Vous êtes sur la page 1sur 11

DOI: 10.1002/ajoc.

201500431

Full Paper

Carbazole Chromophores

Synthesis and Electronic, Optical, and Electrochemical Properties


of a Series of Tetracyanobutadiene-Substituted Carbazoles
Shin-ichiro Kato, Hiroto Noguchi, Satoshi Jin, and Yosuke Nakamura*[a]
Abstract: A series of tetracyanobutadiene (TCBD)-substituted carbazole chromophores were synthesized by formal
[2+
+2] cycloaddition between tetracyanoethylene (TCNE) and
carbazole derivatives containing an acetylenic moiety, followed by retro-electrocyclization. The reactivity of the carbazole varied depending on the conjugation connectivity between the carbazole and acetylenic moieties, the number of
the carbazole moieties, and the length of the acetylenic moi-

eties. All of the products feature intramolecular charge-transfer bands in the visible region. TCBD substitution at the 1position of the carbazole gave rise to small HOMOLUMO
gaps compared to that at the 2- and 3-positions. The relationship between the TCBD substitution pattern of the carbazole moiety and the effectiveness of donoracceptor pconjugation was evaluated, and it was established that the
p-conjugation was effective in order of positions 1 < 2 < 3.

Introduction

of nonplanar pushpull chromophores.[9] They feature intense


ICT interactions regardless of significant nonplanarity, and thus
some were found to have high third-order NLO properties[6b]
and function as DA dyes for solar cells.[6i, m]
Carbazole, which has fine optical and electronic properties
and high chemical stability has found wide application as
a part of functional materials.[10] One of the fascinating features
of carbazole is its strong electron-donating ability,[11] and
hence there have been a number of reports on carbazolebased DA hybrids for various applications,[1214] especially in
recent years, in DSSCs.[15] In this work, we focus our attention
on a class of carbazole-based pushpull chromophores,
namely 1,1,4,4-tetracyanobutadiene (TCBD)-substituted carbazoles 19 (Figure 1). We have recently synthesized various conjugated carbazole dimers (bicarbazoles),[11b] in which the carbazole moieties are connected either directly or via p-conjugated
spacers, such as acetylenic and olefinic spacers, at the 1-, 2-, or
3-position. We have disclosed their detailed structureproperty
relationships, particularly the remarkable effects the conjugation connectivity and the p-spacers have on the effectiveness
of p-conjugation and donor potency. We envisaged that the
CC bonds next to the electron-donating carbazole unit in
ethynylcarbazoles 1012 and bicarbazoles 1318 are sufficiently activated for the reaction with TCNE (Schemes 13). Although Michinobu and co-workers reported the reaction of
alkyne-linked poly(3,6-carbazole)s with TCNE to yield DA-type
conjugated polymers,[16] the reactivity of well-defined carbazole
derivatives containing an acetylenic moiety with TCNE has remained unexplored. Herein, we describe the synthesis of 19
and the characterization of their electronic, optical, and electrochemical properties using electronic absorption spectroscopy, cyclic voltammetry (CV), and quantum chemical calculations. The properties of 19 are discussed in terms of the effects of the TCBD-substitution pattern of the carbazole moiety,

The rational design of novel donoracceptor (DA) compounds, so-called pushpull chromophores,[1] featuring intense
intramolecular charge-transfer (ICT) interactions, has received
an impressive amount of attention because of their potential
application in nonlinear optics (NLOs),[2] organic photovoltaics
(OPVs),[3] dye-sensitized solar cells (DSSCs),[4] and ambipolar organic field-effect transistors (OFETs).[5] Variation of the donor
and acceptor moieties has been widely recognized as a useful
approach for tuning the HOMOLUMO gaps in pushpull systems. From the geometrical point of view, a large number of
fully planar pushpull chromophores have been extensively investigated over the past decades, for which efficient electronic
communication between the donor and acceptor is expected.
Nonplanar molecules have several advantagesthey are generally more soluble, less aggregating, and more readily sublimable than planar ones; these features favor the formation of
the thin films that are used in optoelectronic devices. Diederichs group and others have recently shown that alkynes activated appropriately by electron-donating groups undergo
formal [2+
+2] cycloadditions with electron-withdrawing olefins,[68] such as tetracyanoethylene (TCNE), 7,7,8,8-tetracyanop-quinodimethane (TCNQ), and related powerful electron acceptors, followed by retro-electrocyclization, to afford a class
[a] Dr. S.-i. Kato, H. Noguchi, S. Jin, Prof. Dr. Y. Nakamura
Division of Molecular Science
Faculty of Science and Technology
Gunma University
1-5-1 Tenjin-cho, Kiryu, Gunma 376-8515 (Japan)
Fax: (+ 81) 277-30-1314
E-mail: nakamura@gunma-u.ac.jp
Supporting information for this article is available on the WWW under
http://dx.doi.org/10.1002/ajoc.201500431.
Asian J. Org. Chem. 2016, 5, 246 256

246

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper

Figure 1. Tetracyanobutadiene-substituted carbazoles 19.

the number of the carbazole moieties, and the number of the


ethynylene units between the carbazole and TCBD moieties.
The systematic study of the physicochemical properties of the
series 19 leads to an understanding of DA conjugation in
nonplanar TCBD-carbazole chromophores.

Results and Discussion


Synthesis
We examined the formal [2+
+2] cycloaddition/retro-electrocyclization reactions between 1018 and TCNE and consequently
obtained the corresponding TCBD-substituted carbazoles 19
(Schemes 13). Notably, 1018 differ in their reactivity with
TCNE depending on the conjugation connectivity between the
carbazole and acetylenic moieties, the number of the carbazole
units, and the length of the acetylenic moieties, as described
below.
The reaction of 10 with 1 equivalent of TCNE readily proceeded at room temperature in 1,2-dichloroethane to give 1 in
a quantitative yield (Scheme 1). Similarly, 3 was isolated in
90 % yield by the reaction of 12 with 1 equivalent of TCNE at
room temperature. The reactions of 10 and 12 with TCNE were
complete in 2 h and 15 min, respectively, on the basis of analysis by thin-layer chromatography (TLC). Compound 2 was also
obtained in 69 % yield by the reaction of 11 with TCNE at
room temperature, however, 11 was not entirely consumed
even after 20 h. The reaction in 1,2-dichloroethane at reflux
Asian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

Scheme 1. Synthesis of compounds 13.

(83 8C) led to complete consumption of 11, and 2 was obtained in 85 % yield.
In contrast to 1012, the reaction of bicarbazoles 1318
with TCNE did not proceed at room temperature. Heating a solution of 1318 and TCNE was required for this transformation
to occur; 1,1,2,2-tetrachloroethane, with a boiling point of
147 8C, was used as a reaction solvent (Schemes 2 and 3).
Moreover, an excess amount of TCNE was used for the transformation of bicarbazoles except for 18; we added 1 equivalent of TCNE at a time up to 10 equivalents, while monitoring
247

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper

Scheme 2. Synthesis of compounds 46.

Scheme 3. Synthesis of compounds 79.

the reactions by TLC. The reaction of 13 with TCNE (5 equiv)


proceeded in 1,1,2,2-tetrachloroethane at reflux to afford 4 in
28 % yield (Scheme 2). The analogous reaction of 14 with TCNE
(3 equiv) provided 5 in 43 % yield. Bicarbazoles 13 and 14
were almost entirely consumed, however, unidentified, highly
polar byproducts were formed together with the desired 4 and
5. The reaction of 15 with TCNE (2 equiv) at 100 8C afforded 6
in 80 % yield. The observed lower reactivity toward TCNE of
1315 compared to the corresponding 1012 should reflect
the greater steric crowding around the ethynylene unit due to
the presence of the second carbazole moiety in the former.[17]
Increasing the length of the acetylenic moiety in 16 and 18
relative to 13 and 15, respectively, dramatically increased the
yields of the formal cycloaddition/retro-electrocyclization reactions (Scheme 3); the transformation of 16 and 18 to 7 and 9,
respectively, occurred quantitatively. These remarkable yields
for 7 and 9 are attributed to the decrease of steric crowding in
16 and 18 relative to 13 and 15, respectively, which results
from the increase in the acetylenic unit.[18] Conversely, the reaction of 17 with TCNE (10 equiv) afforded 8 in only 20 % yield
(Scheme 3), which is lower than the yield for 5.[19] Bis-TCNE addition to the butadiynylene moiety was not observed in any of
these cases.[20]
The investigation of the reactions of 1018 with TCNE described above clearly indicates that the reactivity of the compounds with the acetylenic functionality at the 2-position of
carbazole is lower than that at the 1- and 3-positions in any of
the series 1012, 1315, and 1618. The acetylenic moieties in
11, 14, and 17 are located at the meta positions relative to the
electron-rich nitrogen atom of the carbazole moiety, while the
acetylenic moieties of 10, 13, and 16 are located at the ortho
positions, and those of 12, 15, and 18 are at the para posiAsian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

tions. Thus, due to cross-conjugation, the electron density of


the acetylenic moieties at the 2-position of carbazole is lower
than that at the 1- and 3-positions, which should be responsible for the observed lower reactivity of 11, 14, and 17 compared to 10, 12, 13, 15, 16, and 18. The lower reactivity of 13
and 16 compared to 15 and 18, respectively, could be attributable to the close proximity of the Et groups to the acetylenic
moieties in 13 and 16; the Et groups in 13 and 16 kinetically
decrease their reactivity.[21] Compounds 19 were fully characterized by various spectroscopic methods and elemental analysis.[22] They were stable in the solid state at room temperature
under air. Thermogravimetric analyses of 19 under a nitrogen
atmosphere revealed that the temperatures (T95), at which 5 %
weight loss occurred upon heating, ranged from 130 8C (8) to
294 8C (4; Table S1 in the Supporting Information).
Electronic Absorption Spectroscopy
We measured the electronic absorption spectra of TCBD-substituted carbazoles 19 in CH2Cl2 at room temperature. The absorption spectra of 13, 46, and 79 are shown in Figure 2 a,
b, and c, respectively, and the data are summarized in Table 1.
Compounds 19 feature significantly broad absorption bands
around 400900 nm, and the absorption onset of some compounds reaches into the near-infrared region. Their absorption
bands are identified as the ICT bands characteristic of TCBDsubstituted derivatives.[6] The colors of CH2Cl2 solutions of 1, 4,
and 7 are blue, those of 2, 5, and 8 are brown, and those of 3,
6, and 9 are purple or purple-red (Figure S3).[23] We further
measured the electronic absorption spectra of 13 in toluene,
248

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper
conjugation between the carbazole donor and TCBD acceptor
moieties in 1 is less effective than that in 2 and 3 (see below).
The longest lmax values of 46 are blue-shifted relative to
those of 13, respectively, by 3487 nm (4: 650 nm, 5: 412 nm,
6: 455 nm; Figure 2 b). The increase of the optical HOMO
LUMO gaps (DEopt) observed by electronic absorption spectroscopy results from an increase of the LUMO levels relative to
the HOMO levels, as confirmed by CV (see below). Incorporation of the second carbazole chromophore increases the
e values in the ICT bands. The e values in the ICT absorption
band of 4 are significantly low compared to 5 and 6, indicating
that the DA conjugation in 4 is ineffective for the same steric
reason as in the case of 1.
The substantial redshifts of the longest lmax in 8 and 9 as
compared to 5 and 6, respectively, by approximately 40 nm,
are attributed to the extension of p-conjugation (8: 451 nm, 9:
491 nm; Figure 2 c). In sharp contrast, the longest lmax of 7 is
blue-shifted relative to that of 4 by 7 nm (7: 643 nm), suggesting that the large twist between the carbazole and TCBD moieties is present in 7 as in 1 and 4 (see below).
In any of the series 13, 46, and 79, TCBD substitution at
the 1-position of the carbazole leads to the most-redshifted
absorption spectra, and TCBD at the 2-position gives rise to
the least. Thus there is a clear relationship between the TCBDsubstitution pattern of the carbazole moiety and the DEopt
values; the optical gaps generally decrease in order of position
2 > 3 > 1, although we cannot entirely rule out the effect of
the tBu groups in 1, 4, and 7. Consequently, the DEopt value of
1.81 eV for 1 is the lowest in the compounds in this study.
Electrochemistry
We performed CV for 19 as well as 9-ethyl-9H-carbazole (19)
and 3,6-di-tert-butyl-9-ethyl-9H-carbazole (20) in CH2Cl2 containing nBu4NPF6 (0.1 mol L1) as the supporting electrolyte.
Compounds 19 are amphoteric within the available potential
window (Figure 3), and the oxidation potentials (Epa) and reduction potentials (Epc) versus ferrocenium/ferrocene (Fc + /Fc)
are summarized in Table 2.
All TCBD-substituted carbazoles showed oxidation steps centered on the electron-donating carbazole moiety. The oxidation
processes for 1, 4, and 7 bearing the tBu groups at the 3,6-positions of the carbazole moiety are reversible or quasi-reversible, while those for 2, 3, 5, 6, 8, and 9 are irreversible. The tBu
groups at the 3- and 6-positions in 1, 4, and 7 sufficiently increase the kinetic stability of cationic species formed upon
electron release.[10b, 11b,c] The somewhat lower reversibility in 7
than 1 and 4 is attributed to the electron-accepting ethynylene
unit in 7. Irreversible oxidation processes for 2, 3, 5, 6, 8, and 9
indicate the formation of an unstable radical cation, which can
be involved in electrochemical dimerization or polymerization.
Compounds 19 has two reversible, well-resolved reduction
steps characteristic of TCBD-substituted derivatives; each dicyanovinyl moiety is responsible for a 1 e reduction.
As summarized in Table 2, compounds 19 are oxidized at
higher potentials than the corresponding 19 or 20, attributable
to p-conjugation between the carbazole donor and TCBD ac-

Figure 2. Electronic absorption spectra of a) 13, b) 46, and c) 79 in


CH2Cl2 at room temperature.

THF, and DMF to investigate the effect of solvent polarity on


their spectroscopic behavior, and observed negative solvatochromism (Figure S4). These results clearly indicate the occurrence of charge-transfer in the ground state.
As shown in Figure 2 a, the longest absorption maximum
(lmax) of 1 is remarkably redshifted as compared to 2 and 3 by
over 100 nm (1: 684 nm, 2: 480 nm, 3: 542 nm), and the molar
absorption coefficients (e) of 1 in the longer wavelength
region are somewhat small compared to those of 2 and 3.
These characteristic spectral features of 1 indicate that the
steric clash between the Et group and the TCBD moiety in 1 induces the large twist about the CC single bond connecting
the carbazole moiety to the TCBD moiety, and thereby the pAsian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

249

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper
Table 1. Electronic absorption spectral data and calculated lowest excitation energies of 19.
Compd

lmax [nm (eV)][a,b]

e [L mol1 cm1][a]

Calcd lmax [nm (eV)][c]

f[c]

Composition of band[c]

684 (1.81)

800

2
3

480 (2.53)
542 (2.29)

5300
5300

650 (1.90)

2500

412 (3.00)

27 100

745 (1.66)
616 (2.01)
574 (2.16)
649 (1.90)
571 (2.16)
683 (1.81)
650 (1.90)
411 (3.01)

0.036
0.032
0.141
0.088
0.026
0.074
0.065
0.078

455 (2.72)

34 300

406 (3.04)
426 (2.90)

0.087
0.045

416 (2.97)

0.064

610 (2.03)
592 (2.09)
466 (2.65)

0.020
0.216
0.267

450 (2.74)

0.204

481 (2.57)
459 (2.69)

0.083
0.406

H-1!L, 3 %; H!L, 96 %
H-1!L, 96 %; H!L, 3 %
H-1!L, 98 %
H!L, 98 %
H-1!L, 97 %
H!L, 98 %
H-1!L, 97 %;
H-3!L, 3 %; H-3!L + 1, 10 %;
H-2!L, 81 %; H-2!L + 1, 2 %
H-3!L, 21 %; H-3!L + 1, 8 %; H-2!L, 2 %; H-2!L + 1, 65 %
H-1!L, 6 %; H-1!L + 1, 5 %;
H!L, 85 %
H-1!L, 21 %; H-1!L + 1, 12 %;
H!L + 1, 60 %
H-1!L, 93 %; H!L, 4 %
H-1!L, 4 %; H!L, 93 %
H-3!L, 29 %; H-3!L + 1, 2 %;
H-2!L, 66 %
H-3!L, 65 %; H-2!L, 27 %;
H-2!L + 1, 2 %
H-1!L, 62 %; H!L, 34 %
H-1!L, 35 %; H!L, 59 %

643 (1.92)

7900

451 (2.74)

26 300

491 (2.52)

31 900

[a] In CH2Cl2. [b] Only the longest absorption maxima are shown. [c] TD-DFT [TD-PBE0/6-31G(d)] calculations were performed with the use of optimized
structures at the PBE0/6-31G(d) level of theory, and only energies with f > 0.02 are shown; f = oscillator strength; H = HOMO, L = LUMO.

fords a low HOMO level and a high LUMO level.[25] Judging


from the first Epc values for 19, the effectiveness of p-conjugation between the carbazole donor and the TCBD acceptor is
likely to follow the established trend of the TCBD-substitution
pattern of the carbazole moiety, and thus the p-conjugation
becomes effective in order of positions 1 < 2 < 3; accordingly,
the electrochemical HOMOLUMO gaps (DEredox) decrease in
order of positions 3 > 2 > 1.[26] The least effective DA p-conjugation resulting from TCBD substitution at the 1-position of
the carbazole is readily interpreted in terms of the steric reasons described above. TCBD substitution at the 2-position of
carbazole results in the cross-conjugation pathway between
the electron-rich nitrogen atom of the carbazole and the TCBD
moiety, whereas TCBD at the 3-position brings about the linear
conjugation. This alternative conjugation pathway should be
the main reason for the more efficient DA conjugation by
TCBD substitution at the 3-position relative to the 2-position. It
is worth stating that the effect of the TCBD-substitution pattern on the DEredox values is inconsistent with that for the DEopt
values (see above), which was addressed by using the quantum chemical calculations described below.

ceptor moieties in 19. Interestingly, the first Epc values for 19


were more affected by the structural variations compared to
the Epa values; the first Epc values range from 0.96 to 0.56 V,
whereas the Epa values range from + 0.97 to + 1.10 V.[24] Increasing number of the carbazole donor moieties disfavors the
reduction. Thus, 46 exclusively have negative shifts of the
first Epc values compared to the corresponding 13, and the
pronounced shifts of 0.23 and 0.29 V for 5 and 6, respectively,
were observed (2: 0.62 V, 3: 0.67 V, 5: 0.85 V, 6: 0.96 V);
the first Epc value for 4 negatively shifts by only 0.04 V (1:
0.56 V, 4: 0.60 V). These results demonstrate that the electron-accepting ability of the TCBD moiety decreases due to the
presence of the second carbazole moiety. The first Epc values of
8 and 9 were positively shifted relative those to 5 of 0.15 V
and to 6 of 0.16 V (8: 0.70 V, 9: 0.80 V), respectively, due to
the electron-accepting ability of the ethynylene unit. Conversely, 7 has almost the same first Epc value (0.61 V) as that of 4.
Overall, the change of the first Epc values from 1 and 4 to 7 is
less pronounced than that from 2 and 5 to 8 and 3 and 6 to 9.
This finding, together with the electronic absorption spectroscopic results strongly support that the carbazole donor
moiety and the TCBD acceptor moiety in 1, 4, and 7 are only
weakly conjugated due to the distinct twist around the CC
bond that connects the carbazole moiety to the TCBD moiety.
Intriguingly, in any of the series 13, 46, and 79, TCBD
substitution at the 1-position of the carbazole moiety results in
the most positively shifted reduction potentials, whereas TCBD
at the 3-position affords the least. It is well documented that
in a DA-conjugated system, an electron-donating group restricts the electron reduction and an electron-withdrawing
group makes oxidation difficult. Thus, within constitutional isomers of pushpull chromophores, efficient DA conjugation afAsian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

Quantum Chemical Calculations


To estimate the molecular orbitals of 19, we performed calculations using density functional theory (DFT) at the PBE0/6311G(d,p)//PBE0/6-31G(d) level.[2729] The results are summarized in Table 2, and the HOMOs and LUMOs of 19 are shown
in Figure 4. The geometrical optimization supports that the
twist angles (q) of the CC single bond connecting the carbazole moiety to the TCBD moiety in 1, 4, and 7 are larger than
those of the corresponding angles in 2, 3, 5, 6, 8, and 9; for in250

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper
Table 2. Oxidation (Epa) and reduction (Epc) potentials of 19 determined
by CV in CH2Cl2 (0.1 mol L1 nBu4NPF6),[a] theoretically calculated HOMO
and LUMO levels and their gaps (DEHOMOLUMO),[b] and electrochemical
(DEredox)[c] and optical energy gaps (DEopt)[d] .

Compd

Epa
[V]

+ 0.97[e] 0.56[e]
1.32[e]
+ 1.03[f] 0.62[e]
1.27[e]
[g]
+ 1.07
0.67[e]
1.30[e]
+ 1.06[e] 0.60[e]
1.29[e]
+ 1.04[f] 0.85[e]
1.23[e]
+ 1.10[f] 0.96[e]
1.29[e]
+ 0.98[g] 0.61[e]
1.17[e]
[f]
+ 1.02
0.70[e]
1.12[e]
+ 1.05[f] 0.80[e]
1.17[e]
+ 0.89[f] [h]
+ 0.71[e] [h]

2
3
4
5
6
7
8
9
19
20

Epc
[V]

HOMO LUMO DEHOMOLUMO DEredox DEopt


[eV]
[eV]
[eV]
[V]
[eV]
6.40

4.06

2.34

1.53

1.81

6.49

4.16

2.33

1.65

2.53

6.54

4.08

2.46

1.74

2.29

6.21

3.75

2.46

1.66

1.90

6.35

3.04

3.31

1.89

3.00

6.35

2.80

3.55

2.06

2.72

6.12

3.56

2.56

1.59

1.92

6.40

3.46

2.94

1.72

2.74

6.34

3.23

3.11

1.85

2.52

5.76
5.53

0.81
0.77

4.95
4.76

[h]
[h]

[a] All potentials are given versus the Fc + /Fc couple used as external
standard. Scan rate: 100 mV s1. [b] PBE0/6-311G(d,p)//PBE0/6-31G(d).
[c] Electrochemical gap, DEredox, is defined as the potential difference between Epa and Epc. [d] Optical gap, DEopt, is defined as the energy corresponding to the lmax. [e] Reversible wave. [f] Irreversible wave. [g] Quasireversible. [h] No reduction peak within the available potential window.

pattern of the carbazole moiety, the number of the carbazole


units, and the ethynylene unit inserted between the carbazole
and TCBD moieties on the HOMO and LUMO levels and the
gaps; the results obtained from the calculations are generally
in qualitative agreement with those obtained by CV. The calculated HOMOLUMO gaps (DEHOMOLUMO) are apparently dependent on the TCBD-substitution pattern of the carbazole moiety,
and the DEHOMOLUMO values decrease in order of positions 3 >
2 > 1 for almost all the compounds; exceptionally, the DEHOMO
LUMO value of 1 is estimated to be almost the same as that of 2
(1: 2.34 eV, 2: 2.33 eV). Thus, the DA p-conjugation between
the carbazole and TCBD moieties becomes effective in order of
positions 1 < 2 < 3. Upon increasing the number of the carbazole donor moieties from one in 13 to two in the corresponding 49, the LUMO levels essentially elevate relative to the
HOMO levels, and thereby the DEHOMOLUMO values become
large. The ethynylene unit apparently lowers the LUMO levels
sufficiently. Accordingly, the DEHOMOLUMO values of 8 and 9 are
smaller than those of 5 and 6, respectively, although the
DEHOMOLUMO value of 7 is slightly larger than that of 4.
To gain further insight into the electronic transitions, we
conducted time-dependent (TD) DFT calculations (TD/PBE0/631G(d)//PBE0/6-31G(d)) for 19, and the results are summar-

Figure 3. Cyclic voltammograms of a) 13, b) 46, and c) 79 measured in


CH2Cl2 (0.1 mol L1 nBu4NPF6) at a scan rate 100 mV s1.

stance, the q values for 1, 2, and 3 were determined to be 558,


428, and 428, respectively. All of the HOMOs are almost localized in the electron-donating carbazole moiety as expected.
The HOMOs of 1, 3, 4, 6, 7, and 9 also comprise a small but
distinct contribution from the TCBD moiety, whereas those of
2, 5, and 8 have almost no contribution. This difference likely
stems from the fact that the 2-position of the carbazole moiety
in 2, 5, and 8 lies in the nodal plane of the HOMOs. By contrast, all the LUMOs are largely localized in the electron-withdrawing TCBD moiety and have a slight but recognizable contribution from the carbazole moiety.
The calculations for 19 reveal the following three important
findings with respect to the effects of the TCBD-substitution
Asian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

251

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper

Figure 4. Molecular orbital plots [PBE0/6-311G(d,p)//PBE0/6-31G(d)] of a) 1, b) 2, c) 3, d) 4, e) 5, f) 6, g) 7, h) 8, and i) 9. The lower plots represent the HOMOs,
and the upper plots represent LUMOs. The optimized structures are shown in Figure S5.

ized in Table 1. The absorption maxima in the low-energy regions of 1, 3, 4, 6, 7, and 9 are essentially ascribed to transitions from the HOMOs and/or HOMO-1s, mainly localized on
the carbazole moiety, to the LUMOs, mainly localized on the
TCBD moiety, judging from their moderate-to-high oscillator
strengths f, which are 0.0200.406. Compounds 2, 5, and 8
have significantly low f values of 0.0012, 0.0091, and 0.0159, respectively, for the HOMOLUMO transitions. This suggests that
the HOMOLUMO transitions for 2, 5, and 8 are almost forbidden, unlike those of 1, 3, 4, 6, 7, and 9. The absorption
maxima in the low-energy region of 2, 5, and 8 are ascribed to
the transitions from the HOMO-1s, HOMO-2s, and/or HOMO-3s,
mostly localized on the carbazole moiety, to the LUMOs (Figures S6S8). The calculated lmax values of 2, 5, and 8 are blueshifted relative to those of 1 and 3, 4 and 6, and 7 and 9, reAsian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

spectively (Table 1), which is in good agreement with the electronic absorption spectroscopy results. The forbidden HOMO
LUMO transitions in 2, 5, and 8 are explained by the almost no
overlap between the HOMOs and LUMOs in the TCBD moiety,
which should lead to a lack of the transition dipole moment
from the carbazole moiety to the TCBD moiety. In other words,
the small but distinctive overlap between the HOMOs and
LUMOs in both the carbazole and TCBD moieties in 1, 3, 4, 6,
7, and 9 are essential for their allowed HOMOLUMO transitions. Therefore, it can be concluded that the difference in
trend of the TCBD-substitution pattern for the optical gaps
and that for the electrochemical gaps is attributed to the forbidden HOMOLUMO transitions that result from the TCBD
substituent at the 2-position of the carbazole.
252

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper
Conclusions

the reference electrode. All electrochemical measurements were


performed in CH2Cl2 solution (approx. 5 104 mol L1) containing
nBu4NPF6 (0.1 mol L1) at RT. All potentials are referenced to the
Fc + /Fc couple, used as a standard.

We have synthesized a series of TCBD-substituted carbazoles


19 by means of the formal [2+
+2] cycloaddition/retro-electrocyclization reaction between TCNE and carbazole derivatives
1018 containing an acetylenic moiety. Compounds 1018
differ in their reactivity toward TCNE depending on the conjugation connectivity between the carbazole and acetylenic moieties, the number of the carbazole moieties, and the length of
the acetylenic moieties. The electronic structure of 19 was
studied by using electronic absorption spectroscopy, CV, and
DFT calculations, and consequently a picture of the structure
property relationships emerged. They feature ICT bands in the
UV/vis region, and the absorption of some compounds reaches
into the near-infrared region. TCBD substitution at the 1-position of the carbazole leads to the redshifted longest lmax
values, namely the small optical HOMOLUMO gaps as compared to compounds bearing TCBD at the 2- and 3-positions,
however, the former gives rise to the lower e values in the ICT
bands. The optical gaps increase by incorporation of the
second carbazole moiety and decrease by insertion of the
ethynylene unit between the carbazole donor and TCBD acceptor moieties. Noticeably, the relationship between the
TCBD-substitution pattern of the carbazole moiety and the effectiveness of DA conjugation is observed in 19 from the
CV, whereby the conjugation becomes effective in the order of
position 1 < 2 < 3. The experimentally obtained structureproperty relationships were qualitatively corroborated by the DFT
calculations. We believe that the present study provides valuable information for the design and synthesis of carbazolebased pushpull chromophores.

Quantum chemical calculations were performed using the Gaussian 09 program package.[27] Molecular properties in the electronic
ground state were computed using the hybrid PBE0 DFT
method.[30] Geometry optimization of 19 was performed with no
symmetry constraints with the 6-31G(d) basis set.[31] Frequency calculations on these optimized geometries revealed no imaginary
frequencies. Further single-point calculations were performed with
the 6-311G(d,p) basis set[32] to obtain molecular orbital energies
with use of optimized structures at the PBE0/6-31G(d) level of
theory. TD calculations for 19 were performed at the PBE0/631G(d) level of theory with use of optimized structures at the
PBE0/6-31G(d) level of theory. The calculated electronic transitions
are vertical.

[2+
+2] Cycloaddition/Retro-Electrocyclization Reactions
General procedure: One equivalent of TCNE was added to a solution of acetylenic carbazole derivative (1 equiv) in 1,2-dichloroethane or 1,1,2,2-tetrachloroethane, and the resulting mixture was
stirred at 25147 8C under an argon atmosphere. To fully react the
precursor carbazole, 1 equivalent of TCNE was added at a time to
the mixture up to 10 equivalents, while being monitored by TLC.
After the mixture was stirred for 15 min48 h, the solvent was removed under reduced pressure. The residue was purified by washing with hexane, recrystallization, and/or column chromatography.
An analytically pure material was obtained by using recycling GPC,
eluting with CHCl3.
2-(3,6-Di-tert-butyl-9-ethyl-9H-carbazol-1-yl)buta-1,3-diene1,1,4,4-tetracarbonitrile (1): Compound 10 (40 mg, 0.12 mmol)
was allowed to react with TCNE (15 mg, 0.12 mmol) in 1,2-dichloroethane (5 mL) at room temperature for 2 h. The crude material was
dissolved in hexane, and insoluble material was removed by filtration. The filtrate was evaporated under reduced pressure to give
1 as a blue solid (55 mg, quant); m.p. 190193 8C; 1H NMR
(300 MHz, CDCl3): d = 1.25 (t, J = 7.2 Hz, 3 H), 1.45 (s, 9 H), 1.47 (s,
9 H), 3.914.02 (m, 1 H), 4.174.29 (m, 1 H), 7.18 (d, J = 1.8 Hz, 1 H),
7.36 (d, J = 8.7 Hz, 1 H), 7.60 (dd, J = 1.8, 8.7 Hz, 1 H), 8.11 (d, J =
1.8 Hz, 1 H), 8.13 (s, 1 H), 8.37 ppm (d, J = 1.8 Hz, 1 H); 13C NMR
(125 MHz, CDCl3): d = 14.1, 31.8, 32.0, 34.9, 35.0, 40.5, 93.2, 98.2,
107.7, 109.3, 110.8, 110.9, 111.0, 112.2, 116.9, 122.7, 123.0, 124.9,
125.3, 126.8, 135.7, 139.7, 143.6, 144.0, 149.5, 162.2 ppm; UV/vis/
NIR (CH2Cl2): lmax (relative intensity) = 275 (sh, 0.65), 301 (1.00), 360
(sh, 0.18), 684 nm (0.02); HRMS (FAB): m/z: calcd for C30H29N5 :
459.2423 [M] + ; found 459.2422. Elemental analysis calcd (%) for
C30H29N50.05 CHCl3 : C 77.53, H 6.29, N 15.04; found: C 77.68, H
6.55, N 14.82.

Experimental Section
General
Column chromatography and plug filtrations were carried out with
SiO2 60. TLC was conducted on aluminum sheets coated with SiO2
60 F254 ; visualization was performed with a lamp (254 or 365 nm).
Melting points (m.p.) are uncorrected. Residual and deuterated solvent signals in the 1H and 13C NMR spectra were used as internal
references, respectively (CDCl3, 1H: d = 7.26 ppm, 13C: d =
77.16 ppm; 1,1,2,2-[D2]tetrachloroethane, 1H: d = 6.00 ppm; 13C: d =
74.00 ppm). Chemical shifts (d) are given as values in ppm. The
coupling constants (J) are given in Hz. The apparent resonance
multiplicity is described as s (singlet), d (doublet), t (triplet), q
(quartet), or m (multiplet). FAB-MS spectra were recorded in the
positive-ion mode with m-nitrobenzyl alcohol as the matrix.
MALDI-TOF MS spectra were recorded in the positive-ion mode
with dithranol as the matrix. The most important signals are reported in m/z units with M as the molecular ion. The UV/vis spectra
were measured in a 1 cm cuvette at 298 K. The absorption maxima
(lmax) are reported in nm with relative intensities. Shoulders are indicated as sh. Gel permeation chromatography (GPC) was performed with CHCl3 ; an analytically pure material was obtained by
recycling GPC. Compounds 1018 were prepared according to literature procedures.[11b]
CV was performed by using a cell equipped with a platinum working electrode, a platinum wire counter electrode, and Ag/AgNO3 as
Asian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

2-(9-Ethyl-9H-carbazol-2-yl)buta-1,3-diene-1,1,4,4-tetracarbonitrile (2): Compound 11 (30 mg, 0.13 mmol) was allowed to react
with TCNE (17 mg, 0.13 mmol) in 1,2-dichloroethane (10 mL) at
reflux (84 8C) for 3 h. The crude material was purified by column
chromatography (SiO2, CH2Cl2) and recrystallization from CH2Cl2/
hexane to give 2 as a purple solid (40 mg, 85 %); m.p. 97100 8C;
1
H NMR (500 MHz, CDCl3): d = 1.48 (t, J = 7.2 Hz, 3 H), 4.43 (q, J =
7.2 Hz, 2 H), 7.18 (dd, J = 1.8, 8.1 Hz, 1 H), 7.307.34 (m, 1 H), 7.48 (d,
J = 8.1 Hz, 1 H), 7.587.62 (m, 2 H), 8.13 (s, 1 H), 8.16 (d, J = 7.8 Hz,
1 H), 8.27 ppm (d, J = 8.1 Hz, 1 H); 13C NMR (125 MHz, CDCl3): d =
14.1, 38.1, 90.1, 98.3, 108.5, 109.3, 109.6, 111.4, 111.9, 112.1, 119.4,
120.3, 121.9, 121.9, 122.1, 126.5, 128.1, 128.5, 139.6, 141.9, 153.4,
163.1 ppm; UV/vis/NIR (CH2Cl2): lmax (relative intensity) = 267 (1.00),

253

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper
304 (0.79), 360 (sh, 0.15), 480 nm (0.16); HRMS (FAB): m/z: calcd for
C22H13N5 : 347.1171 [M] + ; found 347.1176. Elemental analysis calcd
(%) for C22H13N50.23 CHCl3 : C 71.23, H 3.56, N 18.68; found: C
71.28, H 3.89, N 18.44.

123.6, 124.3, 127.8, 127.9, 140.9, 143.5, 168.2 ppm; UV/vis/NIR


(CH2Cl2): lmax (relative intensity) = 289 (1.00), 333 (0.63), 387 (0.40),
455 nm (0.80); MS (MALDI-TOF): m/z: 541.12 [M+
+H] + ; HRMS (FAB):
+
m/z: calcd for C36H24N6 : 540.2062 [M] ; found: 540.2063. Elemental
analysis calcd (%) for C36H24N60.13 CHCl3 : C 78.03, H 4.37, N 15.11;
found: C 77.90, H 4.75, N 14.75.

2-(9-Ethyl-9H-carbazol-3-yl)buta-1,3-diene-1,1,4,4-tetracarbonitrile (3): Compound 12 (170 mg, 0.77 mmol) was allowed to react
with TCNE (102 mg, 0.80 mmol) in 1,2-dichloroethane (8 mL) at
room temperature for 15 min. The crude material was purified by
recrystallization from CH2Cl2/hexane to give 3 as a purple solid
(241 mg, 90 %); m.p. 233236 8C; 1H NMR (500 MHz, CDCl3): d = 1.50
(t, J = 7.2 Hz, 3 H), 4.43 (q, J = 7.2 Hz, 2 H), 7.357.38 (m, 1 H), 7.49
(d, J = 8.1 Hz, 1 H), 7.557.60 (m, 3 H), 8.15 (d, J = 8.1 Hz, 1 H), 8.16
(s, 1 H), 8.24 ppm (d, J = 1.8 Hz, 1 H); 13C NMR (125 MHz, CDCl3): d =
14.0, 38.3, 86.4, 98.3, 108.7, 109.7, 110.1, 111.8, 112.0, 112.7, 120.9,
121.2, 121.3, 122.6, 122.8, 124.4, 126.9, 127.8, 141.0, 143.3, 154.9,
162.3 ppm; UV/vis/NIR (CH2Cl2): lmax (relative intensity) = 288 (1.00),
330 (sh, 0.50), 350 (sh, 0.29), 450 (sh, 0.12), 542 nm (0.16); HRMS
(FAB): m/z: calcd for C22H13N5 : 347.1171 [M] + ; found: 347.1143. Elemental analysis calcd (%) for C22H13N50.03 CHCl3 : C 75.39, H 3.74, N
19.96; found: C 75.37, H 3.85, N 19.65.

2-(3,6-Di-tert-butyl-9-ethyl-9H-carbazol-1-yl)-3-[2-(3,6-di-tertbutyl-9-ethyl-9H-carbazol-1-yl)ethynyl]buta-1,3-diene-1,1,4,4-tetracarbonitrile (7): Compound 16 (20 mg, 0.030 mmol) was allowed


to react with TCNE (20 mg, 0.15 mmol) in 1,1,2,2-tetrachloroethane
(5 mL) at reflux (147 8C) for 5 h. The crude material was purified by
column chromatography (SiO2, CH2Cl2/hexane 6:4) to give 7 as
a blue solid (23 mg, quant); m.p. 155157 8C; 1H NMR (300 MHz,
CDCl3): d = 1.38 (t, J = 7.2 Hz, 3 H), 1.42 (s, 9 H), 1.47 (s, 9 H), 1.48 (s,
9 H), 1.51 (s, 9 H), 1.53 (br s, 3 H), 4.35 (br s, 2 H), 4.73 (q, J = 7.2 Hz,
2 H), 7.37 (d, J = 8.5 Hz, 1 H), 7.39 (d, J = 8.5 Hz, 1 H), 7.53 (d, J =
1.8 Hz, 1 H), 7.607.63 (m, 3 H), 8.10 (d, J = 1.8 Hz, 1 H), 8.14 (d, J =
1.8 Hz, 1 H), 8.32 (d, J = 1.8 Hz, 1 H), 8.42 ppm (d, J = 1.8 Hz, 1 H);
13
C NMR (125 MHz, CDCl3): d = 14.0, 14.7, 29.8, 31.5, 31.8, 32.0, 32.1,
34.7, 34.9, 35.0, 39.1, 41.2, 88.8, 91.3, 91.6, 99.6, 108.9, 110.0, 110.2,
111.7, 112.4, 113.1, 115.0, 116.5, 116.8, 120.6, 122.2, 123.1, 123.8,
123.9, 125.1, 125.3, 125.5, 126.9, 128.3, 131.9, 136.6, 138.1, 139.0,
139.9, 142.5, 142.8, 143.8, 144.3, 145.9, 164.0 ppm; UV/vis/NIR
(CH2Cl2): lmax (relative intensity) = 289 (1.00), 315 (sh, 0.49), 397
(0.47), 643 nm (0.26); MS (MALDI-TOF): m/z: 788.22 [M] + ; HRMS
(FAB): m/z: calcd for C54H56N6 : 788.4566 [M] + ; found: 788.4567. Elemental analysis calcd (%) for C54H56N60.13 CHCl3 : C 80.80, H 7.03, N
10.45; found: C 80.75, H 7.33, N 10.11.

Bis(3,6-di-tert-butyl-9-ethyl-9H-carbazol-1-yl)buta-1,3-diene1,1,4,4-tetracarbonitrile (4): Compound 13 (30 mg, 0.047 mmol)


was allowed to react with TCNE (20 mg, 0.23 mmol) in 1,1,2,2-tetrachloroethane (8 mL) at reflux (147 8C) for 9 h. The crude material
was purified by column chromatography (SiO2, CH2Cl2/hexane 6:4)
to give 4 (10 mg, 28 %) as a blue solid; m.p. 178181 8C; 1H NMR
(600 MHz, 1,1,2,2-[D2]tetrachloroethane, 120 8C): d = 1.18 (brs, 36 H),
4.17 (brs, 4 H), 7.10 (brs, 2 H), 7.31 (d, J = 6.7 Hz, 2 H), 7.58 (d, J =
6.7 Hz, 2 H), 8.02 ppm (brs, 4 H); 13C NMR (150 MHz, 1,1,2,2[D2]tetrachloroethane, 120 8C): d = 13.6, 31.3, 31.9, 34.4, 34.6, 40.7,
109.7, 111.4, 116.4, 120.4, 121.2, 123.1, 125.0, 135.9, 139.8, 142.7,
144.1 ppm (17 signals out of 21 expected); UV/vis/NIR (CH2Cl2): lmax
(relative intensity) = 270 (sh, 1.00), 290 (sh, 0.83), 315 (sh, 0.56), 376
(sh, 0.26), 650 nm (0.07); MS (MALDI-TOF): m/z: 764.25 [M] + ; Elemental analysis calcd (%) for C52H56N60.11 CHCl3 : C 80.43, H 7.27, N
10.80; found: C 80.75, H 7.56, N 10.44.

2-(9-Ethyl-9H-carbazol-2-yl)-3-[2-(9-ethyl-9H-carbazol-2-yl)ethynyl]buta-1,3-diene-1,1,4,4-tetracarbonitrile (8): Compound 17


(60 mg, 0.137 mmol) was allowed to react with TCNE (180 mg,
1.38 mmol) in 1,1,2,2-tetrachloroethane (15 mL) at reflux (147 8C)
for 48 h. The crude material was purified by column chromatography (SiO2, CH2Cl2/hexane 8:2) to give 8 as a brown solid (15 mg,
20 %); m.p. 130 8C (decomposed); 1H NMR (300 MHz, CDCl3): d =
1.46 (t, J = 7.2 Hz, 3 H), 1.49 (t, J = 7.2 Hz, 3 H), 4.38 (q, J = 7.2 Hz,
2 H), 4.47 (q, J = 7.2 Hz, 2 H), 7.267.34 (m, 2 H), 7.447.51 (m, 4 H),
7.547.60 (m, 1 H), 7.597.64 (m, 1 H), 7.71 (d, J = 0.6 Hz, 1 H), 8.05
(d, J = 1.8 Hz, 1 H), 8.12 (d, J = 7.8 Hz, 2 H), 8.15 (d, J = 7.8 Hz, 1 H),
8.24 ppm (d, J = 8.1 Hz, 1 H); 13C NMR (125 MHz, CDCl3): d = 14.1,
14.2, 37.9, 38.2, 84.6, 86.9, 92.9, 109.2, 109.4, 110.2, 110.6, 111.7,
111.9, 113.0, 114.2, 115.2, 120.1, 120.2, 120.4, 121.1, 121.6, 121.8,
121.9, 121.9, 122.1, 122.4, 124.4, 127.1, 127.4, 128.2, 128.6, 128.8,
139.2, 139.4, 141.7, 142.3, 149.3, 164.9 ppm; UV/vis/NIR (CH2Cl2):
lmax (relative intensity) = 261 (1.00), 311 (0.44), 335 (sh, 0.40), 358
(0.39), 451 nm (0.55); MS (MALDI-TOF): m/z: 565.12 [M+
+H] + ; HRMS
+
(FAB): m/z: calcd for C38H24N6 : 564.2062 [M] ; found: 564.2062. Elemental analysis calcd (%) for C38H24N60.10 CHCl3 : C 79.37, H 4.21, N
14.58; found: C 79.13, H 4.55, N 14.51.

Bis(9-ethyl-9H-carbazol-2-yl)buta-1,3-diene-1,1,4,4-tetracarbonitrile (5): Compound 14 (42 mg, 0.102 mmol) was allowed to react
with TCNE (36 mg, 0.282 mmol) in 1,1,2,2-tetrachloroethane
(10 mL) at reflux (147 8C) for 40 h. The crude material was purified
by column chromatography (SiO2, CH2Cl2) to give 5 as a purple
solid (24 mg, 43 %); m.p. 181183 8C; 1H NMR (300 MHz, CDCl3): d =
1.48 (t, J = 7.2 Hz, 6 H), 4.42 (q, J = 7.2 Hz, 4 H), 7.277.32 (m, 2 H),
7.42 (dd, J = 1.8, 8.4 Hz, 2 H), 7.47 (d, J = 8.4 Hz, 2 H), 7.577.63 (m,
2 H), 8.12 (d, J = 8.4 Hz, 2 H), 8.13 (s, 2 H), 8.20 ppm (d, J = 8.4 Hz,
2 H); 13C NMR (125 MHz, CDCl3): d = 14.1, 38.1, 84.8, 109.4, 110.1,
111.9, 113.1, 120.4, 120.5, 121.8, 121.8, 121.9, 128.1, 128.6, 128.8,
139.4, 142.4, 168.8 ppm; UV/vis/NIR (CH2Cl2): lmax (relative intensity) = 262 (1.00), 306 (0.44), 412 nm (0.63); MS (MALDI-TOF): m/z:
541.18 [M+
+H] + ; HRMS (FAB): m/z: calcd for C36H24N6 : 540.2062 [M] +
; found: 540.2063. Elemental analysis calcd (%) for
C36H24N60.03 CHCl3 : C 79.52, H 4.45, N 15.44; found: C 79.73, H
4.79, N 15.51.

2-(9-Ethyl-9H-carbazol-3-yl)-3-[2-(9-ethyl-9H-carbazol-3-yl)ethynyl]buta-1,3-diene-1,1,4,4-tetracarbonitrile (9): Compound 18


(20 mg, 0.045 mmol) was allowed to react with TCNE (6 mg,
0.046 mmol) in 1,1,2,2-tetrachloroethane (5 mL) at 60 8C for 17 h.
The crude material was purified by column chromatography (SiO2,
CH2Cl2) to give 9 as a purple solid (25 mg, quant); m.p. 137 8C (decomposed); 1H NMR (500 MHz, CDCl3): d = 1.46 (t, J = 7.2 Hz, 3 H),
1.49 (t, J = 7.2 Hz, 3 H), 4.37 (q, J = 7.2 Hz, 2 H), 4.40 (q, J = 7.2 Hz,
2 H), 7.297.60 (m, 8 H), 7.73 (dd, J = 1.5, 8.4 Hz, 1 H), 8.00 (dd, J =
1.8, 8.7 Hz, 1 H), 8.08 (d, J = 7.8 Hz, 1 H), 8.19 (d, J = 7.8 Hz, 1 H), 8.39
(d, J = 1.5 Hz, 1 H), 8.61 ppm (d, J = 1.5 Hz, 1 H); 13C NMR (125 MHz,
CDCl3): d = 14.0, 14.1, 38.1, 38.3, 80.8, 88.1, 90.9, 109.0, 109.4, 109.5,
109.6, 109.8, 111.0, 112.4, 112.5, 113.6, 120.9, 121.1, 121.3, 121.4,
121.5, 122.4, 122.7, 123.4, 123.8, 124.2, 124.3, 127.3, 127.6, 127.6,

Bis(9-ethyl-9H-carbazol-3-yl)buta-1,3-diene-1,1,4,4-tetracarbonitrile (6): Compound 15 (40 mg, 0.096 mmol) was allowed to react
with TCNE (24 mg, 0.192 mmol) in 1,1,2,2-tetrachloroethane
(10 mL) at 100 8C for 6 h. The crude material was purified by
column chromatography (SiO2, CH2Cl2) to give 6 as a purple solid
(42 mg, 80 %); m.p. 223226 8C; 1H NMR (300 MHz, CDCl3): d = 1.49
(t, J = 7.2 Hz, 6 H), 4.40 (q, J = 7.2 Hz, 4 H), 7.307.36 (m, 2 H), 7.46
7.59 (m, 6 H), 8.03 (dd, J = 2.1, 8.7 Hz, 2 H), 8.11 (d, J = 7.8 Hz, 2 H),
8.64 ppm (d, J = 1.8 Hz, 2 H); 13C NMR (125 MHz, CDCl3): d = 14.0,
38.3, 80.9, 109.6, 109.9, 112.6, 113.9, 121.4, 121.4, 122.6, 122.7,
Asian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

254

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper
127.8, 131.6, 140.6, 140.9, 142.4, 143.5, 150.0, 164.2 ppm; UV/vis/
NIR (CH2Cl2): lmax (relative intensity) = 289 (1.00), 331 (0.63), 345 (sh,
0.57), 389 (0.46), 491 nm (0.85); MS (MALDI-TOF): m/z: 565.18
[M+
+H] + ; HRMS (FAB): m/z: calcd for C38H24N6 : 564.2062 [M] + ;
found:
564.2062.
Elemental
analysis
calcd
(%)
for
C38H24N60.05 CHCl3 : C 80.09, H 4.25, N 14.73; found: C 80.37, H
4.43, N 14.35.

Acknowledgements
This work was supported by a Grant-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science
and Technology of Japan (No. 24550040, 15K05416). This work
was performed under the Cooperative Research Program of
Network Joint Research Center for Materials and Devices
(Kyushu University: no. 2015474). We thank Prof. Takayuki
Miyazaki (Gunma University) for generous permission to use
the UV/vis/NIR spectrometer. Quantum chemical calculations
were performed in the Research Center for Computational Science, Japan.

[7]

[8]
[9]

[10]

Keywords: carbazoles conjugation donoracceptor


systems electrochemistry tetracyanoethylene
[1] a) R. Gompper, H. U. Wagner, Angew. Chem. Int. Ed. Engl. 1988, 27, 1437;
Angew. Chem. 1988, 100, 1492; b) H. Meier, Angew. Chem. Int. Ed. 2005,
44, 2482; Angew. Chem. 2005, 117, 2536; c) L. Beverina, G. A. Pagani,
Acc. Chem. Res. 2014, 47, 319.
[2] a) D. R. Kanis, M. A. Ratner, T. J. Marks, Chem. Rev. 1994, 94, 195; b) J.-L.
Brdas, C. Adant, P. Tackx, A. Persoons, B. M. Pierce, Chem. Rev. 1994, 94,
243; c) J. L. Segura, N. Martn, Angew. Chem. Int. Ed. 2001, 40, 1372;
Angew. Chem. 2001, 113, 1416; d) M. Bendikov, F. Wudl, D. F. Perepichka,
Chem. Rev. 2004, 104, 4891; e) S. Barlow, S. R. Marder in Functional Organic Materials (Eds.: T. J. J. Mller, U. H. F. Bunz), Wiley-VCH, Weinheim,
2007, pp. 393.
[3] a) A special issue on organic photovoltaics: Acc. Chem. Res. 2009, 42,
1689 (Eds.: J.-L. Brdas, J. R. Durrant); b) J. L. Delgado, P.-A. Bouit, S. Filippone, M. . Herranz, N. Martn, Chem. Commun. 2010, 46, 4853;
c) P. M. Beaujuge, J. M. J. Frchet, J. Am. Chem. Soc. 2011, 133, 20009;
d) F. Wrthner, K. Meerholz, Chem. Eur. J. 2010, 16, 9366; e) A. Mishra, P.
Buerle, Angew. Chem. Int. Ed. 2012, 51, 2020; Angew. Chem. 2012, 124,
2060; f) J. Roncali, P. Leriche, P. Blanchard, Adv. Mater. 2014, 26, 3821;
g) V. Malytskyi, J.-J. Simon, L. Patrone, J.-M. Raimundo, RSC Adv. 2015, 5,
354.
[4] a) K. Hara, Z.-S. Wang, T. Sato, A. Furube, R. Katoh, H. Sugihara, Y. DanOh, C. Kasada, A. Shinpo, S. Suga, J. Phys. Chem. B 2005, 109, 15476;
b) S. Ito, H. Miura, S. Uchida, M. Takata, K. Sumioka, P. Liska, P. Comte, P.
Pchy, M. Grtzel, Chem. Commun. 2008, 5194; c) S. Erten-Ela, M. D.
Yilmaz, B. Icli, Y. Dede, S. Icli, E. U. Akkaya, Org. Lett. 2008, 10, 3299;
d) G. Zhang, H. Bala, Y. Cheng, D. Shi, X. Lv, Q. Yu, P. Wang, Chem.
Commun. 2009, 2198; e) S. Wenger, P.-A. Bouit, Q. Chen, J. Teuscher, D.
Di Censo, R. Humphry-Baker, J.-E. Moser, J. L. Delgado, N. Martn, S. M.
Zakeeruddin, M. Grtzel, J. Am. Chem. Soc. 2010, 132, 5164.
[5] a) J. Zaumseil, H. Sirringhaus, Chem. Rev. 2007, 107, 1296; b) J. Cornil, J.L. Brdas, J. Zaumseil, H. Sirringhaus, Adv. Mater. 2007, 19, 1791; c) M.
Mas-Torrent, C. Rovira, Chem. Soc. Rev. 2008, 37, 827; d) L. Huang, M.
Stolte, H. Brckstmmer, F. Wrthner, Adv. Mater. 2012, 24, 5750; e) A.
Liess, L. Huang, A. Arjona-Esteban, A. Lv, M. Gsnger, V. Stepanenko, M.
Stolte, F. Wrthner, Adv. Funct. Mater. 2015, 25, 44.
[6] a) Y. Morioka, N. Yoshizawa, J. Nishida, Y. Yamashita, Chem. Lett. 2004,
33, 1190; b) T. Michinobu, J. C. May, J. H. Lim, C. Boudon, J.-P. Gisselbrecht, P. Seiler, M. Gross, I. Biaggio, F. Diederich, Chem. Commun. 2005,
737; c) T. Michinobu, C. Boudon, J.-P. Gisselbrecht, P. Seiler, B. Frank,
N. N. P. Moonen, M. Gross, F. Diederich, Chem. Eur. J. 2006, 12, 1889;
d) M. Kivala, C. Boudon, J.-P. Gisselbrecht, P. Seiler, M. Gross, F. Diederich,
Asian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

[11]

[12]

[13]

[14]

[15]

[16]

255

Angew. Chem. Int. Ed. 2007, 46, 6357; Angew. Chem. 2007, 119, 6473;
e) T. Shoji, S. Ito, K. Toyota, M. Yasunami, N. Morita, Chem. Eur. J. 2008,
14, 8398; f) T. Shoji, J. Higashi, S. Ito, T. Okujima, M. Yasunami, N. Morita,
Chem. Eur. J. 2011, 17, 5116; g) T. M. Pappenfus, D. K. Schneiderman, J.
Casado, J. T. L. Navarrete, C. R. Delgado, G. Zotti, B. Vercelli, M. D. Lovander, L. M. Hinkle, J. N. Bohnsack, K. R. Mann, Chem. Mater. 2011, 23,
823; h) A. Lelige, P. Blanchard, T. Rousseau, J. Roncali, Org. Lett. 2011,
13, 3098; i) S. Niu, G. Ulrich, P. Retailleau, R. Ziessel, Org. Lett. 2011, 13,
4996; j) R. Garca, M. . Herranz, M. R. Torres, P.-A. Bouit, J. L. Delgado, J.
Calbo, P. M. Viruela, E. Ort, N. Martn, J. Org. Chem. 2012, 77, 10707;
k) Y.-L. Wu, M. C. Stuparu, C. Boudon, J.-P. Gisselbrecht, W. B. Schweizer,
K. K. Baldridge, J. S. Siegel, F. Diederich, J. Org. Chem. 2012, 77, 11014;
l) T. Michinobu, N. Satoh, J. Cai, Y. Li, L. Han, J. Mater. Chem. C 2014, 2,
3367.
Donoracceptor-type aromatic polymers were also synthesized: a) T.
Michinobu, J. Am. Chem. Soc. 2008, 130, 14074; see also a review: b) T.
Michinobu, Chem. Soc. Rev. 2011, 40, 2306.
For a review on nonplanar push pull chromophores, see: S.-i. Kato, F.
Diederich, Chem. Commun. 2010, 46, 1994.
Thermal [2+
+2] cycloaddition of TCNE with electron-rich ruthenium
acetylide complexes, followed by ring opening of the initially formed
cyclobutenes to form organometallic TCBD derivatives, was first reported by Bruce and co-workers: M. I. Bruce, J. R. Rodgers, M. R. Snow, A. G.
Swincer, J. Chem. Soc. Chem. Commun. 1981, 271.
For reviews, see: a) J. V. Grazulevicius, P. Strohriegl, J. Pielichowski, K.
Pielichowski, Prog. Polym. Sci. 2003, 28, 1297; b) J.-F. Morin, M. Leclerc,
D. Ads, A. Siove, Macromol. Rapid Commun. 2005, 26, 761; c) N. Blouin,
M. Leclerc, Acc. Chem. Res. 2008, 41, 1110; d) P.-L. T. Boudreault, J.-F.
Morin, M. Leclerc in Design and Synthesis of Conjugated Polymers (Eds.:
M. Leclerc, J.-F. Morin), Wiley-VCH, Weinheim, 2010, pp. 205.
a) S.-i. Kato, S. Shimizu, H. Taguchi, A. Kobayashi, S. Tobita, Y. Nakamura,
J. Org. Chem. 2012, 77, 3222; b) S.-i. Kato, H. Noguchi, A. Kobayashi, T.
Yoshihara, S. Tobita, Y. Nakamura, J. Org. Chem. 2012, 77, 9120; c) S.-i.
Kato, S. Shimizu, A. Kobayashi, T. Yoshihara, S. Tobita, Y. Nakamura, J.
Org. Chem. 2014, 79, 618; d) S.-i. Kato, Y. Yamada, H. Hiyoshi, K. Umezu,
Y. Nakamura, J. Org. Chem. 2015, 80, 9076.
For photorefractive materials, see: a) Y. Zhang, T. Wada, L. Wang, H.
Sasabe, Chem. Mater. 1997, 9, 2798; see also a review: b) Y. Zhang, T.
Wada, H. Sasabe, J. Mater. Chem. 1998, 8, 809.
For selected examples of NLO materials, see: a) J. L. Daz, A. Dobarro, B.
Villacampa, D. Velasco, Chem. Mater. 2001, 13, 2528; b) S. Lee, K. R. J.
Thomas, S. Thayumanavan, C. J. Bardeen, J. Phys. Chem. A 2005, 109,
9767; c) E. De Meulenaere, W.-Q. Chen, S. Van Cleuvenbergen, M.-L.
Zheng, S. Psilodimitrakopoulos, R. Paesen, J.-M. Taymans, M. Ameloot, J.
Vanderleyden, P. Loza-Alvarez, X.-M. Duan, K. Clays, Chem. Sci. 2012, 3,
984.
For selected examples of fluorophores, see: a) X. Sun, Y. Liu, X. Xu, C.
Yang, G. Yu, S. Chen, Z. Zhao, W. Qiu, Y. Li, D. Zhu, J. Phys. Chem. B
2005, 109, 10786; b) C.-H. Zhao, A. Wakamiya, Y. Inukai, S. Yamaguchi, J.
Am. Chem. Soc. 2006, 128, 15934; c) A. P. Kulkarni, X. Kong, S. A. Jenekhe, Adv. Funct. Mater. 2006, 16, 1057; d) R. M. Adhikari, D. C. Neckers,
B. K. Shah, J. Org. Chem. 2009, 74, 3341; e) P. Kotchapradist, N. Prachumrak, R. Tarsang, S. Jungsuttiwong, T. Keawin, T. Sudyoadsuk, V. Promarak,
J. Mater. Chem. C 2013, 1, 4916.
For selected examples of dye-sensitized solar cells, see: a) N. Koumura,
Z.-S. Wang, S. Mori, M. Miyashita, E. Suzuki, K. Hara, J. Am. Chem. Soc.
2006, 128, 14256; b) Z.-S. Wang, N. Koumura, Y. Cui, M. Takahashi, H. Sekiguchi, A. Mori, T. Kubo, A. Furube, K. Hara, Chem. Mater. 2008, 20,
3993; c) N. Koumura, Z.-S. Wang, M. Miyashita, Y. Uemura, H. Sekiguchi,
Y. Cui, A. Mori, S. Mori, K. Hara, J. Mater. Chem. 2009, 19, 4829; d) Y.
Ooyama, Y. Shimada, S. Inoue, T. Nagano, Y. Fujikawa, K. Komaguchi, I.
Imae, Y. Harima, New J. Chem. 2011, 35, 111; e) W. Lee, N. Cho, J. Kwon,
J. Ko, J.-I. Hong, Chem. Asian J. 2012, 7, 343; f) C. Chen, J.-Y. Liao, Z. Chi,
B. Xu, X. Zhang, D.-B. Kuang, Y. Zhang, S. Liu, J. Xu, J. Mater. Chem.
2012, 22, 8994; g) P. Thongkasee, A. Thangthong, N. Janthasing, T. Sudyoadsuk, S. Namuangruk, T. Keawin, S. Jungsuttiwong, V. Promarak,
ACS Appl. Mater. Interfaces 2014, 6, 8212; see also a review: h) N. Koumura, K. Hara, Heterocycles 2013, 87, 275.
T. Michinobu, H. Fujita, Materials 2010, 3, 4773.
2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Paper
[17] Increasing the number of the electron-donating carbazole units from
one in 1012 to two in 1315, is considered to be electronically favorable for the reactions with TCNE.
[18] Compounds 16 and 18 should be electronically less favorable for the
cycloaddition reaction with TCNE than 13 and 15, respectively, reflecting the more electron-withdrawing butadiynylene unit than the ethynylene unit: J.-P. Gisselbrecht, N. N. P. Moonen, C. Boudon, M. B. Nielsen, F.
Diederich, M. Gross, Eur. J. Org. Chem. 2004, 2959.
[19] Compound 17 was not entirely consumed on the basis of a TLC analysis of the reaction mixture.
[20] Double [2+
+2] cycloaddition/retro-electrocyclization reactions between
TCNE and aniline-capped buta-1,3-diynes were reported: B. Breiten, Y.-L.
Wu, P. Jarowski, J.-P. Gisselbrecht, C. Boudon, M. Griesser, C. Onitsch, G.
Gescheidt, W. B. Schweizer, N. Langer, C. Lennartz, F. Diederich, Chem.
Sci. 2011, 2, 88.
[21] The close proximity of the Et groups to the ethynylene moiety in 13 in
the solid state was revealed by X-ray diffraction analysis. See Ref. [11b].
[22] Compound 4 gave broad and complex 1H NMR signals in 1,1,2,2[D2]tetrachloroethane at room temperature. Upon heating to 120 8C,
well-resolved peaks appeared and could be clearly assigned to the proposed structure (Figure S15). This indicates that the rotation of the CC
single bond connecting the carbazole moiety to the TCBD moiety is restricted at room temperature, which should be attributable to the
severe steric hindrance between the Et groups and the dicyanovinyl
moieties.
[23] Compounds 19 are nonfluorescent.
[24] Similar significant effects of the structural variations on the Epc values
compared to the Epa values were also observed in other tetracyanobutadiene-donor systems. For details, see Ref. [6c].
[25] N. N. P. Moonen, F. Diederich, Org. Biomol. Chem. 2004, 2, 2263.
[26] We note that the conformational difference of the TCBD moiety in 19
might also affect their Epc values.
[27] All calculations were performed using Gaussian 09, Revision D.01, 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.

Asian J. Org. Chem. 2016, 5, 246 256

www.AsianJOC.org

[28]

[29]

[30]

[31]
[32]

Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, 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, 2013.
We note that the optimized structures are strongly affected by the initial structures in general; we did not perform the conformation search
for 19. We note that 2,3-diarylated buta-1,3-dienes often adopt an scis conformation with a torsion angle of approximately 65 8; see
refs. [6l] and [29]. Based on this finding, the conformations of 49 are
considered to be s-cis conformation in the TCBD unit, although the optimized structures of 49 with an s-trans conformation can be also obtained. Judging from the reported X-ray crystal structures of 2-arylated
buta-1,3-dienes, such as TCBD-N,N-dimethylaniline conjugates, which
adopt an s-trans conformation, 2 and 3 are considered to have an strans conformation in the TCBD unit; see Ref. [6b]. However, the conformation of 1 should be s-cis-type. 1 cannot adopt an s-trans conformation due to the significant steric clash between the Et group and the dicyanovinyl moieties. The optimized structures of 19 are shown in Figure S5.
a) C. A. van Walree, B. C. van der Wiel, L. W. Jenneskens, M. Lutz, A. L.
Spek, R. W. A. Havenith, J. H. van Lenthe, Eur. J. Org. Chem. 2007, 4746;
b) P. A. Limacher, H. P. Luethi, J. Phys. Chem. A 2008, 112, 2913.
a) J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 1996, 77, 3865;
b) J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 1997, 78, 1396;
c) C. Adamo, V. Barone, J. Chem. Phys. 1999, 110, 6158.
J. S. Binkley, J. A. Pople, W. J. Hehre, J. Am. Chem. Soc. 1980, 102, 939.
a) G. A. Petersson, M. A. Al-Laham, J. Chem. Phys. 1991, 94, 6081; b) G. A.
Petersson, A. Bennett, T. G. Tensfeldt, M. A. Al-Laham, W. A. Shirley, J.
Mangzaris, J. Chem. Phys. 1988, 89, 2193.

Manuscript received: October 20, 2015


Revised: November 10, 2015
Accepted Article published: December 7, 2015
Final Article published: December 29, 2015

256

2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Vous aimerez peut-être aussi