ArticlePDF Available

Amino-substituted diazocines as pincer-type photochromic switches

Authors:

Abstract and Figures

Azobenzenes are robust, reliable, and easy to synthesize photochromic switches. However, their high conformational flexibility is a disadvantage in machine-like applications. The almost free rotation of the phenyl groups can be restricted by bridging two ortho positions with a CH(2)CH(2) group, as realized in the dihydrodibenzo diazocine framework. We present the synthesis and properties of 3,3'-amino- and 3,3'-acetamido substituted diazocines. Upon irradiation with light of 405 and 530 nm they isomerize from the cis to the trans configuration and back, and thereby perform a pincer-like motion. In the thermodynamically more stable cis isomer the lone pairs of the amino nitrogen atoms point towards each other, and in the trans form they point in opposite directions. The distance between the amino nitrogen atoms changes between 8 Å (cis) and 11 Å (trans isomer).
Content may be subject to copyright.
1
Amino-substituted diazocines as pincer-type
photochromic switches
Hanno Sell1, Christian Näther2 and Rainer Herges*1
Full Research Paper Open Access
Address:
1Otto-Diels Institut für Organische Chemie,
Christian-Albrechts-Universität zu Kiel, Otto-Hahn-Platz 4, 24418 Kiel,
Germany and 2Institut für Anorganische Chemie,
Christian-Albrechts-Universität zu Kiel, Max-Eyth-Str. 2, 24418 Kiel,
Germany
Email:
Rainer Herges* - rherges@oc.uni-kiel.de
* Corresponding author
Keywords:
azobenzene; diazocine; molecular pincer; molecular switches;
photochromic compound
Beilstein J. Org. Chem. 2013, 9, 1–7.
doi:10.3762/bjoc.9.1
Received: 04 September 2012
Accepted: 26 November 2012
Published: 02 January 2013
This article is part of the Thematic Series "Molecular switches and cages".
Guest Editor: D. Trauner
© 2013 Sell et al; licensee Beilstein-Institut.
License and terms: see end of document.
Abstract
Azobenzenes are robust, reliable, and easy to synthesize photochromic switches. However, their high conformational flexibility is a
disadvantage in machine-like applications. The almost free rotation of the phenyl groups can be restricted by bridging two ortho
positions with a CH2CH2 group, as realized in the dihydrodibenzo diazocine framework. We present the synthesis and properties of
3,3’-amino- and 3,3’-acetamido substituted diazocines. Upon irradiation with light of 405 and 530 nm they isomerize from the cis
to the trans configuration and back, and thereby perform a pincer-like motion. In the thermodynamically more stable cis isomer the
lone pairs of the amino nitrogen atoms point towards each other, and in the trans form they point in opposite directions. The dis-
tance between the amino nitrogen atoms changes between 8 Å (cis) and 11 Å (trans isomer).
1
Introduction
Azobenzenes probably are the most frequently used
photochromic switches in chemistry. They are employed as
molecular actuators to drive a number of dynamic machine-like
functions [1]. To achieve sophisticated engineering tasks such
as directed motion at the molecular level [2,3], the geometry
change and the force induced during cistrans isomerization has
to be coupled to the environment. In the macroscopic world,
therefore, machines are made from stiff materials. Azobenzene,
however, is a rather floppy molecule. Both phenyl rings can
rotate with a low activation barrier, and isomerization of the
trans form can occur in two different directions, forming two
different isomers (enantiomers in the parent system) [4]. Power
transmission to neighbouring molecules is inefficient because of
energy transfer to internal conformational motion. The first and
probably most simple measure to make azobenzene stiffer
would be to prevent the phenyl groups from rotating.
Beilstein J. Org. Chem. 2013, 9, 1–7.
2
Scheme 1: Synthesis of 3,3’-diamino-EBAB 4 and its acetamide derivative 5.
Connecting both rings with each other via an alkane bridge is
probably the most straightforward way to achieve that. Such a
molecule, 5,6-dihydrodibenzo[c,g][1,2]diazocine (1), has been
known for more than a hundred years [5]. However, only
recently we discovered that the photophysical properties of 1
(quantum yields, photostationary states) are superior to those of
parent azobenzene, and most of its derivatives [6-8]. In contrast
to azobenzene, diazocine 1 is most stable in its cis configur-
ation, which has a boat conformation. The trans isomer comes
in two conformations: a chair and a twist form, of which the
twist is more stable (Figure 1).
Figure 1: Configurations and conformations of 5,6-
dihydrodibenzo[c,g][1,2]diazocine (1), and DFT (B3LYP/6-31G*) calcu-
lated energies (kcal mol1) relative to the most stable cis boat form.
Dipol moments are given in Debye.
Proper substitution of the diazocine molecular framework is
necessary to control the interaction with the environment or
with other molecules. Therefore, we explore different
approaches to prepare diazocine derivatives. Since the nomen-
clature is not unambiguous and, hence, potentially confusing,
we refer to 5,6-dihydrodibenzo[c,g][1,2]diazocine derivatives as
2,2’-ethylene-bridged azobenzenes (EBABs).
Results and Discussion
Synthesis
Woolley et al. very recently published the synthesis of a 4,4’-
diamino-2,2’-ethylene-bridged azobenzene (4,4’-diamino-
EBAB), which exhibits excellent photophysical properties [9].
In planning our synthesis (and not yet being aware of the results
of the above authors) we were concerned about the fact that
amino substituents in the 4-position with respect to the azo
group would impair photochemical conversion, as this is known
for 4-amino-substituted azobenzenes [10]. We therefore set out
to synthesize 3,3’-diamino-EBAB 4 and derivatives thereof.
Key steps of the synthesis are the introduction of the
substituents and formation of the azo bond. Several approaches
were evaluated, changing the order of the steps and the groups
from which the azo unit was generated. The preferred proce-
dure starts from commercially available 1,2-bis(4-amino-
phenyl)ethane (2). Nitration proceeds almost quantitatively in
ortho position to the ethylene bridge, forming 1,2-bis(2-nitro-4-
aminophenyl)ethane (3) [11]. Intramolecular reductive coupling
of the nitro groups to form the azo unit proceeds with notori-
ously low yields. The most frequently used procedure using Zn
as the reducing agent in Ba(OH)2 [12] or NaOH [13] gives irre-
producible and low yields varying from 2% to not more than
19%. Glucose, however, in basic ethanolic solution turned out
to furnish the azo compound 4 reproducibly in more than 20%
yield. The acetamide derivative 5 is formed by treatment of
3,3’-EBAB 4 with acetic anhydride (Scheme 1).
Beilstein J. Org. Chem. 2013, 9, 1–7.
3
Figure 2: Crystal structures of the cis isomers of 3,3’-diamino-EBAB 4 and its acetamide derivative 5. The atoms of the ethylene bridge are disor-
dered.
The structures of the products were confirmed by 1H and
13C NMR spectroscopy, as well as X-ray crystallography
(Figure 2). As in the parent system 1 the amino and acetamido
derivates 4 and 5 are thermodynamically most stable in their cis
form.
Photochromic Properties
trans-Azobenzene exhibits a high-intensity ππ* band at
λmax = 316 nm (ε = 22,000 LM1cm1) and a symmetry-
forbidden n–π* band at 444 nm with a very low extinction coef-
ficient (ε = 440 LM1cm1) [10]. Irradiation with UV-light of
365 nm converts the trans to the cis isomer. The n–π* absorp-
tion of the nonplanar cis-azobenzene is not formally symmetry
forbidden. Even though it has a rather low extinction coeffi-
cient (1250 LM1cm1), irradiation into the n–π* band leads to
complete conversion back to the trans isomer [14]. Photo-
switching of 5,6-dihydrodibenzo[c,g][1,2]diazocine (unsubsti-
tuted EBAB 1), however, is performed differently [6]. Conver-
sion of cis to trans, as well as isomerization of trans to cis, is
achieved by irradiation into the corresponding n–π* bands. This
is possible because the n–π* bands appear well separated at
different wavelengths (cis: 404 nm, trans: 490 nm) and the tran-
sitions are allowed (albeit weak) in both isomers. Since
substituents in the meta position are known to interact less effi-
ciently with each other than those in the ortho or para positions
we decided to examine EBABs that are 3,3’-substituted, hoping
that the excellent switching properties of the parent system
would be retained. For determination of the ratio of isomers in
photostationary states we used 1H NMR (for details see
Supporting Information File 1).
The UV spectrum of cis-3,3’-diamino-EBAB 4 exhibits a ππ*
transition at 350 nm, and a shoulder at approximately 400 nm,
which arises from the n–π* transition of the azobenzene
Figure 3: UV–vis spectra of the diazocine derivatives 3,3’-diamino-
EBAB 4 and its bisamide derivative 5 in acetonitrile.
substructure (Figure 3). Irradiation of a solution of 4 in acetoni-
trile with light of 405 nm leads to isomerization to the trans
isomer with 34% trans form in the photostationary state (deter-
mined by 1H NMR). The rather low conversion rate is probably
Beilstein J. Org. Chem. 2013, 9, 1–7.
4
Figure 4: Absorbances of solutions of 4 and 5 in acetonitrile at 405 nm (red) and 485 nm (blue) in the corresponding photostationary states upon
alternating irradiation at 405 and 530 nm.
Table 1: Half-life and photostationary states of EBAB 1 and derivatives.
compound half life [h] PSS 405 nm
% trans
PSS 520 nm
% trans
EBAB 14.5a92% <1%
3,3’-diamino-EBAB 474b34% <1%
3,3’-diacetamido-EBAB 546b54% <1%
4,4’-diacetamido-EBAB 4.8c70% <1%
a300 K, n-hexane [6]; b300 K, MeCN; c293 K, DMSO [9].
due to the overlap of the ππ* and n–π* transitions in the cis
form. In the trans isomer the n–π* transition is shifted to longer
wavelengths (~500 nm) and is clearly separated from any other
absorption. Irradiation with light at 450–600 nm therefore
converts the trans isomer completely back to the cis form. As
compared to the bisamino-substituted EBAB 4 the ππ* and the
n–π* bands of the cis isomer of the acetamido derivate 5 are
better separated (shoulder at 320 nm and 400 nm) (Figure 3).
The photostationary state upon irradiation with 405 nm there-
fore rises to 54% trans form. As in the bisamino-substituted
system 4, conversion of 5 back to the cis isomer is quantitative
within the error limits of 1H NMR and UV–vis spectroscopy
(Figure 4).
Thermal stability of the trans isomers
Amino and alkylamino substituents in para position to the azo
group reduce the lifetime of the cis states of azobenzenes
[15,16]. A dramatically shortened half-life of an ortho-dimeth-
ylamino-substituted cis-phenylazopyridine (cis-4-N,N-dimethyl-
amino(3-phenylazo)pyridine) has also been observed [17]. The
reverse effect was observed in our meta-substituted trans-3,3’-
diamino-EBAB 4. While the unsubstituted trans-EBAB 1
exhibits a half-life of 4.5 h in n-hexane solution at room
temperature, trans-3,3’-EBAB 4 isomerizes to the more stable
cis form with a half-life of 74 h. The corresponding half-life of
trans-3,3’-acetamido-EBAB 5 is 46 h (Table 1).
Application as a molecular pincer
In the 1H NMR spectra of cis-4 and cis-5 the four protons of the
ethylene bridge yield a centred multiplet. This symmetry of the
fine structure shows that they divide into two chemically
unequal groups of two chemically equal protons. Hence, there is
no boat inversion at room temperature on the NMR time scale.
According to the X-ray crystal structures the amino nitrogen
lone pairs of cis-3,3’-diamino EBAB 4 point towards each other
and (if extended to more than 6.5 Å) would intersect with an
angle of about 65°. If the acetamido groups in 5 are rotated
appropriately, the N–H bonds (extended to 10.5 Å) intersect
with an angle of about 45°. Thus, the EBAB derivatives 4 and 5
should be suitable as molecular pincers.
To demonstrate this property we studied the binding of ethyl-
enediamine to the two isomers of the acetamido derivative 5
(Figure 5).
We carried out 1H NMR titrations of cis-5 as well as of the
photostationary mixture of cis-5 and trans-5 upon irradiation
with light of a wavelength of 405 nm with ethylenediamine in
acetonitrile. The spectra showed a significant chemically
induced shift (CIS) of the acetamide protons of both isomers
upon addition of ethylenediamine. Equilibrium analysis with
respect to the CIS binding isotherms by means of nonlinear
least-squares methods (for details see Supporting Information
Beilstein J. Org. Chem. 2013, 9, 1–7.
5
Figure 5: DFT-calculated structure (B3LYP/6-31+G**) of a complex of
5 with ethylenediamine as a conceivable model of the binding mode of
3,3’-diacetamido-EBAB 5.
File 1) [18,19] yielded a binding constant for the 1:1 ethylenedi-
amine complex of the cis isomer of Ka,cis = 0.88 ± 0.03 M1.
For the corresponding complex of the trans isomer a slightly
lower binding constant of Ka,trans = 0.61 ± 0.05 M1 was deter-
mined in acetonitrile-d3 at 25 °C.
Conclusion
We prepared ethylene-bridged azobenzene (EBAB) derivatives
with amino and acetamido substituents in the meta position with
respect to the azo group (3,3’-diamino-EBAB 4, and 3,3’-di-
acetamido-EBAB 5). In contrast to azobenzene, and in agree-
ment with the parent EBAB, the cis isomer is more stable than
the trans form. Compared to the parent EBAB, which is a very
efficient photoswitch, the conversion from the cis to the trans
isomer upon irradiation with 405 nm is reduced to 34%
(diamino derivative 4) and 54% (diacetamido derivative 5, cf.
92% in the parent compound). The thermal half-life of the trans
isomer, however, is drastically increased (3,3’-diamino-EBAB
4: 74 h, 3,3’-diacetamido-EBAB 5: 46 h). The EBAB deriva-
tives upon photoisomerization perform a pincer-like motion. Di-
acetamido derivative 5 binds ethylenediamine better in its cis
(closed) form than in its trans configuration (open form).
Experimental
General remarks
All chemicals were purchased from commercial sources and
used without further purification. All NMR spectra were
recorded with instruments of the company Bruker (AC 200,
DRX 500, and AV 600). The assignments of the NMR signals
were confirmed by the evaluation of COSY, HSQC and HMBC
spectra. The chemical shifts of the signals were all referenced to
residual solvent peaks. The mass spectra were recorded on a
Finnigan MAT 8230 instrument. IR spectra were recorded with
a Spectrum 100 instrument from Perkin-Elmer equipped with an
ATR unit from the company Loriot-Oriel. Melting points were
taken without correction. UV–vis spectra were recorded on a
Perkin-Elmer Lambda 14 spectrometer.
Synthesis
1,2-Bis(2-nitro-4-aminophenyl)ethane (3): A solution of 5.0 g
1,2-bis(4-aminophenyl)ethane (24.0 mmol) in 40 mL of sulfuric
acid was warmed up to 60 °C and a solution of 4.4 g
(52.0 mmol) finely grounded sodium nitrate in 45 mL of
sulfuric acid was added dropwise. The mixture was stirred at
60 °C for 6 h and afterwards poured into 200 mL of ice–water.
The resulting suspension was neutralised by the addition of an
aqueous ammonia solution (32%). The red precipitate was
filtered off, washed with water and dried in vacuum over CaCl2.
Yield: 7.2 g (23.6 mmol, 99%). Mp 247–249 °C; 1H NMR
(500 MHz, DMSO-d6) δ 7.06 (d, 4J6,2 = 2.0 Hz, 2H, 6-H), 6.99
(d, 3J3,2 = 8.3 Hz, 2H, 3-H), 6.78 (dd, 3J2,3 = 8.3 Hz, 4J2,6 =
2.1 Hz, 2H, 2-H), 5.58 (s, 4H, 1-NH2), 2.86 (s, 4H, 7-H);
13C NMR (126 MHz, DMSO-d6) δ 149.90 (Cq, C-5), 148.58
(Cq, C-1), 132.85 (d, C-3), 121.85 (Cq, C-4), 119.21 (d, C-2),
108.48 (d, C-6), 33.25 (t, C-7); IR (ATR): 3444 (m), 3363 (s),
3234 (w), 3061 (w), 2947 (w), 2877 (w), 1622 (s), 1513 (vs),
1495 (vs), 1324 (vs), 1272 (s), 1263 (s), 829 (s), 818 (s) cm1;
EIMS (70 eV) m/z (% relative intensity): 302 (16) [M]+, 151
(100); CIMS (isobutane) m/z (% relative intensity): 303 (100)
[M + H]+.
(Z)-11,12-Dihydrodibenzo[c,g][1,2]diazocine-3,8-diamine
(4): A suspension of 1,2-bis(2-nitro-4-aminophenyl)ethane (3)
(1.059 g, 3.5 mmol) in a mixture of 140 mL ethanol and a solu-
tion of 8.8 g (220 mmol) sodium hydroxide in 35 mL water was
heated to 70 °C. A solution of 6.5 g (36 mmol) glucose in
20 mL water was added, and the reaction mixture was stirred
overnight. After cooling to room temperature, 500 mL water
was added, and the resulting mixture was extracted three times
with 100 mL of ethyl acetate. The organic phase was separated
and dried over sodium sulfate, and the solvent was evaporated
in vacuum. From the obtained residue the product was isolated
by flash chromatography (silica gel, cyclohexane/ethyl acetate
1:1) (254 mg, 1.1 mmol, 30%). Mp 193–196 °C; 1H NMR
(600 MHz, DMSO-d6) δ 6.67 (d, 3J6,5 = 8.2 Hz, 2H, 6-H), 6.24
(dd, 3J5,6 = 8.2 Hz, 4J5,3 = 2.3 Hz, 2H, 5-H), 5.97 (d, 4J3,5 =
2.3 Hz, 2H, 3-H), 5.15 (br.s, 4H, 4-NH2), 2.63 (mc, 4H, 7-Ha,
7-Hb,); 13C NMR (150 MHz, DMSO-d6) δ 155.87 (Cq, C-1),
146.76 (Cq, C-2), 130.15 (d, C-6), 115.18 (Cq, C-4), 112.88 (d,
C-5), 103.11 (d, C-3), 30.44 (t, C-7); IR (ATR): 3433 (m), 3344
(m), 3433 (m), 2952 (m), 2850 (m), 1703 (w), 1609 (vs), 1570
(m), 1497 (vs), 1455 (s), 1435 (m), 1303 (s), 1272 (s), 1170
(m), 1142 (m), 1094 (m), 1020 (m), 930 (w), 900 (m), 856 (m),
808 (vs) cm1; EIMS (70 eV) m/z (% relative intensity): 238
(100) [M]+, 209 (92), 193 (46); CIMS (isobutane) m/z (% rela-
tive intensity): 239 (100) [M + H]+.
Beilstein J. Org. Chem. 2013, 9, 1–7.
6
(Z)-N,N'-(11,12-Dihydrodibenzo[c,g][1,2]diazocine-3,8-
diyl)diacetamide (5): In acetic acid anhydride (25 mL), (Z)-
11,12-dihydrodibenzo[c,g][1,2]diazocine-3,8-diamine (4)
(5 mg, 20 mmol) was dissolved. The solution was stirred at
room temperature overnight. Afterwards the solvent was evapo-
rated in vacuum, and the product was obtained as a pale yellow
solid (7 mg, 20 mmol, 100%). Mp 220–221 °C; 1H NMR
(500 MHz, MeCN-d3) δ 8.25 (s, 2H, 5-NH), 7.06 (d, 4J3,5 =
2.2 Hz, 2H, 3-H), 7.03 (dd, 3J5,6 = 8.2 Hz, 4J5,3 = 2.2 Hz, 2H,
5-H), 6.87 (d, 3J3,5 = 2.2 Hz, 2H, 3-H), 2.73 (m, 2H, 7-Ha,),
2.70 (mc, 4H, 7-Ha, 7-Hb), 2.00 (s, 6H, 9-H); 13C NMR
(125 MHz, MeCN-d3) δ 168.36 (Cq, C-8), 155.27 (Cq, C-1),
137.45 (Cq, C-2), 129.94 (d, C-6), 123.17 (Cq, C-4), 117.44 (d,
C-5), 108.59 (d, C-3), 30.29 (t, C-7), 22.96 (q, C-9); IR (ATR):
3253 (m), 3174 (m), 3102 (m), 3048 (m), 2924 (m), 1711 (m),
1680 (m), 1300 (s), 1260 (s), 1020 (s), 980 (m), 957 (m), 899
(m), 883 (m), 814 (s), 763 (m) cm1. EIMS (70 eV) m/z (%
relative intensity): 322 (50) [M]+, 252 (92), 209 (100); CIMS
(isobutane): m/z (% relative intensity) 323 (100) [M + H]+.
Supporting Information
1H NMR spectra of 4 and 5 before and after the irradiation
with 405 nm, and 1H NMR binding study of
3,3-acetamido-EBAB (5) with ethylenediamine. cif-Files of
X-ray crystal structures of cis-4 and cis-5, and gaussian09
input file of the geometry optimization of the complex of
cis-5 and ethylenediamine (DFT B3LYP/6-31+G**).
Supporting Information File 1
Additional NMR spectra and 1H NMR binding study of
3,3-acetamido-EBAB (5) with ethylenediamine.
[http://www.beilstein-journals.org/bjoc/content/
supplementary/1860-5397-9-1-S1.pdf]
Supporting Information File 2
Crystallographic information file of compound cis-4.
[http://www.beilstein-journals.org/bjoc/content/
supplementary/1860-5397-9-1-S2.cif]
Supporting Information File 3
Crystallographic information file of compound cis-5.
[http://www.beilstein-journals.org/bjoc/content/
supplementary/1860-5397-9-1-S3.cif]
Supporting Information File 4
Gaussian09 input file of the geometry optimization of the
complex of cis-5 and ethylenediamine.
[http://www.beilstein-journals.org/bjoc/content/
supplementary/1860-5397-9-1-S4.gjf]
Acknowledgements
We would like to thank the Deutsche Forschungsgemeinschaft
(DFG) for funding through SFB 677 (Function by Switching).
References
1. Merino, E.; Ribagorda, M. Beilstein J. Org. Chem. 2012, 8, 1071–1090.
doi:10.3762/bjoc.8.119
2. Purcell, E. M. Am. J. Phys. 1977, 45, 3–11. doi:10.1119/1.10903
3. Hänggi, P.; Marchesoni, F. Rev. Mod. Phys. 2009, 81, 387–442.
doi:10.1103/RevModPhys.81.387
4. Haberhauer, G.; Kallweit, C. Angew. Chem. 2010, 122, 2468–2471.
doi:10.1002/ange.200906731
Angew. Chem., Int. Ed. 2010, 49, 2418-2421.
doi:10.1002/anie.200906731
5. Duval, H. Bull. Soc. Chim. Fr. 1910, 7, 727.
6. Siewertsen, R.; Neumann, H.; Buchheim-Stehn, B.; Herges, R.;
Näther, C.; Renth, F.; Temps, F. J. Am. Chem. Soc. 2009, 131,
15594–15595. doi:10.1021/ja906547d
7. Jiang, C.-W.; Xie, R.-H.; Li, F.-L.; Allen, R. E. J. Phys. Chem. A 2011,
115, 244–249. doi:10.1021/jp107991a
8. Böckmann, M.; Doltsinis, N. L.; Marx, D. Angew. Chem., Int. Ed. 2010,
49, 3382–3384. doi:10.1002/anie.200907039
9. Samanta, S.; Qin, C.; Lough, A. J.; Woolley, G. A. Angew. Chem. 2012,
124, 6558–6561. doi:10.1002/ange.201202383
Angew. Chem., Int. Ed. 2012, 51, 6452–6455.
doi:10.1002/anie.201202383
10. Rau, H. In Azocompounds in Photochromism, Molecules and Systems;
Dürr, H.; Bouas-Laurent, H., Eds.; Elsevier: Amsterdam, 1990; p 165.
The p-Amino substitution leads to a bathochromic shift of the p–p*
transition and thus to an overlap of the p–p* and n–p* transitions.
11. Matei, S. Rev. Roum. Chim. 1966, 11, 843.
12. Paudler, W. W.; Zeiler, A. G. J. Org. Chem. 1969, 34, 3237–3239.
doi:10.1021/jo01263a004
13. Shine, H. J.; Chamness, J. T. J. Org. Chem. 1963, 28, 1232–1236.
doi:10.1021/jo01040a016
14. Knoll, H. Photoisomerism of Azobenzenes. CRC Handbook of Organic
Photochemistry and Photobiology, 2nd ed.; CRC Press LLC: Boca
Raton, 2004; 89-1–89-16.
15. Schulte-Frohlinde, D. Justus Liebigs Ann. Chem. 1958, 612, 138–152.
doi:10.1002/jlac.19586120115
16. Mourot, A.; Kienzler, M. A.; Banghart, M. R.; Fehrentz, T.;
Huber, F. M. E.; Stein, M.; Kramer, R. H.; Trauner, D.
ACS Chem. Neurosci. 2011, 2, 536–543. doi:10.1021/cn200037p
17. Thies, S.; Sell, H.; Bornholdt, C.; Schütt, C.; Köhler, F.; Tuczek, F.;
Herges, R. Chem.–Eur. J. doi:10.1002/chem.201201698.
18. del Piero, S.; Melchior, A.; Polese, P.; Portanova, R.; Tolazzi, M.
Ann. Chim. (Rome, Italy) 2006, 96, 29–49.
doi:10.1002/adic.200690005
19. de Levie, R. J. Chem. Educ. 1999, 76, 1594–1598.
doi:10.1021/ed076p1594
Beilstein J. Org. Chem. 2013, 9, 1–7.
7
License and Terms
This is an Open Access article under the terms of the
Creative Commons Attribution License
(http://creativecommons.org/licenses/by/2.0), which
permits unrestricted use, distribution, and reproduction in
any medium, provided the original work is properly cited.
The license is subject to the Beilstein Journal of Organic
Chemistry terms and conditions:
(http://www.beilstein-journals.org/bjoc)
The definitive version of this article is the electronic one
which can be found at:
doi:10.3762/bjoc.9.1
... . According to the X-ray crystal structures of the 3,3 -diaminosubstituted diazocine provided by Sell and co-workers, the amino nitrogen atoms were further apart upon switching from Z (8 Å) to E (11 Å) configuration [28]. This molecular pincer motion was exploited to gain photocontrol over macromolecular systems such as peptides [29] and oligonucleotides [30]. ...
... The photochromism of diazocine is reversible with green light with a wavelength of 525 nm or via thermal relaxation with a half-life in hexane of 4.5 h at 28.5 °C [23]. According to the X-ray crystal structures of the 3,3'-diaminosubstituted diazocine provided by Sell and co-workers, the amino nitrogen atoms were further apart upon switching from Z (8 Å) to E (11 Å) configuration [28]. This molecular pincer motion was exploited to gain photocontrol over macromolecular systems such as peptides [29] and oligonucleotides [30]. ...
Article
Full-text available
Unlike azobenzene, the photoisomerization behavior of its ethylene-bridged derivative, diazocine, has hardly been explored in synthetic polymers. In this communication, linear photoresponsive poly(thioether)s containing diazocine moieties in the polymer backbone with different spacer lengths are reported. They were synthesized in thiol-ene polyadditions between a diazocine diacrylate and 1,6-hexanedithiol. The diazocine units could be reversibly photoswitched between the (Z)- and (E)-configurations with light at 405 nm and 525 nm, respectively. Based on the chemical structure of the diazocine diacrylates, the resulting polymer chains differed in their thermal relaxation kinetics and molecular weights (7.4 vs. 43 kDa) but maintained a clearly visible photoswitchability in the solid state. Gel permeation chromatography (GPC) measurements indicated a hydrodynamic size expansion of the individual polymer coils as a result of the Z→E pincer-like diazocine switching motion on a molecular scale. Our work establishes diazocine as an elongating actuator that can be used in macromolecular systems and smart materials.
... However, the switching efficiencies of functionalized b-Azo compounds are also generally low, similar to azobenzene. This has been reported for an amine-substituted b-Azo [20,21], Molecules 2022, 27, 3296 2 of 9 the electronic coupling of which leads to overlapping absorption bands. The degree of isomerization (the proportion of the metastable E-isomers accounted for) of this compound is only 30% after illumination at its excitation wavelength [16][17][18][19]22]. ...
... The successful synthesis of the four compounds was determined by Fourier transform infrared spectroscopy (FTIR), high-resolution mass spectra (HRMS), 1 H NMR and 13 switching efficiencies of functionalized b-Azo compounds are also generally low, similar to azobenzene. This has been reported for an amine-substituted b-Azo [20,21], the electronic coupling of which leads to overlapping absorption bands. The degree of isomerization (the proportion of the metastable E-isomers accounted for) of this compound is only 30% after illumination at its excitation wavelength [16][17][18][19]22]. ...
Article
Full-text available
Molecular photoswitches are considered to be important candidates in the field of solar energy storage due to their sensitive and reversible bidirectional optical response. Nevertheless, it is still a daunting challenge to design a molecular photoswitch to improve the low solar spectrum utilization and quantum yields while achieving charging and discharging of heat without solvent assistance. Herein, a series of visible-light-driven ethylene-bridged azobenzene (b-Azo) chromophores with different alkyne substituents which can undergo isomerization reactions promoted in both directions by visible light are reported. Their visible light responsiveness improves their solar spectrum utilization while also having high quantum yields. In addition, as the compounds are liquids, there is no need to dissolve the compounds in order to exploit this switching. The photoisomerization of b-Azo can be adjusted by alkyne-related substituents, and hexyne-substituted b-Azo is able to store and release photothermal energy with a high density of 106.1 J·g−1, and can achieve a temperature increase of 1.8 °C at a low temperature of −1 °C.
... 210 Exchanging an azobenzene scaffold for a diazocine poses a way to invert the properties of a given system and gives access to additional tunability. 215,216 Moreover, in diazocines, the absorption bands associated with the S 0 -S 1 transition are well-separated, giving rise to azobenzenes that can be switched bidirectionally with visible light with high quantum yields. 210 The strain that the E isomer experiences results in t 1/2 of a few hours in the original CH 2 -CH 2 -bridged diazocines. ...
Article
Full-text available
Molecular photoswitches enable dynamic control of processes with high spatiotemporal precision, using light as external stimulus, and hence are ideal tools for different research areas spanning from chemical biology to smart materials. Photoswitches are typically organic molecules that feature extended aromatic systems to make them responsive to (visible) light. However, this renders them inherently lipophilic, while water-solubility is of crucial importance to apply photoswitchable organic molecules in biological systems, like in the rapidly emerging field of photopharmacology. Several strategies for solubilizing organic molecules in water are known, but there are not yet clear rules for applying them to photoswitchable molecules. Importantly, rendering photoswitches water-soluble has a serious impact on both their photophysical and biological properties, which must be taken into consideration when designing new systems. Altogether, these aspects pose considerable challenges for successfully applying molecular photoswitches in aqueous systems, and in particular in biologically relevant media. In this review, we focus on fully water-soluble photoswitches, such as those used in biological environments, in both in vitro and in vivo studies. We discuss the design principles and prospects for water-soluble photoswitches to inspire and enable their future applications.
Article
Full-text available
A macrocyclic azobenzene with unique thermal stability, demonstrating potential for use in optical data storage material is presented: The Z‐isomer of this novel photoswitch exhibits unparalleled thermal stability, with a thermal half‐life surpassing 120 years at 25 °C. This stability is attributed to the strategic fluorination at two ortho‐ and both para‐positions. Comparative analyses involving its non‐fluorinated counterpart, ortho‐only‐fluorinated variant, and open‐chain analog are performed. Employing NMR and UV–vis spectroscopy, X‐ray diffraction, alongside Arrhenius, Eyring, and DFT calculations, revealed insights into its extraordinary stability. Furthermore, when incorporated into poly(methylmethacrylate), this material showcase efficient switching with visible light in the solid state, emphasizing its potential for optical data storage applications.
Article
Full-text available
Diazocines are azobenzene derived macrocyclic photoswitches with well resolved photostationary states for the (E)- and (Z)-isomers, which improves their addressability by light. In this work, effective procedures for the stannylation and borylation of diazocines in different positions are reported. Their use in Stille cross-coupling and Suzuki cross-coupling reactions with organic bromides is demonstrated in yields of 47–94% (Stille cross-coupling) and 0–95% (Suzuki cross-coupling), respectively.
Thesis
In many areas of the observational and experimental sciences data is scarce. Observation in high-energy astrophysics is disrupted by celestial occlusions and limited telescope time while laboratory experiments in synthetic chemistry and materials science are both time and cost-intensive. On the other hand, knowledge about the data-generation mechanism is often available in the experimental sciences, such as the measurement error of a piece of laboratory apparatus. Both characteristics make Gaussian processes (GPs) ideal candidates for fitting such datasets. GPs can make predictions with consideration of uncertainty, for example in the virtual screening of molecules and materials, and can also make inferences about incomplete data such as the latent emission signature from a black hole accretion disc. Furthermore, GPs are currently the workhorse model for Bayesian optimisation, a methodology foreseen to be a vehicle for guiding laboratory experiments in scientific discovery campaigns. The first contribution of this thesis is to use GP modelling to reason about the latent emission signature from the Seyfert galaxy Markarian 335, and by extension, to reason about the applicability of various theoretical models of black hole accretion discs. The second contribution is to deliver on the promised applications of GPs in scientific data modelling by leveraging them to discover novel and performant molecules. The third contribution is to extend the GP framework to operate on molecular and chemical reaction representations and to provide an open-source software library to enable the framework to be used by scientists. The fourth contribution is to extend current GP and Bayesian optimisation methodology by introducing a Bayesian optimisation scheme capable of modelling aleatoric uncertainty, and hence theoretically capable of identifying molecules and materials that are robust to industrial scale fabrication processes.
Article
Full-text available
Photoswitchable molecules display two or more isomeric forms that may be accessed using light. Separating the electronic absorption bands of these isomers is key to selectively addressing a specific isomer and achieving high photostationary states whilst overall red-shifting the absorption bands serves to limit material damage due to UV-exposure and increases penetration depth in photopharmacological applications. Engineering these properties into a system through synthetic design however, remains a challenge. Here, we present a data-driven discovery pipeline for molecular photoswitches underpinned by dataset curation and multitask learning with Gaussian processes. In the prediction of electronic transition wavelengths, we demonstrate that a multioutput Gaussian process (MOGP) trained using labels from four photoswitch transition wavelengths yields the strongest predictive performance relative to single-task models as well as operationally outperforming time-dependent density functional theory (TD-DFT) in terms of the wall-clock time for prediction. We validate our proposed approach experimentally by screening a library of commercially available photoswitchable molecules. Through this screen, we identified several motifs that displayed separated electronic absorption bands of their isomers, exhibited red-shifted absorptions, and are suited for information transfer and photopharmacological applications. Our curated dataset, code, as well as all models are made available at https://github.com/Ryan-Rhys/The-Photoswitch-Dataset.
Chapter
Diazocine was first synthesized more than 110 years ago; however, its photochemical properties were not investigated until 2009. Structurally, diazocines are related to azobenzenes. The two phenyl rings are bridged by a CH 2 CH 2 group in ortho position forming an 8‐membered ring. The small structural variation gives rise to improved photophysical properties, such as bathochromically shifted switching wavelengths, very high quantum yields, superior photoconversion yields, and very high fatigue resistance. Within the last decade, new synthetic strategies were developed and a number of derivatives were synthesized. Diazocines found applications in material science such as photoactuators and mechanochromes, in biochemistry for light‐induced protein and DNA structure switching, and particularly in photopharmocology.
Article
Full-text available
Control over molecular motion represents an important objective in modern chemistry. Aromatic azobenzenes are excellent candidates as molecular switches since they can exist in two forms, namely the cis (Z) and trans (E) isomers, which can interconvert both photochemically and thermally. This transformation induces a molecular movement and a significant geometric change, therefore the azobenzene unit is an excellent candidate to build dynamic molecular devices. We describe selected examples of systems containing an azobenzene moiety and their motions and geometrical changes caused by external stimuli.
Article
Photochromism is simply defined as the light induced reversible change of colour. The field has developed rapidly during the past decade as a result of attempts to improve the established materials and to discover new devices for applications. As photochromism bridges molecular, supramolecular and solid state chemistry, as well as organic, inorganic and physical chemistry, such a treatment requires a multidisciplinary approach and a broad presentation. The first edition (1990) provided an enormous amount of new concepts and data, such as the presentation of main families based on the pericyclic reaction mechanism, the review of new families, some bimolecular photocycloadditions and some promising systems. This new edition provides an efficient entry into this flourishing field, with the core content retained from the original work to provide a basic introduction into the different subjects.
Article
Alter Hund lernt neuen Trick: Beim chiralen Azobenzol‐Derivat 1 konnte eine lichtinduzierte trans→cis‐Isomerisierung erstmalig auch räumlich gerichtet (unidirektional) verwirklicht werden (siehe Schema). Die Anwendungsbreite von Azobenzol‐Derivaten wird damit um einen wichtigen Effekt erweitert.
Article
4,4′-Divinylhydrazobenzene (I) has been prepared. The rates of the acid-catalyzed rearrangement of I in 95% ethanol at 0° and 25°, and in 75% t-butyl alcohol at 25° were determined. The rearrangement was followed by both the Bindschedler's Green titration method and spectrophotometrically. In each case the rearrangement was first-order in acid, over the range of acid concentrations 0.001 to 0.05 M. The rate of rearrangement of I is faster than that of hydrazobenzene. The product of rearrangement is neither the semidine nor the bis-p-aminophenylbutadiene. The product decomposes slowly with charring above 300°, is insoluble in ethanol, and has no unsaturation. The rearrangement is not accompanied by extensive disproportionation as is observed in the rearrangement of other di-p-substituted hydrazobenzenes. It is proposed that the product is a polymer that accompanies the o-benzidine type of rearrangement. We have observed, with the rearrangement of 2,2′-hydrazonaphthalene, the connection between acid order and acid concentration which Banthorpe, Hughes, and Ingold2 discovered in the rearrangement of 1-phenyl-2-β-naphthylhydrazine; that is, that the trend is from second-order at higher to first-order at lower acid concentrations.
Article
The bistability of spin states (e.g., spin crossover) in bulk materials is well investigated and understood. We recently extended spin-state switching to isolated molecules at room temperature (light-driven coordination-induced spin-state switching, or LD-CISSS). Whereas bistability and hysteresis in conventional spin-crossover materials are caused by cooperative effects in the crystal lattice, spin switching in LD-CISSS is achieved by reversibly changing the coordination number of a metal complex by means of a photochromic ligand that binds in one configuration but dissociates in the other form. We present mathematical proof that the maximum efficiency in property switching by such a photodissociable ligand (PDL) is only dependent on the ratio of the association constants of both configurations. Rational design by using DFT calculations was applied to develop a photoswitchable ligand with a high switching efficiency. The starting point was a nickel-porphyrin as the transition-metal complex and 3-phenylazopyridine as the photodissociable ligand. Calculations and experiments were performed in two iterative steps to find a substitution pattern at the phenylazopyridine ligand that provided optimum performance. Following this strategy, we synthesized an improved photodissociable ligand that binds to the Ni-porphyrin with an association constant that is 5.36 times higher in its trans form than in the cis form. The switching efficiency between the diamagnetic and paramagnetic state is efficient as well (72 % paramagnetic Ni-porphyrin after irradiation at 365 nm, 32 % paramagnetic species after irradiation at 440 nm). Potential applications arise from the fact that the LD-CISSS approach for the first time allows reversible switching of the magnetic susceptibility of a homogeneous solution. Photoswitchable contrast agents for magnetic resonance imaging and light-controlled magnetic levitation are conceivable applications.
Article
The synthesis of dibenzo[b,f][1,2]diazocine (1) and its reduction to the 5,6-dihydrodibenzo[b,f][1,2]diazocine (2) is described. The spectral properties of the latter compound indicate that its central ring, because of the presence of 10 π electrons, is somewhat resonance-stabilized.
Article
A macro is described that computes the precision of the parameters obtained with Microsoft Excel Solver, and several examples illustrate its use. Keywords (Audience): Upper-Division Undergraduate
Article
Photochromic channel blockers provide a conceptually simple and convenient way to modulate neuronal activity with light. We have recently described a family of azobenzenes that function as tonic blockers of K(v) channels but require UV-A light to unblock and need to be actively switched by toggling between two different wavelengths. We now introduce red-shifted compounds that fully operate in the visible region of the spectrum and quickly turn themselves off in the dark. Furthermore, we have developed a version that does not block effectively in the dark-adapted state, can be switched to a blocking state with blue light, and reverts to the inactive state automatically. Photochromic blockers of this type could be useful for the photopharmacological control of neuronal activity under mild conditions.
Article
Es wurde die thermische cis → trans-Umlagerung substituierter Azobenzole in verschiedenen Lösungsmitteln gemessen. Die erhaltenen Werte für die Aktivierungsenergie Ea liegen zwischen 20 und 23.5 kcal/Mol, die Werte für die Aktionskonstanten A zwischen 1010.2 und 1012 sec−1. Viel niedrigere Werte wurden nur festgestellt, wenn Katalyse nachgewiesen werden konnte. Die Ea-Werte für die katalytische Umlagerung liegen zwischen 5.5 und 12 kcal/Mol, die A-Werte zwischen 100.28 und 104.14 sec −1. Eine nicht-adiabatische Umwandlung über einen Triplett-Zustand ist nicht Ursache für das Auftreten kleiner A-Konstanten.