran, Y.-Y. Yuan, Z. H. Lu, S. Wang, J. Mater. Chem. 2005, 15,
3326 3333; d) S.-B. Zhao, T. McCormick, S. Wang, Inorg. Chem.
2007, 46, 10965 10967; e) D.-R. Bai, X.-Y. Liu, S. Wang, Chem. Eur.
J. 2007, 13, 5713 5723; f) X.-Y. Liu, D.-R. Bai, S. Wang, Angew.
Chem. 2006, 118, 5601 5604; Angew. Chem. Int. Ed. 2006, 45, 5475
5478; g) Z. M. Hudson, X.-Y. Liu, S. Wang, Org. Lett. 2011, 13, 300
303.
a) S. Yamaguchi, S. Akiyama, K. Tamao, J. Am. Chem. Soc. 2001,
123, 11372 11375; b) S. Yamaguchi, T. Shirasaka, S. Akiyama, K.
Tamao, J. Am. Chem. Soc. 2002, 124, 8816 8817; c) Y. Kubo, M. Yamamoto, M. Ikeda, M. Takeuchi, S. Shinkai, S. Yamaguchi, Angew.
Chem. 2003, 115, 2082 2086; Angew. Chem. Int. Ed. 2003, 42, 2036
2040.
a) S. Sol, F. P. Gabbai, Chem. Commun. 2004, 1284 1285; b) M.
Melaimi, F. P. Gabbai, J. Am. Chem. Soc. 2005, 127, 9680 9681;
c) C.-W. Chiu, F. P. Gabbai, J. Am. Chem. Soc. 2006, 128, 14248
14249; d) M. H. Lee, T. Agou, J. Kobayashi, T. Kawashima, F. P.
Gabbai, Chem. Commun. 2007, 1133 1135; e) T. W. Hudnall, M.
Melaimi, F. P. Gabbai, Org. Lett. 2006, 8, 2747 2749.
a) A. Sundararaman, M. Victor, R. Varughese, F. J kle, J. Am.
Chem. Soc. 2005, 127, 13748 13749; b) K. Parab, K. Venkatasubbaiah, F. Jkle, J. Am. Chem. Soc. 2006, 128, 12879 12885; c) H. Li,
K. Sundararaman, K. Venkatasubbaiah, F. J kle, J. Am. Chem. Soc.
2007, 129, 5792 5793.
Y.-H. Zhao, H. Pan, G.-L. Fu, J.-M. Lin, C.-H. Zhao, Tetrahedron
Lett. 2011, 52, 3832 3835.
a) M. Mazzeo, V. Vitale, F. Della Sala, M. Anni, G. Barbarella, L.
Favaretto, G. Sotgui, R. Cingolani, G. Gigli, Adv. Mater. 2005, 17,
34 39; b) Y. Liu, X. Xu, F. Zheng, Y. Cui, Angew. Chem. 2008, 120,
4614 4617; Angew. Chem. Int. Ed. 2008, 47, 4538 4541; c) J. C.
Doty, B. Babb, P. J. Grisdale, M. E. Glogowski, J. L. R. Williams, J.
Organomet. Chem. 1972, 38, 229 236; d) E. Sakuda, A. Funahashi,
N. Kitamura, Inorg. Chem. 2006, 45, 10670 10677.
H. Pan, G. L. Fu, Y. H. Zhao, C. H. Zhao, Org. Lett. 2011, 13, 4830
4833.
C. C. Chen, L. Y. Chin, S. C. Yang, M. J. Wu, Org. Lett. 2010, 12,
5652 5655.
A. Fukazawa, H. Yamada, S. Yamaguchi, Angew. Chem. 2008, 120,
5664 5667; Angew. Chem. Int. Ed. 2008, 47, 5582 5585.
M. C. Hong, Y. K. Kim, J. Y. Choi, S. Q. Yang, H. Rhee, Y. H. Ryu,
T. H. Choi, G. J. Cheon, G. I. An, H. Y. Kim, Y. Kim, D. J. Kim, J. S.
Lee, Y. T. Chang, K. C. Lee, Bioorg. Med. Chem. 2010, 18, 7724
7730.
a) J. R. Lakowicz, Principles of Fluorescence Spectroscopy, 2nd ed.,
Kluwer Academic Press/Plenum Publishers, New York, 1999,
pp. 187 194; b) E. Lippert, Z. Elektrochem. 1957, 61, 962 975;
c) N. Mataga, Y. Kaifu, M. Koizumi, Bull. Chem. Soc. Jpn. 1956, 29,
465 470.
L. Zhang, R. J. Clark, L. Zhu, Chem. Eur. J. 2008, 14, 2894 2903.
Gaussian 09, Revision A.02, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V.
Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X.
Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa,
M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven,
J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J.
Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J.
Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar,
J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox,
J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E.
Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W.
Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A.
Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O.
Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox, Gaussian,
Inc., Wallingford CT, 2009.
Crystal data for 1 b, 2 a, b, 3 a, b: see the Supporting Information.
CCDC 942707 (1 b), CCDC 942708 (2 a), CCDC 942709 (2 b),

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

[36] SHELX-97, Program for the Refinement of Crystal Structure, G. M.


Sheldrick, University of Gttingen, Gttingen, Germany, 1997.

CCDC 942710 (3 a) and CCDC 942711 (3 b) contain the supplementary crystallographic data for this paper. These data can be obtained
free of charge from The Cambridge Crysallographic Data Centre
via www.ccdc.cam.ac.uk/data_request/cif.

Chem. Asian J. 2013, 8, 3164 3176

Jian-Wu Wang, Cui-Hua Zhao et al.

Received: July 2, 2013


Published online: September 10, 2013

3176

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Você também pode gostar