Você está na página 1de 8

DOI: 10.1002/chem.

201102981

Light-Harvesting in Multichromophoric Rotaxanes


Maria E. Gallina,[a] Bilge Baytekin,[b, c] Christoph Schalley,*[b] and Paola Ceroni*[a]
Abstract: Two rotaxanes with benzyl
ether axles and tetralactam wheels
were synthesized through an anion
template effect. They carry naphthalene chromophores attached to the
stopper groups and a pyrene chromophore attached to the wheel. The difference between the two rotaxanes is
represented by the connecting unit of

the naphthyl chromophore to the rotaxane axle: a triazole or an alkynyl


group. Both rotaxanes exhibit excellent
light-harvesting properties: excitation
Keywords: energy transfer molecular antenna photochemistry
photophysics rotaxanes

Introduction
Rotaxanes, catenanes, and other interlocked molecules[1]
have been extensively studied for the development of mechanically bonded molecular machines,[2] such as shuttles,
switches, and even muscles. On the other hand, only few examples have been reported on rotaxanes containing multiple
chromophoric units either in the axle or in the wheel component, in which energy transfer processes take place upon
light excitation.[3] In this view, we would like to couple the
concept of light-harvesting antennae and rotaxanes.
An antenna for light harvesting is an organized multicomponent system in which many chromophoric molecular units
absorb the incident light and then channel the excitation
energy to a common acceptor component.[4] Most of the artificial light-harvesting antennae are based on dendrimers,[5]
well-defined tree-like macromolecules that are characterized
by a high degree of order and the possibility to contain selected chemical units in predetermined sites of their structure.[6] Examples of rotaxanes containing dendritic wedges
in the axle and/or in the wheel,[7] as well as dendrimers
based on rotaxanes[8] or pseudo-rotaxanes[9] branching units
are present in the literature, but none of them has been designed to perform as a molecular antenna. To make a rotax-

[a] Dr. M. E. Gallina, Prof. Dr. P. Ceroni


Department of Chemistry G. Ciamician, Via Selmi 2
40126 Bologna (Italy)
E-mail: paola.ceroni@unibo.it
[b] Dr. B. Baytekin, Prof. Dr. C. Schalley
Institut fr Chemie und Biochemie der Freien Universitt
Takustr. 3, 14195 Berlin (Germany)
E-mail: christoph@schalley-lab.de
[c] Dr. B. Baytekin
Present address: Non-equilibrium Energy Research Center
Northwestern University, 2145 Sheridan Rd.
Evanston, IL 60208 (USA)

1528

of the naphthalene chromophores is


followed by energy transfer to the
pyrene unit with efficiency higher than
90 % in both cases. This represents an
example of light-harvesting function
among chromophores belonging to mechanically interlocked components,
that is, the axle and the wheel of the
rotaxanes.

ane working as a light-harvesting antenna, a possible strategy is to place several donor chromophores on dendritic
stoppers and the acceptor one on the wheel of the rotaxane.
This construction enabled us to investigate the energy transfer process between mechanically bonded functional units.
By proper design of this kind of supramolecular structures,
future developments along this research line can result in
light-activated molecular shuttles in which the light absorbed by the antenna is used to promote a molecular
movement. This will increase the efficiency of light absorption.
In the present paper, the two rotaxanes investigated (R1
and R2 in Scheme 1) consist of six naphthyl units in the
stoppers and a pyrene chromophore appended to the
wheel.[10, 11] The difference between the two rotaxanes is represented by the connecting unit of the naphthyl chromophore to the rotaxane axle, a triazole or an alkynyl group in
the rotaxanes R1 and R2, respectively. The stoppers S1 and
S2, axle A1, macrocycles W and W-Py, methylnaphthalene
and methylpyrene (Scheme 1) serve as model compounds
and were investigated for comparison purposes. The results
obtained evidence a highly efficient energy transfer from
the naphthalene chromophores of the stoppers to the
pyrene chromophore bound to the wheel.

Results and Discussion


Syntheses: Scheme 1 shows the two rotaxanes R1 and R2 together with their syntheses. Starting with parafuchsine (1),
the trityl stopper 5 carrying three acetylene groups can be
synthesized as the key intermediate through a sequence of
diazotation and iodination providing compound 2, Friedel
Crafts alkylation yielding the tris-iodotritylphenol 3 followed by Sonogashira coupling to give compound 4, and deprotection of the acetylenes.[12] Intermediate 5 can then
either be coupled to naphthalene through a second Sonoga-

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eur. J. 2012, 18, 1528 1535

FULL PAPER

Scheme 1. Synthesis of the photoactive rotaxanes R1 and R2. a) 1) NaNO2, 0 8C, 1 H; H2O; 2) KI, 80 8C, 2 H; H2O, 82 %. b) H2SO4 (cat.), 80 8C, 4 H;
76 %. c) [PdACHTUNGRE(PPh3)2Cl2] (cat.), PPh3, CuI, TMSacetylene, 40 8C, 4 H; NEt3, DMF, 40 %. d) KOH, RT, MeOH/H2O (1:1), 85 %. e) [PdACHTUNGRE(PPh3)2Cl2] (cat.),
CuI, RT, 21 H; NEt3, DMF, 56 %. f) CuSO4H2O, Na-ascorbate, RT, 2 d, H2O/THF (1:1), 88 %. g) dibenzo[18]crown-6, K2CO3, 1,4-di-(bromomethyl)benzene, CH2Cl2, RT, 6 d, R1: 82 %, R2: 85 %. Methylpyrene (Py), (1-methyl)naphthalene (Naph), and the axle A1 of rotaxane R1 were used as control
compounds.

shiraHagihara coupling reaction[13] to yield stopper S1 or it


can be converted to the tris-triazole stopper S2 with the corresponding azide through click chemistry.[14] With these two
stoppers in hand, the two rotaxanes R1 and R2 are prepared

Chem. Eur. J. 2012, 18, 1528 1535

with quite high yields of > 80 % through the Vgtle anion


template effect.[15]

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemeurj.org

1529

P. Ceroni, C. Schalley et al.

Photophysical properties: All experiments were performed


in dichloromethane/acetonitrile (1:1, v/v) at 298 K and in dichloromethane/chloroform (1:1, v/v) at 77 K to get a transparent matrix. The most relevant photophysical properties
of the rotaxanes R1 and R2, and of their model compounds
are reported in Table 1.
Table 1. Photophysical properties of the rotaxanes R1 and R2, and their
model compounds in air-equilibrated dichloromethane/acetonitrile (1:1,
(v/v) at 298 K and in dichloromethane/chloroform (1:1, v/v) rigid matrix
at 77 K.

R1
R2
W-Py
Py
A1
S1
Naph
S2

298 K
absorption
fluorescence
lmax
Fem[b]
e
t
lmax
[nm][a] [ m1 cm1] [nm]
[ns]

77 K
phosphorescence
t
lmax
[nm]
[ms]

328
346
346
343
328
328
276
265

610
610
610
602
550
550
475
475

152 300
25 200
24 800
41 200
136 000
68 000
4900
27 200

411
411
407
378
371
371
336
336

0.20
0.23
0.24
0.01
0.58
0.54
0.08
0.04

4.6
5.0
5.7
16.3
1.3
1.5
10
14

250
300
250
250
310
210
1320
[c]

[a] Lowest energy absorption band. [b] Excitation was performed in the
lowest energy absorption band. [c] The phosphorescence intensity is too
weak to measure the excited state lifetime.

Let us consider at first the wheel and the axle components


separately, as well as their model compounds, to better understand the more complex behavior of the two rotaxanes.
The wheel component: Compound W-Py (Scheme 1) was investigated in comparison to Py, and the macrocycle with no
appended functionality (W in Scheme 1).
The macrocycle W-Py shows two intense absorption
bands with maxima at l = 280 and 346 nm (black solid line
in Figure 1). These spectral features are reminiscent of those
observed for Py (dashed line in Figure 1), but they are characterized by a lower intensity, a less-resolved vibrational
structure, and a slight red-shift. The absorption spectrum of
the unfunctionalized macrocycle W (gray line in Figure 1) is
a tail-like band below l = 320 nm. From these data it is evident that the main chromophoric unit of W-Py is the pyrene
component. Its photophysical properties are affected by the
delocalization of the p electrons extended to the directly
linked aromatic group of the macrocycle.
As to the emission properties, W is not luminescent,
whereas W-Py and Py exhibit an emission band with
maxima at l = 407 and 377 nm, respectively. The differences
between the two molecules are not only the red-shift and
less-resolved vibrational structure of the W-Py emission
band, but also the much higher emission quantum yield
(Fem = 24 % and 1 % for W-Py and Py, respectively) and the
lower excited-state lifetime (t = 5.7 and 16.3 ns for W-Py
and Py, respectively). In the case of W-Py, the fluorescence
quantum yield is sensitive to the excitation wavelength: it is
24 % upon excitation at the lowest energy absorption band
and it decreases upon excitation at l < 300 nm. This is due

1530

www.chemeurj.org

Figure 1. Absorption (left axis) and emission spectra (right axis) of W-Py
(black solid lines) and Py (black dashed lines), as well as absorption spectrum of W (gray solid line) in air-equilibrated acetonitrile/dichloromethane (1:1, v/v) at 298 K. Emission spectra are normalized at their maximum. lex = 320 nm.

to the fact that in this spectral region, light is absorbed not


only by the pyrene unit but also by the tetralactam macrocycle (see W for comparison), which is not effective to populate the luminescent excited state of the pyrene chromophore. This is confirmed by the excitation spectrum performed at lem = 420 nm: it shows an intensity ratio of the
bands at l = 280 and 346 nm lower than that observed in the
absorption spectrum.
At 77 K in dichloromethane/chloroform rigid matrix, WPy presents a phosphorescence band with a maximum at l =
610 nm, similar to that of Py (lmax = 602 nm). The lifetime of
the phosphorescent state is 250 ms for both compounds.
The two axle components: The axle of rotaxane R1, hereafter called A1, has been examined in comparison with the
corresponding stopper S1.
The absorption spectra of A1 and S1 (black solid line in
Figure 2) have identical shape. Due to the limited amount of
sample available for A1, its molar absorption coefficient
was not measured, but it can be assumed to be the double
of that of S1. This hypothesis, based on the expected similar
photophysical properties of the two compounds, is corroborated by the strong analogy found also in their luminescent
properties. Their emission spectra are superimposed (for S1,
see the black solid line in Figure 2), the corresponding fluorescence quantum yields are very close (54 and 58 % for S1
and A1, respectively) and the luminescent excited-state lifetimes (Table 1) are identical within the experimental error.
Phosphorescence was observed in dichloromethane/chloroform at 77 K for both A1 and S1 with maxima located at l =
550 nm with a slightly longer lifetime of the phosphorescent
excited state in the case of A1 (Table 1).
The photophysical properties of A1 and S1 are significantly different from those of Naph (dashed lines in Figure 2),
which exhibits a strongly blue-shifted absorption with lower

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eur. J. 2012, 18, 1528 1535

Light-Harvesting in Rotaxanes

FULL PAPER
Indeed, the lowest singlet excited state of the naphthyl chromophore in the stopper has one more deactivation pathway,
that is, the excimer formation.
A weak phosphorescence was observed in rigid matrix at
77 K for S2, very similar in shape and position to that of
naphthalene, further confirming the similarity between the
photophysical properties of the two species.
Rotaxanes: Rotaxane R1 presents an intense and structured
absorption band in the UV region with maximum at l =
328 nm (black lines in Figure 3) and its absorption spectrum

Figure 2. Absorption (left axis) and emission spectra (right axis) of S1


(black solid lines), S2 (gray lines), and Naph (black dashed lines) in airequilibrated acetonitrile/dichloromethane (1:1, v/v) at 298 K. Note that
the molar absorption coefficient of Naph is multiplied by three for comparison purposes. Emission spectra are normalized at their maximum.
lex = 320 nm for S1 and lex = 275 nm for S2 and Naph.

intensity, even if we take into account the number of naphthyl units in the stopper component S1. Also the fluorescence band is blue shifted and characterized by a lower
emission quantum yield (Table 1). As a result, we can conclude that the spectroscopic features of the naphthyl units
present in A1 and S1 are strongly modified by the presence
of the phenylacetylene group, which participates strongly
to the delocalization of p electrons. This is confirmed by the
appreciable shift of the absorption spectrum towards higher
wavelengths that was reported in the literature[16] in going
from naphthalene to 1-ethinyl-naphtalene in dimethylsulfoxide solution. The discrepancy between the properties of
naphthalene and those of A1 and S1 can be found in the
analysis of the triplet excited states as well: the phosphorescence of Naph is strongly blue-shifted (the emission maximum is located at l = 475 nm) and characterized by a much
longer lifetime (Table 1).
In the case of the rotaxane R2, the axle component was
not available and only the stopper unit S2 was investigated.
In contrast to S1, the absorption and emission spectra of
S2 (gray lines in Figure 2) are very similar to that of Naph.
This result points out that the functionalization with a triazole group through a methylene spacer does not significantly
affect the photophysical properties of the naphthalene chromophore.
The molar absorption coefficient of S2 is somewhat
higher than that expected for three naphthyl units, particularly in the l = 250280 nm region. This can be partly explained by the contribution of the other chromophoric units
of the stopper. The emission spectrum of S2 shows a tail
above l = 350 nm, which can be related to naphthalene excimers, as commonly observed for supramolecular systems
containing multiple naphthyl units linked by flexible
spacers.[17] This hypothesis is supported by the lower emission quantum yield observed for S2 compared to Naph.

Chem. Eur. J. 2012, 18, 1528 1535

Figure 3. Absorption (left axis) and emission spectra (right axis) of rotaxane R1 (black solid lines), A1 (gray solid lines), and W-Py (gray dashed
lines) in air-equilibrated acetonitrile/dichloromethane (1:1, v/v) at 298 K.
Emission spectra are normalized at their maximum value. lex = 328 nm.

is superimposed to the sum of the absorption spectra of the


axle A1[18] (gray line in Figure 3) and the wheel W-Py
(dotted line in Figure 3), demonstrating that there is no
ground-state interaction between the two mechanically interlocked components of the rotaxane.
Upon excitation at l = 324 nm, where light is absorbed by
both the naphthalene units of the axle (91 %) and the
pyrene group of the wheel (9 %), R1 presents an emission
spectrum very similar to that of W-Py. The emission maximum is slightly red-shifted and the corresponding excitedstate lifetime is slightly lower (Table 1). This result demonstrates a small effect of the close axle component on the excited-state properties of the pyrenyl chromophore. A very
small band is observed in the spectral region where the
naphthyl chromophores of A1 emit. The corresponding lifetime is below 0.5 ns, that is, the equipment time resolution,
evidencing that the axle emission is strongly quenched and
the pyrene emission is sensitized.
To get more quantitative data, the emission quantum
yield of R1 was measured also upon excitation at l = 342 nm
where 17 % of the light are absorbed by the pyrenyl chromophore of the wheel. The value of Fem obtained upon excitation at l = 324 nm was the same (0.20) within experimental
error. Moreover, a close match of the absorption and excita-

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemeurj.org

1531

P. Ceroni, C. Schalley et al.

tion spectrum performed at lem = 450 nm suggests an almost


quantitative energy transfer efficiency (hET > 90 %).
The phosphorescence spectrum of R1 in rigid matrix at
77 K is superimposed to that of W-Py upon excitation at l =
328 nm, where both A1 and W-Py absorb. No phosphorescence of the naphthyl chromophore was observed in agreement with an efficient energy transfer process, which deactivates its lowest singlet excited state, thus preventing intersystem crossing to populate the lower lying triplet excited
state.
Rotaxane R2 has a strong absorption with the lowest
energy band at l = 346 nm perfectly coincident in shape and
intensity with that of the wheel W-Py. The absorption spectrum of R2 is similar to the sum of two stopper components
S2 (the axle component is not available) and a wheel component W-Py.
Upon excitation at l = 348 nm, where only the wheel absorbs light, the emission spectrum of R2 presents a maximum at l = 411 nm, very similar to that of W-Py with a
slight red-shift and an identical emission quantum yield
(Table 1) within the experimental error. Upon excitation at
l = 267 nm, where approximately 55 % of the light are absorbed by the axle component, a strong emission at l =
411 nm and an emission band with a maximum at l =
335 nm, typical of the naphthyl chromophores of the axle is
observed (Figure 4). The excited-state lifetime measured at

Figure 4. Emission spectra of rotaxane R2 (black solid line), S2 (gray


solid line), and W-Py (dashed gray line) in air-equilibrated acetonitrile/dichloromethane (1:1, v/v) at 298 K. The emission spectra of R2 and W-Py
are normalized at their maximum for comparison purposes. lex = 267 nm.

lem = 411 nm is very close to that observed for W-Py, whereas the emission intensity decay measured at l = 335 nm is
biexponential: t1 around 0.8 ns and t2 = 14 ns. The first component can be assigned to the quenched emission of the
naphthyl chromophore. The second component (t2) is identical to the fluorescent excited-state lifetime measured for S2,
suggesting that an impurity of axle not threaded into the
wheel is present.

1532

www.chemeurj.org

From the value of the shorter lifetime, we can estimate


the efficiency of the quenching process (hq) by the following
Equation (1):
hq 1t=t0

where t and t0 are the lifetime in the presence and in the


absence of the wheel component.
In the present case, the value of hq is 94 %. Therefore,
also in this case a very efficient energy transfer takes place
from the stopper chromophores to the wheel unit.
The phosphorescence of R2 observed in rigid matrix at
77 K with lex = 348 nm is very similar to that observed for
W-Py.

Conclusion
The investigated rotaxanes consist of a multichromophoric
axle containing six naphthyl units in the two stoppers and a
pyrenyl unit appended to the wheel. In the case of the rotaxane R1, the naphthyl chromophores are strongly perturbed
by the alkynyl bridge, so that their absorption and emission
bands occur at lower energy, with molar absorption coefficient and emission quantum yield higher than those of (1methyl)naphthalene. On the other hand, for the rotaxane
R2, the naphthyl chromophores are linked by a methylene
spacer to a triazole group: the photophysical properties are
very similar to those of the model compound. For both rotaxanes, no ground-state interaction was evidenced between
the wheel and the axle, but strong interactions take place in
the excited states. Indeed, upon excitation of the naphthyl
chromophores of the axle, a very efficient energy transfer to
the pyrenyl unit of the wheel occurs.
R1 and R2 couple the function of mechanically interlocked molecules to that of light-harvesting antennae. To
the best of our knowledge, it is the first time in which
energy transfer processes have been reported between multiple donor chromophores placed on the axle and an acceptor one on the wheel. This study can have future developments in the field of molecular machines. Up to now, lightpowered molecular machines are based on the absorption of
light by a single chromophore, followed, for example, by an
electron transfer reaction which drives the molecular movement.[1] Properly designed dendritic light-harvesting stoppers
would offer the advantage to absorb a larger fraction of the
incident light and to extend the spectral window useful to
drive the mechanical movement.

Experimental Section
General methods: Reagents were purchased from Aldrich, ACROS, or
Fluka and used without further purification. The macrocycles W and WPy were synthesized according to literature procedures.[8, 19] CH2Cl2,
MeOH, EtOAc, and toluene were dried and distilled prior to use by
usual laboratory methods, whereas all other solvents were used from
commercial sources. Thin-layer chromatography (TLC) was performed

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eur. J. 2012, 18, 1528 1535

Light-Harvesting in Rotaxanes

FULL PAPER

on pre-coated silica gel 60/F254 plates (Merck KGaA). Silica gel (0.04
0.063 mm), (0.630.100 mm), Merck KGaA, and Al2O3 (neutral) (Riedel
de Han) were used for column chromatography.

(62.5 MHz, CDCl3): d = 0.00, 64.2, 94.6, 104.8, 114.6, 121.0, 130.8, 131.3,
132.2, 137.8, 146.7, 153.9 ppm; MS (EI) (220 8C) m/z (%): 624.3 (100)
[M]C + .

4-[Tris(4-ethynylphenyl)methyl]phenol (5): Compound 4 (154 mg,


0.245 mmol) and KOH (138 mg, 2.46 mmol) were stirred in a mixture of
methanol and water (1:1, 10 mL) overnight. The mixture was neutralized
by a few drops of HCl (conc.) and the desired compound was extracted
into dichloromethane, dried over MgSO4, filtered, and the solvent was
evaporated. When necessary, a short silica column was employed to further purify the compound, by eluting with dichloromethane to give the
desired product (85.4 mg, 85 %). 1H NMR (250 MHz, CDCl3): d = 3.05 (s,
3 H), 6.64 (d, J = 9 Hz, 2 H; PhHphenol), 6.85 (d, J = 9 Hz, 2 H; PhHphenol),
7.04 (d, J = 8.3 Hz, 2 H; PhH), 7.28 ppm (d, J = 8.3 Hz, 2 H; PhH);
13
C NMR (100 MHz, CDCl3) d = 64.24, 83.39, 99.98, 114.75, 120.07,
130.87, 131.58, 132. 19, 137.64, 147.02, 154.09 ppm; MS (ESI-FTICR) m/z
(%): 407.15 (100) [MH] , 815.30 (5) [2 MH] .

H and 13C NMR spectra were recorded with Bruker AC 250 and AM
400 instruments. All chemical shifts are reported in ppm with solvent signals taken as internal standards; coupling constants are in Hertz. Electrospray mass spectra (ESI-MS) were recorded on an Agilent 6210 ESITOF, Agilent Technologies, Santa Clara, CA, USA and Varian/Ionspec
FTICR mass spectrometers. The solvent flow rate was adjusted to
4 mL min1, the spray voltage set to 4 kV. Drying gas flow rate was set to
15 psi (1 bar). All other parameters were adjusted for a maximum abundance of the desired ions. Electron ionization mass spectra (EI-MS) were
measured on a MAT 711 spectrometer, Varian MAT, Bremen. The electron energy was set to 80 eV. Elemental analyses of the macrocyclic compounds reported here almost always fail due to solvent molecules that
are encapsulated inside the macrocycles cavities. These solvents cannot
be removed by high vacuum, not even, when the substances are heated.
They appear in the NMR spectra and are also visible in the crystal structures with quite some disorder. Therefore, elemental compositions were
assessed by isotope pattern analysis of the corresponding ions observed
in the mass spectra.
2-(Azidomethyl)naphthalene:
2-(Bromomethyl)naphthalene
(1.1 g,
5 mmol) and sodium azide (2.6 g, 40 mmol) were heated in dry DMSO
under an argon atmosphere at 45 8C for 17 h. After cooling to RT water
(200 mL) was added and the product was extracted with dichloromethane
(3  100 mL). The combined organic phases were dried with MgSO4, filtered, and the solvent evaporated. The desired compound was obtained
as a solid (0.89 g, 97 %). M.p. 4649 8C; 1H NMR (250 MHz, CDCl3): d =
4.46 (s, 2H; CH2), 7.45- 7.53 (m, 3 H; Ar), 7.817.86 ppm (m, 4 H; Ar);
13
C NMR (100 MHz, CDCl3): d = 55.06, 125.98, 126.46, 126.58, 127.27,
127.86, 128.06, 128.85, 128.86, 125.87, 132.93, 133.17, 133.34 ppm; MS
(EI) (30 8C) m/z (%): 183.2 (66.6) [M]C + , 141.2 (100) [MN3]C + , 154.2
(47.8) [MN2]C + .
Tris-(4-iodophenyl)methanol (2): Concentrated sulfuric acid (4 mL) was
added to parafuchsin hydrochloride (1.77 g, 5.46 mmol) in water (10 mL)
and the mixture was cooled to 0 8C. A solution of sodium nitrite (1.24 g,
18 mmol) in water (4 mL) was added to the solution dropwise and the
stirring was continued for 1 h. Potassium iodide (10 g, 60 mmol) dissolved
in water (10 mL) was then added and the reaction was first let to room
temperature and then kept at 80 8C for 2 h. The dark purple precipitate
formed was filtered, washed with water, and dried. Column chromatography on silica gel by eluting with dichloromethane yielded the pure product (2.80 g, 82 %); Rf = 0.8; 1H NMR (250 MHz, CDCl3): d = 6.88 (d, J =
8.7 Hz, 6 H), 7.54 ppm (d, J = 8.7 Hz, 6 H); 13C NMR (62.5 MHz, CDCl3):
d = 81.6, 94.0, 129.9, 137.6, 145.8 ppm.
4-[Tris(4-iodophenyl)methyl]phenol (3): Sulfuric acid (ten drops) was
added to compound 2 (5 g, 7.84 mmol) and phenol (4 g, 43 mmol) and
the mixture was stirred under argon for 4 h at 80 8C. The melting phenol
served as the solvent; no other solvent was added. After cooling down to
room temperature, an aqueous solution of NaOH (10 %, 50 mL) was
added and the precipitate formed was filtered off. Column chromatography on silica gel by eluting with petroleum ether/ethyl acetate (5:1) yielded the pure product (4.25 g, 76 %). Rf = 0.7; 1H NMR (250 MHz, CDCl3):
d = 6.61 (d, J = 8.8 Hz, 2 H), 6.82 (d, J = 8.8 Hz, 2 H), 6.86 (d, J = 8.6 Hz,
6 H), 7.60 ppm (d, J = 8.6 Hz, 6 H); 13C NMR (62.5 MHz, CDCl3): d =
62.9, 91.6, 114.3, 130.9, 132.2, 134.8, 136.1, 145.6, 155.3 ppm; MS (EI)
(220 8C) m/z (%): 714.3 [M]C + (74), 511.2 [MPhI]C + .
4-(Tris{4-[(trimethylsilyl)ethynyl]phenyl}methyl)phenol (4): Compound 3
(2.3 g, 3.22 mmol), trimethylsilylacetylene (2.75 mL, 19.5 mmol), bis(triphenylphosphine) palladium(II)dichloride (371 mg, 0.53 mmol), and triphenylphosphine (338 mg, 1.29 mmol) were dissolved in dry DMF
(20 mL) and dry triethylamine (90 mL). The reaction mixture was kept at
40 8C for 4 h. After all the solvent was evaporated in vacuo the residue
was purified by column chromatography on a silica column by eluting
with petroleum ether/ethyl acetate (5:1) to give the desired product
(0.8 g, 40 %). Rf = 0.6; 1H NMR (250 MHz, CDCl3): d = 0.00 (s, 27 H;
CH3), 4.77 (br s, 1 H; OH), 6.46 (d, J = 8.8 Hz, 2 H), 6.71 (d, J = 8.8 Hz,
2 H), 6.83 (d, J = 8.6 Hz, 6 H), 7.10 ppm (d, J = 8.6 Hz, 6 H); 13C NMR

Chem. Eur. J. 2012, 18, 1528 1535

4-{Tris[4-(naphthalen-1-ylethynyl)phenyl]methyl}phenol (S1): Dry NEt3


(10 mL) was added to compound 5 (98 mg, 0.24 mmol), 1-iodonaphthalene (198 mg, 0.78 mmol), CuI (28.5 mg), and [PdACHTUNGRE(PPh3)2Cl2] (35 mg) in
DMF (10 mL) and the reaction mixture stirred under an argon atmosphere for 21 h. The solvents were evaporated in vacuo and the residue
was purified on a silica column by eluting with dichloromethane. The
product was obtained after excess iodonaphthalene (first band) as the
second band from the column (105.7 mg, 56 %). Rf = 0.7; 1H NMR
(250 MHz, CDCl3): d = 6.77 (d, 2 H; J = 8.3 Hz, PhHphenol), 7.12 (d, J =
8.3 Hz, 2 H; PhHphenol), 7.30 (d, J = 9 Hz, 6 H; PhH), 7.427.62 (m, 9 H;
Arnaph), 7.747.88 (m, 9 H; Arnaph), 8.43 ppm (d, J = 7.7 Hz, 3 H; PhH);
13
C NMR (100 MHz, CDCl3): d = 64.44, 87.92, 94.16, 114.84, 120.95,
121.35, 125.39, 126.30, 126.54, 126.90, 128.42, 128.89, 130.49, 131.09,
131.15, 132.30, 133.29, 133.36, 137.90, 146.84, 154.20 ppm; MS (EI)
(300 8C) m/z (%): 786.3 (32) [M]C + .
4-(Tris{4-[1-(naphthalen-2-ylmethyl)-1 H-1,2,3-triazol-4-yl]phenyl}methACHTUNGREyl)phenol (S2): 2-(Azidomethyl)naphthalene (143 mg, 0.78 mmol), compound 5 (100 mg, 0.245 mmol), CuSO45 H2O (1.84 mg, 7.35 mmol), and
sodium ascorbate (7.28 mg, 36.8 mmmol) were stirred in THF/H2O (1:1,
10 mL) at RT for 2 d, by which time the reaction mixture became clearer
and the dispersed product mixture separated from the aqueous phase as
a yellow liquid after stirring was terminated. The product was extracted
into dichloromethane (3  20 mL), washed thoroughly with water (2 
50 mL), dried with MgSO4, filtered, and the solvent was evaporated. The
yellow residue was purified on a silica column and eluted first with pure
dichloromethane to remove low-polarity impurities. Then, the eluent was
changed to CH2Cl2/MeOH (1:1) to obtain the desired product (206.4 mg,
88 %). 1H NMR (250 MHz, CDCl3): d = 5.60 (s, 6 H; CH2), 6.55 (d, 2 H;
J = 9 Hz, PhHphenol), 6.87 (d, J = 9 Hz, 2 H; PhHphenol), 7.11 (d, J = 8.3 Hz,
6 H; PhH), 7.24 (d, 3 H; Arnaph), 7.377.43 (m, 6 H; Arnaph), 7.48 (d, J =
8.3 Hz, 6 H; PhH), 7.63 (s, 3 H; CH), 7.65 (s, 3 H; Arnaph), 7.687.74 ppm
(m, 9 H; Arnaph); 13C NMR (62.5 MHz, CDCl3): d = 54.40, 63.9, 120.26,
124.86, 125.16, 126.71, 127.32, 127.72, 127.89, 129.13, 131.38, 131.86,
131.96, 133.16, 133.19, 137.03, 147.04, 147.81 ppm; MS (ESI-TOF) m/z =
958.41 [M+H] + , 980.39 [M+Na] + .
[2]-{1,4-Bis[(4-ACHTUNGRE{tris[4-(naphthalen-1-ylethynyl)phenyl]methyl}phenyloxy)methyl]benzene}{10-tert-butyl-29-pyrenyl-5,17,23,35,38,40,43,45octamethyldispiro[cyclohexane-1,2-7,15,25,33-tetraazaheptacyclo[32.2.2.23, 6.216, 19.221, 24.19, 13.127, 31]hexatetraconta-3,5,9,11,13ACHTUNGRE(44),16,
18,21,23,27,29,31ACHTUNGRE(39),34,36,37,40,42,45-octadecaen-20,1-cyclohexane]-8,14,26,32-tetraon}rotaxane (R1): Dibenzo[18]crown-6 (3.9 mg,
0.011 mmol), pyrene macrocycle W-Py (20.0 mg, 0.017 mmol), K2CO3
(59 mg, 0.43 mmol), 1,4-dibromomethylbenzene (11.0 mg, 0.043 mmol),
and compound S2 (27.2 mg, 0.034 mmol) were stirred in dry dichloromethane (10 mL) for a week. The solvent was then evaporated and the residue was purified on a silica column with dichloromethane. The second
band was collected as the pure rotaxane (39.6 mg, 82 %). 1H NMR
(250 MHz, CDCl3): d = 1.32 (s, 9 H; CCH3), 1.451.70 (br s, 12 H; Cy = cylohexyl), 1.93 (s, 12 H; PhCH3), 1.96 (s, 12 H; PhCH3), 2.32 (br s, 8 H; Cy),
4.44 (s, 4 H; OCH2), 5.93 (s, 4 H; PhHaxle), 6.52 (d, J = 8.3 Hz, 4 H;
PhHstopper), 7.05 (s, 8 H; PhH), 7.10 (d, J = 8.3 Hz, 4 H; PhHstopper), 7.19
8.17 (m, 93 H; ArH), 8.38 ppm (d, 12 H; 2-Naph); 13C NMR (62.5 MHz,

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemeurj.org

1533

P. Ceroni, C. Schalley et al.

CDCl3): d = 14.19, 18.82, 22.77, 23.10, 26.38, 29.43, 29.74, 31.22, 31.99,
35.41, 35.78, 45.49, 64.38, 71.07, 88.23, 93.87, 113.84, 120. 81, 121.71,
124.86, 125.35, 126.22, 126.53, 126.90, 127.02, 127.07, 127.84, 127.86,
128.10, 128.25, 128.35, 128.40, 128.46, 128.96, 129.29, 130.54, 130.60,
130.95, 131.10, 131.24, 131.34, 131.59, 132.65, 133.28, 133.32, 133.97,
134.71, 135.01, 135.15, 135.18, 135.68, 140.14, 146.23, 156.23, 165.37,
165.85 ppm; MS (ESI-FTICR) m/z (%): 1520.75 (22) [M+2 HNEt3]2 + ,
2939.34 (18) [M+HNEt3] + .
1,4-Bis{[4-ACHTUNGRE(tris{4-[1-(naphthalen-2-ylmethyl)-1H-1,2,3-triazol-4-yl]phenyl}methyl)phenoxy]methyl}benzene (A1): The axle was obtained as a
byproduct (5 %) of the rotaxane synthesis. Due to the low available
amount, it was characterized only by 1H NMR spectroscopy. Rf = 0.9;
1
H NMR (250 MHz, CDCl3): d = 4.96 (s, 4 H; CH2), 6.81 (d, J = 7.0 Hz,
2 H; PhH), 7.06 (d, J = 7.0 Hz, 2 H; PhH), 7.18 (d, J = 6.8 Hz, 6 H; PhH),
7.46 (d, J = 6.8 Hz, 6 H; PhH), 7.317.74 (m, 36 H; Naph), 8.31 ppm (d,
J = 6.5 Hz, 6 H; 2-Naph).
[2]-(1,4-Bis{[4-ACHTUNGRE(tris{4-[1-(naphthalen-2-ylmethyl)-1H-1,2,3-triazol-4-yl]phenyl}methyl)phenoxy]methyl}benzene)-{10-tert-butyl-29-pyrenyl5,17,23,35,38,40,43,45-octamethyldispiro[cyclohexane-1,2-7,15,
25,33-tetraazaheptacyclo[32.2.2.23, 6.216, 19.221, 24.19, 13.127, 31]hexatetraACHTUNGREconta-3,5,9,11,13ACHTUNGRE(44),16,18,21,23,27,29,31ACHTUNGRE(39),34,36,37,40,42,45octaACHTUNGREdecaen-20,1-cyclohexan]-8,14,26,32-tetraon}rotaxane (R2): Dibenzo[18]crown-6 (3.9 mg, 0.011 mmol), pyrene macrocycle W-Py (50 mg,
0.043 mmol), K2CO3 (59 mg, 0.43 mmol), 1,4-dibromomethylbenzene
(11.0 mg, 0.043 mmol), and compound S2 (82 mg, 0.086 mmol) were
stirred in dry dichloromethane (10 mL) for one week. The solvent was
then evaporated and the residue was purified on a silica column with dichloromethane/methanol (20:1). The first band was collected as the pure
rotaxane (116.2 mg, 85 %). 1H NMR (250 MHz, CDCl3/CD3OD = 5:1):
d = 1.11 (s, 9 H; CCH3), 1.401.52 (br s, 12 H; Cy), 1.77 (s, 24 H; PhCH3),
2.11 (br s, 8 H; Cy), 4.20 (s, 4 H; OCH2), 5.57 (s, 12 H; CH2), 5.91 (s, 4 H;
PhHaxle), 6.34 (d, J = 8.3 Hz, 4 H; PhHstopper), 6.96 (s, 8 H; PhH), 7.00 (d,
J = 8.3 Hz, 4 H; PhHstopper), 7.118.02 (m, 99 H; ArH), 7.92 (s, 2 H; 4,6isophth), 8.11 ppm (s, 1 H; 2-isophth); 13C NMR (62.5 MHz, CDCl3) d =
14.04, 18.65, 22.66, 22.93, 29.33, 29.65, 30.95, 31.91, 35.15, 45.30, 53.47,
54.42, 63.93, 70.43, 113.67, 120.09, 124.73, 124.97, 125.06, 125.13, 125.17,
125.19, 126.73, 127.31, 127.34, 127.74, 127.76, 127.83, 127.89, 127.91,
127.95, 128.18, 128.23, 129.02, 129.17, 131.02, 131.10, 131.28, 131.33,
131.38, 131.83, 131.89, 132.39, 133.17, 133.19, 133.21, 134.23, 134.87,
135.06, 135.77, 140.22, 146.35, 146.82, 147.67, 147.86, 166.35, 167.07 ppm;
MS (ESI-FTICR) m/z (%): 1691.71 (100) [M+2 HNEt3]2 + .
Photophysical experiments: The photophysical experiments were performed in dichloromethane/acetonitrile (1:1, v/v) under air-equilibrated
conditions at 298 K and in a dichloromethane/chloroform (1:1, v/v) rigid
matrix at 77 K. UV/Vis absorption spectra were recorded with a Perkin
Elmer l40 spectrophotometer. Fluorescence spectra were obtained with
a PerkinElmer LS-50 spectrofluorimeter, equipped with a Hamamatsu
R928 phototube for the UV/Vis range, the same instrument was used to
record phosphorescence lifetimes. The emission quantum yields of W-Py,
Py, A1, S1, R1, and R2 were determined by the method of Demas and
Crosby by using anthracene in ethanol as a reference.[20] In the case of
Naph and S2 naphthalene in cyclohexane[20] was used as a standard to determine the emission quantum yield. Fluorescence lifetime measurements
were performed by using an Edinburgh FLS920 spectrofluorimeter
equipped with a TCC900 card for data acquisition in time-correlated
single-photon counting experiments (0.5 ns time resolution) with a D2
lamp.
The estimated experimental errors are: 2 nm on the band maximum, 5 %
on the molar absorption coefficient, on the emission quantum yield and
the lifetimes.

Acknowledgements
This work has been supported in Italy by Fondazione Carisbo (Dispositivi nanometrici basati su dendrimeri e nanoparticelle), bilateral project
ItalyChina 2011 (MAE DGPCC), and MIUR (PRIN 20085ZXFEE).

1534

www.chemeurj.org

We are also grateful to the Deutsche Forschungsgemeinschaft for financial support (SFB 765 Multivalency).

[1] For reviews, see: a) S. Bonnet, J.-P. Collin, Chem. Soc. Rev. 2008, 37,
1207 1217; b) B. Champin, P. Mobian, J.-P. Sauvage, Chem. Soc.
Rev. 2007, 36, 358 366; c) C. A. Schalley, T. Weilandt, J. Brggemann, F. Vgtle, Top. Curr. Chem. 2004, 248, 141 200; d) M.
Blanco, M. Consuelo Jimnez, J.-C. Chambron, V. Heitz, M. Linke,
J.-P. Sauvage, Chem. Soc. Rev. 1999, 28, 293 305; e) F. M. Raymo,
J. F. Stoddart, Chem. Rev. 1999, 99, 1643 1663; f) J.-P. Sauvage, Acc.
Chem. Res. 1998, 31, 611 619; g) S. A. Nepogodiev, J. F. Stoddart,
Chem. Rev. 1998, 98, 1959 1976; h) R. Jger, F. Vgtle, Angew.
Chem. 1997, 109, 966 980; Angew. Chem. Int. Ed. Engl. 1997, 36,
930 944; i) R. Hoss, F. Vgtle, Angew. Chem. 1994, 106, 389 398;
Angew. Chem. Int. Ed. Engl. 1994, 33, 375 384; j) S. Anderson,
H. L. Anderson, J. K. M. Sanders, Acc. Chem. Res. 1993, 26, 469
475; k) J.-P. Sauvage, Acc. Chem. Res. 1990, 23, 319 327; l) C. O.
Dietrich-Buchecker, J.-P. Sauvage, Chem. Rev. 1987, 87, 795 810.
[2] a) P. Ceroni, A. Credi, M. Venturi, V. Balzani, Photochem. Photobiol. Sci. 2010, 9, 1561 1573; b) S. Durot, F. Reviriego, J.-P. Sauvage,
Dalton Trans. 2010, 39, 10557 10570; c) P. Bodis, M. R. Panman,
B. H. Bakker, A. Mateo-Alonso, M. Prato, W. J. Buma, A. M.
Brouwer, E. R. Kay, D. A. Leigh, S. Woutersen, Acc. Chem. Res.
2009, 42, 1462 1469; d) S. Silvi, M. Venturi, A. Credi, J. Mater.
Chem. 2009, 19, 2279 2294; e) V. Balzani, A. Credi, M. Venturi,
Molecular Devices and MachinesConcepts and Perspectives for the
Nanoworld, Wiley-VCH, Weinheim, 2008; f) E. R. Kay, D. A. Leigh,
F. Zerbetto, Angew. Chem. 2007, 119, 72 196; Angew. Chem. Int.
Ed. 2007, 46, 72 191; g) A. Mateo-Alonso, D. M. Guldi, F. Paolucci,
M. Prato, Angew. Chem. 2007, 119, 8266 8272; Angew. Chem. Int.
Ed. 2007, 46, 8120 8126; h) A. Credi, Angew. Chem. 2007, 119,
5568 5572; Angew. Chem. Int. Ed. 2007, 46, 5472 5475; i) V. Balzani, A. Credi, S. Silvi, M. Venturi, Chem. Soc. Rev. 2006, 35, 1135
1149; j) M. Clemente-Le n, A. Credi, M.-V. Mart
nez-D
az, C. Mingotaud, J. F. Stoddart, Adv. Mater. 2006, 18, 1291 1296; k) S.
Bonnet, J.-P. Collin, M. Koizumi, P. Mobian, J.-P. Sauvage, Adv.
Mater. 2006, 18, 1239 1250; l) W. R. Browne, B. L. Feringa, Nat.
Nanotechnol. 2006, 1, 25 35; m) A. R. Pease, J. O. Jeppesen, J. F.
Stoddart, Y. Luo, C. P. Collier, J. R. Heath, Acc. Chem. Res. 2001,
34, 433 444; n) R. Ballardini, V. Balzani, A. Credi, M. T. Gandolfi,
M. Venturi, Acc. Chem. Res. 2001, 34, 445 455; o) A. Harada, Acc.
Chem. Res. 2001, 34, 456 464; p) C. A. Schalley, K. Beizai, F.
Vgtle, Acc. Chem. Res. 2001, 34, 465 476; q) J.-P. Collin, C. Dietrich-Buchecker, P. Gavia, M. Consuelo Jimenez-Molero, J.-P.
Sauvage, Acc. Chem. Res. 2001, 34, 477 487; r) V. Balzani, A. Credi,
F. M. Raymo, J. F. Stoddart, Angew. Chem. 2000, 112, 3484 3530;
Angew. Chem. Int. Ed. 2000, 39, 3348 3391.
[3] a) J.-Y. Wang, J.-M. Han, J. Yan, Y. Ma, J. Pei, Chem. Eur. J. 2009,
15, 3585 3594; b) H. Onagi, J. Rebek, Jr., Chem. Commun. 2005,
4604 4606.
[4] For reviews, see for example: a) V. Balzani, P. Ceroni, M. Maestri,
V. Vicinelli Curr. Opin. Chem. Biol. 2003, 7, 657; b) F. Wrthner,
T. E. Kaiser, C. R. Saha-Mller, Angew. Chem. Int. Ed. 2011, 50,
3376; c) S. V. Eliseevaa, J.-C. G. Bnzli, Chem. Soc. Rev. 2010, 39,
189 227; d) M. R. Wasielewski, Acc. Chem. Res. 2009, 42, 1910
1921; e) N. Aratani, D. Kim, A. Osuka, Acc. Chem. Res. 2009, 42,
1922 1934; f) F. Scandola, C. Chiorboli, A. Prodi, E. Iengo, E. Alessio, Coord. Chem. Rev. 2006, 250, 1471 1496.
[5] For some recent examples, see for example: a) A. Uetomo, M.
Kozaki, S. Suzuki, K. Yamanaka, O. Ito, K. Okada, J. Am. Chem.
Soc. 2011, 133, 13276 13279; b) B. Branchi, G. Bergamini, L. Fiandro, P. Ceroni, F. Vgtle, F.-G. Klrner, Chem. Commun. 2010, 46,
3571 3573; c) D. G. Kuroda, C. P. Singh, Z. Peng, V. D. Kleiman,
Science 2009, 326, 263 267; d) C. Giansante, P. Ceroni, V. Balzani,
F. Vgtle, Angew. Chem. 2008, 120, 5502 5505; Angew. Chem. Int.
Ed. 2008, 47, 5422 5425; e) J.-L. Wang, J. Yan, Z.-M. Tang, Q. Xiao,
Y. Ma, J. Pei, J. Am. Chem. Soc. 2008, 130, 9952 9962.

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eur. J. 2012, 18, 1528 1535

Light-Harvesting in Rotaxanes

FULL PAPER

[6] a) V. Balzani, G. Bergamini, P. Ceroni, E. Marchi, New J. Chem.


2011, 35, 1944; b) A.-M. Caminade, A. Hameau, J.-P. Majoral,
Chem. Eur. J. 2009, 15, 9270 9285.
[7] a) D. A. Tramontozzi, N. D. Suhan, S. H. Eichhorn, S. J. Loeb, Chem.
Eur. J. 2010, 16, 4466 4476; b) A. M. Elizarov, T. Chang, S.-H. Chiu,
J. F. Stoddart, Org. Lett. 2002, 4, 3565 3568.
[8] A. M. Elizarov, S.-H. Chiu, P. T. Glink, J. F. Stoddart, Org. Lett.
2002, 4, 679 682.
[9] a) Y. Zeng, Y. Li, M. Li, G. Yang, Y. Li, J. Am. Chem. Soc. 2009,
131, 9100 9106; b) S.-Y. Kim, Y. H. Ko, J. W. Lee, S. Sakamoto, K.
Yamaguchi, K. Kim, Chem. Asian J. 2007, 2, 747 754.
[10] a) C. A. Hunter, J. Chem. Soc. Chem. Commun. 1991, 749 751;
b) C. A. Hunter, J. Am. Chem. Soc. 1992, 114, 5303 5311; c) F.
Vgtle, S. Meier, R. Hoss, Angew. Chem. 1992, 104, 1628 1631;
Angew. Chem. Int. Ed. Engl. 1992, 31, 1619 1622; d) S. Ottens-Hildebrandt, S. Meier, W. Schmidt, F. Vgtle, Angew. Chem. 1994, 106,
1818 1821; Angew. Chem. Int. Ed. Engl. 1994, 33, 1767 1770.
[11] B. Baytekin, S. S. Zhu, B. Brusilowskij, J. Illigen, J. Ranta, J. Huskonen, L. Russo, K. Rissanen, L Kauffman, C. A. Schalley, Chem. Eur.
J. 2008, 14, 10012 10028.
[12] S. Sengupta, S. K. Sadhukhan, S. Muhuri, Synlett 2003, 2329 2332.
[13] a) K. Sonogashira, Y. Tohda, N. Hagihara, Tetrahedron Lett. 1975,
16, 4467 4470; b) R. Chinchilla, C. Najera, Chem. Rev. 2007, 107,
874 922.
[14] a) H. C. Kolb, M. G. Finn, K. B. Sharpless, Angew. Chem. 2001, 113,
2056 2075; Angew. Chem. Int. Ed. 2001, 40, 2004 2021; b) C. W.
Torne, C. Christensen, M. Meldal, J. Org. Chem. 2002, 67, 3057
3064.

Chem. Eur. J. 2012, 18, 1528 1535

[15] a) G. M. Hbner, J. Glser, C. Seel, F. Vgtle, Angew. Chem. 1999,


111, 395 398; Angew. Chem. Int. Ed. 1999, 38, 383 386; b) G. M.
Hbner, G. Nachtsheim, Q. Y. Li, C. Seel, F. Vgtle, Angew. Chem.
2000, 112, 1315 1318; Angew. Chem. Int. Ed. 2000, 39, 1269 1272.
Also, see: c) P. Ghosh, G. Federwisch, M. Kogej, C. A. Schalley, D.
Haase, W. Saak, A. Ltzen, R. A. Gschwind, Org. Biomol. Chem.
2005, 3, 2691 2700; d) P. Ghosh, O. Mermagen, C. A. Schalley,
Chem. Commun. 2002, 2628 2629; e) A. Affeld, G. M. Hbner, C.
Seel, C. A. Schalley, Eur. J. Org. Chem. 2001, 2877 2890.
[16] N. D. Marsh, C. J. Mikolajczak, M. J. Wornat, Spectrochim. Acta Part
A 2000, 56, 1499 1511.
[17] See, for example: a) L. Rodr
guez, E. Delgado-Pinar, A. SornosaTen, J. Alarcon, E. Garcia-Espana, M. Cano, J. C. Lima, F. Pina, J.
Phys. Chem. B 2009, 113, 15 455 15 459; b) C. Saudan, V. Balzani, P.
Ceroni, M. Gorka, M. Maestri, V. Vicinelli, F. Vgtle, Tetrahedron
2003, 59, 3845 3852.
[18] As previously discussed, the molar absorption coefficient of A1 was
assumed to be double of that of S1.
[19] a) H. Meyer, Monatsh. Chem. 1901, 22, 415 442; b) R. Jger, T.
Schmidt, D. Karbach, F. Vgtle, Synlett 1996, 723 725; c) R. Jger,
Ph.D. Thesis, Universitt Bonn (Germany), 1997.
[20] M. Montalti, A. Credi, L. Prodi, M. T. Gandolfi, Handbook of Photochemistry, 3rd ed., CRC Press, Boca Raton, 2006, Chapter 7.

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Received: September 22, 2011


Published online: December 28, 2011

www.chemeurj.org

1535

Você também pode gostar