You are on page 1of 12

Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

View Article Online / Journal Homepage / Table of Contents for this issue

Showcasing the laboratory of Will Skene at the


Photophysical Characterization of Conjugated As featured in:
Materials Laboratory, Université de Montréal.
Title: Solvatochromic investigation of highly fluorescent
2-aminobithiophene derivatives

The fluorescence of 2-aminobithiophenes was turned-on when


substituted with electron withdrawing groups in the 2’-position.
These push-pull bithiophenes fluoresced in near quantitative yields
regardless of the solvent polarity. Meanwhile, their emission and
colour could be tailored contingent on solvent polarity.

See Skene, Phys. Chem. Chem.


Phys., 2012, 14, 6946.

www.rsc.org/pccp
Registered Charity Number 207890
View Article Online
PCCP Dynamic Article Links

Cite this: Phys. Chem. Chem. Phys., 2012, 14, 6946–6956

www.rsc.org/pccp PAPER
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

Solvatochromic investigation of highly fluorescent 2-aminobithiophene


derivativesw
Andréanne Bolduc, Yanmei Dong, Amélie Guérinz and W. G. Skene*
Received 30th January 2012, Accepted 9th March 2012
DOI: 10.1039/c2cp40293a

The solvatochromic and electrochemical properties of electronic push–pull 2-aminobithiophenes


consisting of an aldehyde and nitro withdrawing groups were examined. With the use of an
integrating sphere, the absolute quantum yields of the bithiophenes were measured. They were
found to be highly fluorescent (Ffl > 70%), provided the nitro group was not located in the
4 0 -position. High fluorescence yields were observed regardless of solvent, except for alcohols,
notably methanol and ethanol. Cryofluorescence was used to probe the bithiophene temperature
dependent excited state deactivation modes. The singlet excited state deactivation mode other
than fluorescence was found to be internal conversion involving rotation around the
thiophene–thiophene bond. Deactivation by intersystem crossing to the triplet state occurred in
ca. 40% only for the unsubstituted 2-aminobithiophene. In contrast, the fluorescence was
quenched by photoinduced intramolecular electron transfer when the nitro group was located in
the 4 0 -position of the bithiophene. Both the absorbance and fluorescence of the bithiophenes were
found to be solvatochromic with more pronounced solvent dependent shifts being observed with
the fluorescence. In fact, both the fluorescence and Stokes shifts were linearly dependent on the
ET(30) solvent parameter. Deviations from the linear trend of the Stokes shift with ET(30) were
observed in ethanol and methanol as a result of intermolecular hydrogen abstraction from the
solvent and by the excited nitro group. The oxidation potential of the bithiophenes was also
highly dependent on the type and number of the electron withdrawing substituents, with values
ranging between 0.8 and 1.2 V vs. SCE.

Introduction Among the interesting opto-electronic properties of conjugated


thiophenes is their fluorescence. However, high fluorescence
Conjugated materials have received much attention because of quantum yields (Ffl) are only possible with thiophenes having
their opto-electronic properties that are compatible for use in degrees of oligomerization greater than 6.5 Intersystem crossing
plastic electronics. This is particularly true for thiophene (ISC) to the triplet state is the major deactivation mode of the
derivatives whose low oxidation states can easily be accessed excited singlet manifold for oligothiophenes consisting of six
by chemical doping. Also, their doped states are highly stable thiophene repeating units or less. This efficient pathway results
and can be reversibly generated. Meanwhile, the properties of in their quenched fluorescence.6–8 While thiophenes are interesting
conjugated thiophenes have resulted in their use in a wide for fluorescence sensors, the high degree of conjugation required
range of applications including conductive layers in tactile to take advantage of their high fluorescence often limits their use
applications,1 photoactive layers in photovoltaics,2,3 and in because of solubility problems. Fluorescent bithiophene derivatives
anti-static packaging.4 would be advantageous over their oligomeric or polymeric
counterparts because of their higher solubility in a wide range
of solvents, provided they fluoresce in appreciable amounts.
Laboratoire de caractérisation photophysique des mate´riaux conjugués,
De´partement de Chimie, Universite´ de Montréal, CP 6128, The inherent and efficient deactivation of bithiophenes by ISC
Centre-ville, Montreal, QC, Canada. E-mail: w.skene@umontreal.ca must therefore be suppressed to make bithiophenes fluorescent.
w Electronic supplementary information (ESI) available: 1H and 13C This is possible by incorporating electronic groups that perturb
NMR of all new compounds, absorbance, fluorescence, and cryofluores-
cence spectra, Lippert–Mataga and ET(30) plots, cyclic voltammograms, the singlet–triplet energy gap. We recently demonstrated that
laser flash photolysis results, and additional fluorescence quenching data. incorporating complementary electron withdrawing and donating
CCDC reference number 865250. For ESI and crystallographic data in groups in conjugation led to highly fluorescent bithiophenes.9
CIF or other electronic format see DOI: 10.1039/c2cp40293a Interestingly, the 2-aminoderivatives 2 and 3 (Chart 1) fluoresced
z Current address: ENSCBP-École Nationale Supérieure de Chimie,
Biologie et de Physique, Institut Polytechnique de Bordeaux, 16, in near unity. This is in contrast to their unsubstituted
avenue Pey Berland, 33607 Pessac Cedex, France. bithiophene counterpart that fluoresced in less than 2%.5

6946 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 This journal is c the Owner Societies 2012
View Article Online

Spectroscopic measurements
Absorption measurements were done on a Cary 500 spectro-
photometer and fluorescence measurements were performed
on an Edinburgh Instruments FLS-920 fluorimeter after
deaerating the solvent for 20 minutes. Absolute quantum
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

yields were measured using an integrating sphere system from


Edinburgh Instruments.

Cryofluorescence
Cryofluorescence was performed using cryocuvettes from
NSG Precision Glass with an OptistatDN cryostat from
Oxford Instruments. The measurements were done in anhydrous
and deaerated tetrahydrofuran. The change in tetrahydrofuran’s
refractive index (nxK) with temperature was corrected using
eqn (1) according to:19
nxK = 1.55868  5.14857  104 K1 (1)
The temperature dependent fluorescence quantum yields (FxK)
were corrected using eqn (2), which takes into account the
temperature dependent change in the refractive index relative to
the absolute quantum yield at room temperature (F300K):20,21
Chart 1 Bithiophenes investigated and the bithiophene numbering  
scheme. nxK 2
FxK ¼ F300K ð2Þ
n300K
The bithiophenes are further chemically and photochemically inert,
owing to the electron accepting substituent in the 3-position.10–12 Cyclic voltammetry
An additional advantage of the complementary electronic
push–pull bithiophenes is their photophysical properties, Cyclic voltammetry measurements were made on a Bioanalytical
which are found to be sensitive to changes in their local System potentiostat. Compounds were dissolved in deaerated
environment, especially fluctuations in polarity.13,14 These dichloromethane at 104 M with 0.5 M NBu4PF6. A platinum
photophysical properties combined with their chemical and electrode was used as the working electrode with a platinum wire
photochemical stability make them interesting candidates for as the auxiliary electrode. The reference electrode was a silver
solvatochromic probes for potential biochemical monitoring wire electrode. Ferrocene was added to the solution as an internal
such as protein dynamics and changes in the protein reference for calibrating the measured oxidation potentials to
conformation.15–18 SCE using Epa = 435 mV vs. SCE for ferrocene.22
Given the suitability of aminothiophenes as solvatochromic Laser flash photolysis
fluorescence probes, assessing their solvatochromism is of
importance for determining their limitation in such applications. Laser flash photolysis measurements were done with a Luzchem
Also, solvatochromic effects within the visible region of the mini-LFP system excited at 355 nm with the third harmonic of a
spectrum would be advantageous for polarity probe applications. Nd-YAG laser. The solutions were prepared with an absorbance
Large solvent induced absorbance and fluorescence shifts of between 0.3 and 0.4 at 355 nm. The transient absorbance spectra
the conjugated push–pull systems should be possible with were generated by averaging the maximum absorbance over
solvents of varying polarity. The absorbance and fluorescence 5–7 shots per wavelength that were recorded at different
solvatochromism of the electronic push–pull aminothiophenes intervals after the laser pulse. Quenching kinetics were done
1–6 were therefore investigated. The effect of the placement by measuring the change in the first order rate constant as a
of the strong electron withdrawing nitro group on the opto- function of the quencher added to the sample. The quantum
electronic properties is additionally described. Such studies yield of ISC (FISC) of 1, which is equal to the quantum yield of
are important for providing insight into the structural and triplet formation (FTT), was determined by relative actinometry
electronic requirements for making new and highly fluorescent using benzophenone (FTT = 1).23,24 Optically matched samples
thiophenes with low degrees of oligomerization, especially of benzophenone and 1 at 355 nm were prepared in anhydrous
functionalized bithiophenes. and deaerated dichloromethane. Equimolar amounts of
2-methylnaphthalene were added to both samples to quench
95% of the triplets of benzophenone and 1 by energy transfer
Experimental section within the laser pulse. The required concentration of 2-methyl-
naphthalene was determined from 20t1 1
o kq , where to is the
General procedures
triplet lifetime of either benzophenone or 1 in the absence of
All reagents were commercially available from Aldrich unless 2-methylnaphthalene, and kq is their corresponding bimolecular
otherwise stated. Anhydrous and deaerated solvents were quenching rate constant with 2-methylnaphthalene determined
obtained from a Glass Contour solvent purification system. from Fig. S53 (ESIw). The resulting triplet 2-methylnaphthalene

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 6947
View Article Online

produced by energy transfer from benzophenone and 1 was Ethyl 5-amino-3 0 ,5 0 -dinitro-[2,2 0 -bithiophene]-4-carboxylate
monitored at 420 nm. The maximum Dabsorbance at 420 nm (5). To trifluoroacetic anhydride (5 mL) at 0 1C was added
was examined as a function of laser power. The FTT of 1 was fuming nitric acid (264 mg, 4 mmol). The mixture was stirred
calculated from the ratio of the resulting slopes as per Fig. 8. for 0.5 h at 0 1C. Meanwhile, 1 (0.506 g, 2 mmol) was dissolved
in anhydrous dichloromethane (10 mL) and trifluoroacetic
Crystal structure determination anhydride (3 mL) was added dropwise at room temperature.
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

Suitable monocrystals of 5 for X-ray diffraction were obtained The mixture was stirred for 0.5 h and then it was added
by the slow solvent evaporation of hexane and ethyl acetate. dropwise into a flask of fuming nitric acid in trifluoroacetic
Diffraction data were collected on a Bruker FR591 diffractometer anhydride at 0 1C. The final reaction mixture was stirred for
using a graphite-monochromatized CuKa source with 1.54178 Å. 1 h at 0 1C, poured into ice water (20 mL), and extracted with
The structure was solved by direct methods (SHELXS97). All dichloromethane. The organic layer was separated and the
non-hydrogen atoms were refined using the Fobs2 (SHELXS97) aqueous layer was extracted with dichloromethane (20 mL  3).
method. Hydrogen atoms were refined at the calculated positions The combined organic layers were washed with water and then
with fixed isotropic U using the riding model techniques (Table 1). mixed with aqueous potassium carbonate (20%, 20 mL) and
stirred for 1 h at room temperature. The organic layer was
Synthesis separated, washed with brine, dried over Na2SO4, and then
1, 2, 3 and 6 were prepared as previously reported.9 filtered. The solvent was then evaporated and product 5 was
obtained as a dark red solid (0.45 g, 66%) after purifying it over a
Ethyl 5-amino-3 0 -nitro-[2,2 0 -bithiophene]-4-carboxylate (4). short column of silica gel. Mp 215–217 1C. 1H NMR (400 MHz,
In anhydrous dichloromethane (10 mL) was dissolved 1 acetone-d6) d = 8.37 (s, 1H), 7.80 (s, 1H), 7.89 (bs, 2H), 4.32
(0.506 g, 2 mmol) and trifluoroacetic anhydride (3 mL) was (quart, J = 7.1 Hz, 2H), 1.35 (t, J = 7.1 Hz, 3H). 13C NMR
added dropwise at room temperature. The mixture was stirred (100 MHz, acetone-d6) d = 170.6, 165.8, 146.9, 144.2, 138.8,
for 0.5 h and nitric acid (20% aqueous, 14 mL) was added 137.0, 128.2, 111.5, 109.8, 61.9, 15.7. HR-MS (ESI): calcd for
dropwise with vigorous stirring at 0 1C. The mixture was stirred C11H9N3O6S2 (M+H)+: 344.0005, found: 344.0003.
for 1 h at 0 1C and extracted with dichloromethane. The organic
layer was separated and the aqueous layer was extracted with
dichloromethane (3  20 mL). The combined organic layers were Results and discussion
washed with water and then mixed with aqueous potassium
Crystallography
carbonate (20%, 20 mL) and stirred for 1 h at room temperature.
The organic layer was separated, washed with brine, dried by While mass spectrometry and NMR analyses confirmed the
Na2SO4, and filtered. After the solvent was evaporated, product 4 identity of products 1–6, assigning the absolute regioisomer of
was isolated as a dark red solid (0.17 g, 29%) by purifying it over the nitro substituents of 5 is not straightforward. Absolute
a short column of silica gel. Mp 179–181 1C. 1H NMR (400 MHz, assignment of the regioisomer could however be found by
acetone-d6/D2O) d = 7.59 (d, J = 5.8 Hz, 1H), 7.59 (s, 1H), 7.46 X-ray diffraction (XRD). Crystals suitable for X-ray diffraction
(d, J = 5.8 Hz, 1H), 4.30 (q, J = 7.1 Hz, 2H), 1.34 (t, J = 7.1 Hz, were therefore grown from the slow evaporation of hexane and
3H). 13C NMR (100 MHz, acetone-d6) d = 168.2, 166.3, 142.1, ethyl acetate. As seen in the top panel of Fig. 1, the resolved
141.6, 132.4, 126.8, 124.6, 112.7, 107.7, 61.5, 15.6. HR-MS (ESI): structure confirms the 2 0 ,40 -dinitro substitution of 5. The
calcd for C11H10N2O4S2 (M+H)+: 299.0155, found: 299.0157. resolved structure further shows that the two thiophenes adopt
an anti-parallel configuration. This is in contrast to 2, whose
Table 1 Details of crystal structure determination of 5 thiophenes adopt an atypical syn orientation.9
Interestingly, the two thiophenes of 5 are coplanar, as
Formula C11H9N3O6S2
shown in the lower panel of Fig. 1. This is in part responsible
1
Mw (g mol ); F(000) 343.33 g mol1; 352 for its high degree of conjugation. The coplanarity of the
Crystal colour and form Blue needle bithiophenes is quantitatively supported by the mean planes
Crystal size/mm 0.25  0.06  0.04
T/K; dcalcd/g cm3 200 (2); 1.715 described by the two thiophenes that are twisted by 2.41(1)1.
Crystal system Triclinic Similarly, the mean plane angle of the thiophene and the terminal
Space group P1% amine is twisted by only 0.59(1)1. A small twisting angle, 1.05(1)1,
Unit cell: a/Å 5.0101(1) was also observed for the mean planes of the thiophene and the
b/Å 10.6411(2)
c/(Å) 13.3104(2) terminal nitro group. The coplanarity of the terminal groups and
a/1 73.901(1) the bithiophene implies a high degree of conjugation.
b/1 79.183(1) The organisation of the molecules in the crystal lattice is
g/1 80.485(1)
V/Å3; Z 664.83(2); 2 driven by hydrogen bonding. Two hydrogen bonds were found
y range (1); completeness 3.50 to 72.47; 0.937 to occur between the amine and the ester oxygen of two
Reflections: collected/independent; Rint 8772/2538; 0.033 different molecules at a distance of 2.829(5) and 3.077(2) Å.
m/mm1 3.994 The resulting hydrogen bonded dimer is seen in Fig. 2. Hydrogen
Abs. corr. Semi-empirical
R1(F); wR(F2) [I > 2s(I)] 0.0433; 0.1139 bonding further occurs between the terminal amine and an
R1(F); wR(F2) (all data) 0.0460; 0.1168 oxygen of the nitro group with a distance of 3.125(2) Å. It was
GoF(F2) 1.014 further found that two molecules crystallize per lattice and the
Max. residual e density 0.368 e Å3
molecules crystallize in a ladder-type lattice, as shown in Fig. 3.

6948 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 This journal is c the Owner Societies 2012
View Article Online

Absorbance and fluorescence


The extensive photophysical properties of 1–6 were analyzed
given that only the cursory spectroscopic properties of 1, 2 and
6 were previously examined.9 The absorbance and fluorescence
of the compounds were first examined in dichloromethane.
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

This was chosen as a suitable reference solvent for examining


the effect of the different substituents on the photophysical
properties owing to the high solubility of the compounds in
this solvent. The effect of the nitro group and its placement on
the bithiophene on the spectroscopic properties is possibly
studied by comparing the photophysical properties of 3–5 to
the unsubstituted benchmark aminothiophene 1. As seen in
Table 2, both the absorbance and fluorescence maxima of 2–5
are bathochromically shifted relative to 1. This is a result of the
electronic push–pull effect of the electron donor and acceptor
groups that are conjugated, resulting in an intramolecular charge
transfer. The spectroscopic shifts are more pronounced for 3 and
5 than for 2, owing to the inherent stronger electron withdrawing
capacity of the nitro group relative to the aldehyde. The similar
absorbance and fluorescence maxima for 3 and 5 imply that the
additional nitro group of 5 does not enhance the charge transfer
that is developed between the donor and acceptor moieties. This
is in contrast to 4, whose absorbance lies between that of 2 and 3.
Fig. 1 Face view for the resolved XRD structure of 5 (top) and edge Interestingly, its fluorescence is the most red shifted for the
view showing the coplanarity of the thiophenes (bottom). bithiophene series studied. The strong red shift implies a high
stabilization of the excited state (vide infra), most likely from
excited state intramolecular charge transfer between the amine
and the nitro substituents, which are in close proximity, unlike
with 3. The absorbance and fluorescence are also sensitive to
substitution on the amine. This is evident from the absorbance
and fluorescence spectra of 6 that are bathochromically shifted
from 2. Interestingly, the Stokes shift of 2 is larger than 6. This
implies that the excited state of 2 is more polar than 6, owing to
the stronger donating character of the primary amine over that
of the secondary amine.
While the effect of the substitution on both the absorbance
and fluorescence is collectively summarized in Table 2, it can
also visually be seen in the photographs in Fig. 4. Most
striking is the colour change of both the absorbance and
fluorescence of the nitro derivatives 3 and 4 relative to 2.
The effect of bithiophene substitution on the fluorescence yield
also can be clearly seen in the photographs. It is evident that 2,
Fig. 2 Figure showing the intermolecular hydrogen bonding between 3, and 6 fluoresce while the fluorescence of 1, 4, and 5 is
the molecules in the lattice for 5. quenched. This trend was confirmed by measuring their
absolute quantum yields (Ffl) that were obtained with an
integrating sphere. This method is preferred over relative
actinometry because a reference having a similar absorbance,
fluorescence, and quantum yield to the compound of study is
not required. The Ffl values determined with an integrating
sphere are absolute values with much higher precision than the
values determined by relative actinometry. As seen in Table 2,
the push–pull bithiophenes fluoresce in excess of 70%. This is
in contrast to 1, 4, and 5 whose Ffl are less than 6%. The
quenched fluorescence of 4 and 5 is clearly a result of the nitro
substituent in the 4 0 -position, especially given that 3 strongly
fluoresces. Meanwhile, the quenched fluorescence observed for
1 suggests that an electron donor–acceptor configuration is
Fig. 3 Crystal lattice packing of 5 shown along the c axis. required to suppress the inherent fluorescence deactivation.

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 6949
View Article Online

Table 2 Spectroscopic and electrochemical data for 1–6 in dichloromethane

labs/nm lem/nm Stokes shift/cm1 Ffla (180 K) Egspec b/eV Epac/V Epcc/V HOMOd/eV LUMOd/eV Egelec e/eV
1 330 427 6890 0.06 (0.38) 3.1 0.79 1.21 5.1 3.1 2.0
2 406 498 4550 0.83 (1.0) 2.6 1.07 1.13 5.4 3.2 2.2
3 494 600 5170 0.76 (1.0) 2.2 1.15 0.83/1.11 5.5 3.5 2.0
4 450 615 5960 0.03 (0.02) 2.1 1.18 1.03 5.5 3.2 2.2
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

5 494 600 3575 0.02 (0.05) 2.1 1.40 1.98 5.7 3.3 2.4
6 422 507 3970 0.74 (1.0) 2.4 0.98/1.08 — 5.3 — —
Bithiophenef 303 362 5380 0.01 3.5 1.32 — 5.7 — —
a
Absolute quantum yield at room temperature determined with an integrating sphere. Values in parentheses refer to absolute Ffl measured at 180 K in
THF. b Spectroscopically derived energy-gap. c Relative to SCE. d Relative to vacuum calculated from the corresponding redox onset values and correcting
for 4.3 eV. e Electrochemically derived energy-gap from the electrochemical redox onset values. f Taken from ref. 5, 8 and 25 in dioxane and acetonitrile.

to identify the room temperature deactivation processes, in parti-


cular those occurring by internal conversion such as rotation
around the thiophene–thiophene bond.26,27 Deactivation by this
process is suppressed at reduced temperatures resulting in fluores-
cence enhancement. This is in contrast to singlet excited state
quenching by ISC to the triplet state that is an inherent property.
As a result, it is temperature independent. The cryofluorescence was
done in THF because of its low freezing temperature, low polarity,
and aprocity ensuring high fluorescence. Despite this, the measur-
able cryofluorescence temperature range was 300 to 180 K. While
lower temperatures would be beneficial for gaining information
about the absolute temperature required to completely suppress the
intramolecular deactivation processes, the lower temperature limit
for the cryofluorescence measurements had to be maintained at
10 K above the solvent freezing point. This was done to avoid the
large solvent volume change at the solvent freezing point that
Fig. 4 Photographs of 1 to 6 in dichloromethane under ambient light could potentially crack the cuvette during the measurements.
(top) and irradiated with a 350 nm UV lamp (bottom). Nonetheless, intramolecular molecular dynamic modes of fluores-
cence quenching were expected to occur within the operating
Cryofluorescence
temperature range of the cryofluorescence measurements.28
To further investigate the room temperature fluorescence The temperature dependent fluorescence of 1 is seen in
deactivation modes of 1, 4, and 5, their cryofluorescence was Fig. 5. It is evident that the fluorescence increases as the
carried out. Low temperature fluorescence provides the means temperature is decreased. This trend points towards fluorescence

Fig. 5 Cryofluorescence of 1 in anhydrous tetrahydrofuran from 300 ( ) to 180 K (TT), uncorrected for the temperature dependent refractive
index change of the solvent. Inset: temperature dependent fluorescence quantum yield of 1 in anhydrous and deaerated tetrahydrofuran, corrected
for the temperature dependent solvent refractive index change.

6950 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 This journal is c the Owner Societies 2012
View Article Online

deactivation by rotation around the thiophene–thiophene bond. reorganization, resulting in its reduced capacity to stabilize the
While deactivation via the loose belt effect of the amine cannot fluorophore’s excited state.13 However, the red shift implies an
be ruled out, the high Ffl measured for both 2 and 3 implies increased stability of the excited state at reduced temperatures.
that rotation around the thiophene–amine bond is a minor This is a result of increased planarization of the bithiophenes by
deactivation mode. adopting the anti configuration.29 This places the nitro group
Using the absolute Ffl measured at 300 K, the absolute Ffl as proximal to the thiophene for 4 and 5. The O4  S2 distance
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

a function of temperature could be calculated according to calculated from the XRD data is 2.616(1) Å. The crystal data
eqn (2) when taking into account the temperature dependent suggest that an O  S interaction occurs between the nitro and
refractive index of the solvent described by eqn (1). The thiophene given that the measured distance is smaller than the
temperature effect on the Ffl is seen in the inset of Fig. 5. It van der Walls radius (3.25 Å).30 However, this interaction is not
is clear that the Ffl of 1 at 180 K increased 6.5 times relative expected in the ground state because of poor molecular overlap of
to the fluorescence at 300 K. While the exact fluorescence the heteroatoms caused by constrained angles. The absence of
enhancement possible at temperatures lower than the operating O  S interaction in the ground state is supported by the lack
limit of the cryofluorescence measurements cannot be determined, of temperature dependent changes in the absorbance spectra
the Ffl of 1 nonetheless increased from 6% at 300 K to 38% at (see ESIw). In contrast, the bond elongation and slight bending
180 K. This fluorescence enhancement confirms that bond of the thiophene–thiophene bond in the excited state favour the
rotation is a significant deactivation mode for 1. It further O  S interaction, which is in part responsible for the large red shift.
confirms that deactivation of the excited state by ISC is not the In contrast to 1, both 4 and 5 exhibited little fluorescence
exclusive fluorescence quenching mode, as is the case with enhancement at low temperatures. The fluorescence increase
unsubstituted bithiophenes at room temperature. for 4 and 5 was 2.4 and 5.3 times, respectively. Both internal
The cryofluorescence of 2, 3 and 6 was also done as control conversion and fluorescence therefore account for less than
experiments. These were done to ensure that the measured 10% of the singlet excited state deactivation modes. The weak
cryofluorescence enhancement was a true phenomenon and fluorescence is not surprising given that the nitro group is
not from solvent effects. The three compounds were examined a well known fluorescence quencher owing to its electron
because of their high fluorescence at room temperature. Little inductive and mesomeric withdrawing properties.13,31,32 These
fluorescence enhancement was therefore expected at reduced lead to efficient singlet excited state deactivation by photo-
temperatures. The fluorescence of 2, 3 and 6 at 180 K induced electron transfer. To determine whether this was the
increased 1.3 times relative to the fluorescence at 300 K. The case for both 4 and 5, the fluorescence quenching of 3 was
corrected fluorescence enhancement at low temperatures leads examined with nitromethane and 2-nitrothiophene. 3 was
to Ffl E 1. The unity fluorescence yield confirms that the slight chosen because accurate quenching studies are possible owing
reduction in fluorescence at room temperature observed for 2, to its high fluorescence. Also, its similar structure to 4 and 5
3, and 6 is a result of molecular dynamics such as bond rotation. makes it a suitable model compound.
Unlike 1, the cryofluorescence of 2–6 led to red shifts at Both nitromethane and 2-nitrothiophene were selected as
reduced temperature. The most pronounced shift was 45 nm quenchers because they are transparent in the spectral window
for 5 (Fig. 6), while 3 and 4 were shifted by 20 nm. The used for the quenching studies. Nitromethane is further known
fluorescence at 180 K of both 2 and 6 was shifted by only 10 nm as an efficient fluorescence quencher, while 2-nitrothiophene is
relative to their fluorescence at 300 K. Normally, a hypsochromic isoelectronic to the nitro group substituted at the 4 0 -position of
shift occurs at reduced temperatures because of slower solvent 4 and 5.13,26,27 The measured bimolecular quenching constants for

Fig. 6 Cryofluorescence of 5 in anhydrous tetrahydrofuran between 300 ( ) and 180 K (TT).

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 6951
View Article Online

nitromethane and 2-nitrothiophene were 1.6  109 and 9.1 


1011 M1 s1, respectively. The diffusion controlled quenching with
2-nitrothiophene implies rapid and efficient fluorescence quenching
by photoinduced electron transfer between the bithiophene and the
nitro quencher. The photoinduced electron transfer process is
slower for nitromethane owing to its higher reduction potential
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

compared to 2-nitrothiophene. Nonetheless, the observed


quenching confirms that the fluorescence of both 4 and 5 is
quenched by intramolecular photoinduced electron transfer
involving the nitro group in the 4 0 -position at room temperature.
This deactivation process occurs because the nitro group is
electronically isolated. This is in contrast to 3, where the nitro
group is conjugated with the heterocycles.

Laser flash photolysis


Laser flash photolysis of 1 to 6 was additionally investigated
for providing more insight into the deactivation modes of the
bithiophenes. This transient absorbance method provides the Fig. 8 Maximum triplet absorbance of 2-methylnaphthalene as a
function of laser power monitored at 420 nm for 1 ( ) and benzo-
means to spectroscopically detect light induced transients such as
phenone (’) in dichloromethane. Inset: triplet decay of 2-methyl-
triplets, radical ions, or photoproducts. As seen in Fig. 7, both 1
naphthalene generated by energy transfer from 1 ( ) and
and 2 produced transients that absorbed at 425 and 450 nm, benzophenone (’) at 7.4 mJ laser power at 355 nm.
respectively, upon excitation at 355 nm. While the first order decay
(inset, Fig. 7) suggests that the transient of 2 is a triplet, its identity Given that both 1 and 2 produced triplets visible with
was confirmed by quenching with known triplet quenchers. For the LFP system, their yield of triplet formation (FTT) was
example, the transient was quenched both with 1,3-cyclohexadiene determined by relative actinometry against benzophenone
and 2-methylnaphthalene with bimolecular rate constants of 1.1  (FTT = 1).23,24 Normally, the molar absorptivity of the triplet
109 and 5.6  108 M1 s1, respectively. to be determined must be accurately known for determining
In contrast to 2, the transient of 1 dissipated with mixed first the FTT directly against benzophenone.34 However, the
and second order kinetics. The rapid first order component measurements can be simplified by choosing a triplet quencher
was quenched with both 1,3-cyclohexadiene and 2-methyl- that produces the same transient in both the reference and sample
naphthalene with bimolecular rate constants of 3  109 and to be analysed. This was done by preparing samples of benzo-
6.4  109 M1 s1, respectively. The longer lived component phenone, 1 and 2 that were optically matched to within 1% at the
was unaffected with the addition of triplet quenchers. While excitation wavelength of 355 nm. Identical amounts of 2-methyl-
the first order fast kinetics was assigned to the triplet, the long naphthalene were added to each sample such that 95% of the
lived second order component most likely is a radical cation. triplets produced would be rapidly quenched within the laser
This intermediate is not surprising since bithiophenes are pulse. Upon excitation, the triplet of the sample to be analyzed is
known to readily photo-oxidize.33 Of particular interest are rapidly deactivated by energy transfer to produce the triplet of
3–6 that did not produce detectable transients when irradiated. methylnaphthalene, which is visible at 420 nm. The maximum
signal corresponding to the triplet of methylnaphthalene is then
examined as a function of laser power. The triplet quantum yield
can then be derived from the ratio of the slopes in Fig. 8,
according to eqn (3). Using this method, the FTT of 1 was
calculated to be 0.4. The combined LFP and temperature
dependent fluorescence data confirm that both ISC and nonradia-
tive IC are the predominate fluorescence deactivation modes of 1.
Unfortunately, the FTT of 2 could not be determined because
of its ground state bleaching that overlapped with the methyl-
naphthalene triplet at 420 nm. Despite this, its FTT is expected
to be negligible given that the cryofluorescence confirmed that
the singlet excited state is deactivated by fluorescence in 83%
and ca. 17% by internal conversion.
SlopeSample
FSample
TT ¼ FBenzophenone ð3Þ
SlopeBenzophenone TT

Fig. 7 Transient absorbance spectra of 1 (’) and 2 ( ) recorded at Solvatochromism


1.3 and 2.4 ms, respectively, after the laser pulse at 355 nm in
acetonitrile. Inset: transient absorbance kinetics of 1 (TT) and 2 ( ) The spectroscopic properties of 2–6 should be solvent dependent
recorded at 420 and 450 nm, respectively. owing to their donor–acceptor arrangement. In particular, both

6952 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 This journal is c the Owner Societies 2012
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM. View Article Online

Fig. 9 Top: normalized absorbance (left) and normalized fluorescence (right) of 2 in (’), ether ( ), tetrahydrofuran ( ), ethyl acetate ( ),
chloroform ( ), dichloromethane ( ), dimethyl formamide ( ), acetonitrile ( ), dimethyl sulfoxide ( ) and ethanol ( ). Photographs of 2
(middle panel) and 3 (top panel) taken under ambient light (left) and irradiated with a 350 nm UV lamp (right) in solvents of increasing polarity
upon going from left to right.

the absorbance and fluorescence of these compounds should yellow to red. The larger solvent induced bathochromic shifts
undergo solvent induced changes. Red shifts in the absorbance for the fluorescence compared to those for the absorbance
and fluorescence are expected with polar solvents owing to the confirm that the excited state is more polar than the corre-
stabilization of the intramolecular charge transfer (ICT) sponding ground state for 2–6. The larger bathochromic shift
occurring between the terminal donor–acceptor groups. To observed for 3 relative to 2 and 6 suggests that an ICT state is
confirm this, the absorbance and fluorescence spectra of 2–6 formed in the excited state. Nonetheless, the fluorescence
were investigated in both aprotic and protic organic solvents yields of both 2 and 3 are solvent independent, with the
of varying polarity. The effect of solvent polarity is seen in the exception being alcohols (vide infra). The consistent Ffl
representative absorbance spectra for 2 in Fig. 9. Bathochromic confirms that deactivation by ISC is not a competitive fluores-
shifts ranging between 13 and 45 nm were observed for 2–6. As cence quenching deactivation mode for 2, 3, and 6. Otherwise,
expected, smaller shifts were observed for 2 and 6 owing to the significant fluorescence quenching would be seen in polar
weaker electron withdrawing carbonyl group. In contrast, the solvents with these bithiophenes, owing to the narrowing of
largest shift of 45 nm was observed with 3. This is a result of the singlet–triplet energy gap, which favors ISC to the triplet
the stronger withdrawing character of the nitro group and the manifold.
collective electronic push–pull effect. The effect of the solvent The solvent induced changes were examined according to
polarity on the absorbance of 2 and 3 is evident in the lower the Lippert–Mataga method.13 This involves plotting the
panels of Fig. 9. While slight colour changes between transparent observed Stokes shift (vst) as a function of the solvent’s
and green were possible with 2, the solvent induced colour orientation polarizability (Df), which is described by the
changes were more pronounced with 3, with colours ranging solvent’s refractive index (n) and its dielectric constant (e),
from yellow to red. according to eqn (4).
The solvent induced shifts were also observed with the
e1 n2  1
fluorescence of 2–6. Similar to what was observed with the Df ¼  ð4Þ
2e þ 1 2n2 þ 1
absorbance, the smallest red shifts (67 nm) were seen with 2
and 6. Meanwhile, the largest solvent induced shift of 92 nm The Stokes shift is related to Df according to eqn (5), where m
was observed for 3, whilst 4 and 5 were both shifted by a is the solvent’s molecular weight, d is the solvent density, Na is
maximum of 80 nm. The effect of solvent polarity on the Avogadro’s number, c is the speed of light, h is Planck’s
emission colour is evident in Fig. 9 with the emission colour of constant, and me and mg are the excited and ground state
2 varying between blue and green. In contrast, the range of dipoles, respectively. Therefore, the Lippert–Mataga plot
colours is more extensive with 3, with colours ranging from should be linear, provided the solvent–solute interactions are

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 6953
View Article Online

(%)
Fflb


64
64
89
61
63

60
74

12
2
shift/cm1
Stokes

3205
3545
3976
4006
3912

4796
3970

5336
5889

6
lem/
nm
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

476
477
497
492
504

515
507

538
544

labs/
nm
413
408
415
411
421

413
422

418
412

(%)
Fflb

B0

B0

B0
4
1

2
1
2

1
shift/cm1
Stokes

2699
2525
2932
3067
3477
3575
4310
3617
3158
3009
5
lem/
nm
566
567
590
585
595
600
626
645
598
583
Fig. 10 Lippert–Mataga plot showing the Stokes shift of 2 as a
function of the solvent orientation polarizability (Df).

labs/
nm
491
496
503
496
493
494
493
523
503
496
exclusively dipolar. As seen in Fig. 10, there is a poor linear

(%)
Fflb

B0
4
2
2
2
2
3

1
2
2
correlation of the Stokes shift of 2 with the solvent orientation
polarizability. The linear regression of the Lippert–Mataga

shift/cm1
Stokes
plots for 2–6 ranged from 0.08 to 0.71. Poor linear correlations

4541
5128
4990
5325
5573
5960
6153
5393
5579
5346
were observed with 4–5, implying that strong solute–solvent

4
interactions other than dipole–dipole interactions occur in the

Absolute fluorescence quantum yields determined with an integrating sphere.


excited state. Also, the lack of good linear correlation for 2–6

lem/
nm
564
585
592
597
606
615
630
644
608
603
further suggests that strong solute–solvent interactions take

labs/
place for the bithiophenes studied (Table 3).
nm
449
450
457
453
453
450
454
478
454
456
8Df pdNa
ðmE  mG Þ2 þ constant
(%)
Fflb

B0
B0
uSt ¼ ð5Þ
97
92
85
79
62
76
34

3hcm
shift/cm1

The Reichardt–Dimroth’s ET(30) solvent index is a more


Stokes

4310
4121
4253
4475
5042
5170
5474
4703
4739
5024
versatile parameter for correlating the solvent induced fluores-
cence changes given that it takes into account both general
3

polarity and hydrogen bonding effects.37 The fluorescence


lem/
nm
566
560
589
584
604
600
631
652
608
617
maximum was subsequently examined as a function of
ET(30). As seen in Fig. 11, a good correlation with this solvent
labs/
nm
455
455
471
463
463
494
469
499
472
471
parameter was obtained for all the bithiophenes. The linear
trend is valid for all the solvents, with the exception of
(%)
Fflb

82
78
78
81
74
83
76
73
17
4

methanol and ethanol, which are grouped within the red circle
in Fig. 11. It should be noted that similar trends were also
Table 3 Spectroscopic properties of 2–6 in different solvents

shift/cm1
Stokes

observed for the Stokes shift as a function of ET(30) for all the
3850
4110
4006
4097
4710
4550
5001
4717
5554
5714

bithiophenes. The observed fluorescence shifts are less than


2

predicted for 3–5 for these two protic solvents. Only in the case
of 2 and 6 was the linear trend extended to include the two
lem/
nm
477
483
492
487
502
498
511
530
541
544

polar protic solvents. The linear trend observed for 2 and 6


suggests that hydrogen bonding is not the reason for the
labs/
nm
403
403
411
406
406
406
407
424
416
415

observed deviation from linearity for the protic solvents


b
Values taken from ref. 35 and 36.

for 3–5. While hydrogen bonding does not account for the
a
ET(30)
33.9
34.5
37.4
38.1
39.1
40.7
45.6
45.4
51.9
55.4

deviation from linearity for 3–5 in protic solvents, the origin is


most likely a result of the nitro group.
To further investigate the origins of the decreased Stokes
0.013
0.167
0.210
0.200
0.149
0.218
0.305
0.263
0.290
0.309
Df

shift with the polar protic solvents, the fluorescence quenching


of 3 with methanol was examined in dichloromethane.
Dichloromethane

Dimethyl sulfoxide
Tetrahydrofuran

Aliquots of methanol were added to a maximum of 5 vol%.


Ethyl acetate

This was to ensure that the absorbance did not change, which
Chloroform

Acetonitrile

Methanol

would otherwise require correcting the fluorescence intensities.


Toluene

Ethanol
Solvent

Ether

The fluorescence of 3 was quenched with the addition of


methanol as seen in Fig. 12. The fluorescence lifetime also
a

6954 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 This journal is c the Owner Societies 2012
View Article Online

important role in deactivation. As seen in the inset of Fig. 12,


the quenching of 3 with deuterated methanol is 2.6 times
slower than with methanol. The difference confirms that
hydrogen abstraction from the solvent by the excited nitro is
the primary deactivation mode.27 This was further confirmed
by similar quenching of 3 with isopropanol. The same quenching
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

kinetics were observed when measuring the change in fluores-


cence lifetimes of 3 as a function of added quencher, confirming
that the quenching is process dynamic. Hydrogen abstraction by
the nitro group is therefore responsible for the deviation from
linearity for the fluorescence of 3–5 vs. ET(30) plot, especially
since this deactivation is not accounted for in the ET(30)
parameters.

Conclusion
It was found that highly fluorescent bithiophenes were possible
Fig. 11 The change of fluorescence maximum of 2 (’), 3 ( ), 4 ( ), by incorporating complementary electron donating and
5 ( ) and 6 ( ) as a function of ET(30) of various solvents. accepting groups in the 2,2 0 -positions. The complementary
electronic push–pull groups, when conjugated, suppressed the
otherwise efficient ISC fluorescence quenching mode in mono-
and unsubstituted bithiophenes. The stronger electron with-
drawing character of the nitro group led to intensely coloured
absorbance and fluorescence in the visible when it was placed
in the 2 0 -position. Both the absorbance and fluorescence could
be varied with solvent polarity with colours covering a large
portion of the visible spectrum. These solvatochromic properties
concomitant with high fluorescence quantum yields make the
push–pull bithiophenes interesting polarity probes for potential
biological imaging and sensor applications. Meanwhile, the
cryofluorescence and quenching studies provided insight into
the structural and electronic requirements for suppressing
competitive fluorescence deactivation modes. This knowledge
is pivotal for the design and subsequent preparation of highly
fluorescent oligothiophenes for potential use as functional
materials, in particular, in plastic electronics.

Fig. 12 Fluorescence quenching of 3 as a function of methanol (’), Acknowledgements


deuterated methanol ( ) and isopropanol ( ). Inset: the fluorescence
change of 3 measured in dichloromethane as a function of added Financial support from the Natural Sciences and Engineering
methanol. Research Council of Canada is acknowledged for Discovery,
Strategic Research, and Research Tools and Instruments
shortened with the addition of methanol. Both the steady-state grants. The Canada Foundation for Innovation is also
and lifetime measurements confirm that methanol quenches acknowledged for additional equipment and infrastructure.
the excited state of 3 with 3.7  109 M1 s1 kinetics, based on The Centre for Self-Assembled Chemical Structures is further
the measured fluorescence lifetime of 2.7 ns. The slower acknowledged. A.B. thanks both NSERC and the Université
diffusion controlled rate constant implies that the deactivation de Montréal for a graduate scholarship. Prof. G. Cosa and
process is not efficient. However, the quenching process is Ms. K. Krumova are acknowledged for their help with LFP
more efficient when the alcohol quencher is used as the solvent measurements as well as Prof. G. Hanan and Mr. D. Chartrand
because of its much higher effective concentration than the for assistance with the lifetime measurements.
excited state concentration of the fluorophore.
While the exact mode of deactivation cannot be assigned References
from these quenching measurements, the quenching was
1 T. F. Otero, I. Boyano, G. Vazquez and M. T. Cortes, Proc. SPIE,
further done with deuterated methanol. Slower quenching 2004, 5385, 425.
kinetics with deuterated methanol would confirm that the 2 J. Chen and Y. Cao, Acc. Chem. Res., 2009, 42, 1709–1718.
protic solvent quenches the fluorescence by intermolecular 3 J. M. Szarko, J. Guo, B. S. Rolczynski and L. X. Chen, J. Mater.
hydrogen abstraction. In contrast, similar quenching kinetics Chem., 2011, 21, 7849–7857.
4 K. C. Persaud, Mater. Today, 2005, 8, 38–44.
for both the deuterated and non-deuterated solvents would 5 R. S. Becker, J. Seixas de Melo, A. L. Maçanita and F. Elisei,
suggest that hydrogen bonding in the excited state plays an J. Phys. Chem., 1996, 100, 18683–18695.

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 6955
View Article Online

6 J. Seixas de Melo, F. Elisei, C. Gartner, G. G. Aloisi and 22 N. G. Connelly and W. E. Geiger, Chem. Rev., 1996, 96, 877–910.
R. S. Becker, J. Phys. Chem. A, 2000, 104, 6907–6911. 23 I. Carmichael and G. L. Hug, J. Phys. Chem. Ref. Data, 1986, 15,
7 J. Seixas de Melo, L. M. Silva, L. G. Arnaut and R. S. Becker, 1–250.
J. Chem. Phys., 1999, 111, 5427–5433. 24 S. A. Pérez Guarı̀n, S. Dufresne, D. Tsang, A. Sylla and
8 J. Seixas de Melo, F. Elisei and R. S. Becker, J. Chem. Phys., 2002, W. G. Skene, J. Mater. Chem., 2007, 17, 2801–2811.
117, 4428–4435. 25 R. J. Waltman, J. Bargon and A. F. Diaz, J. Phys. Chem., 1983, 87,
9 Y. Dong, A. Bolduc, N. McGregor and W. G. Skene, Org. Lett., 1459–1463.
Published on 28 March 2012. Downloaded by University of Tennessee at Knoxville on 6/18/2018 10:02:56 PM.

2011, 13, 1844–1847. 26 N. J. Turro, V. Ramamurthy and J. C. Scaiano, Principles of


10 A. Bolduc, S. Dufresne and W. G. Skene, J. Mater. Chem., 2012, Molecular Photochemistry: An Introduction, University Science
22, 5053–5064. Books, Sausalito, 2009.
11 T. Tshibaka, S. Bishop, I. U. Roche, S. Dufresne, W. D. Lubell 27 N. J. Turro, V. Ramamurthy and J. C. Scaiano, Modern Molecular
and W. G. Skene, Chem.–Eur. J., 2011, 17, 10879–10888. Photochemistry Of Organic Molecules, University Science Books,
12 S. Dufresne and W. G. Skene, J. Phys. Org. Chem., 2012, 25, 211–221. 2010.
13 J. R. Lakowicz, Principles of Fluorescence Spectroscopy, Springer, 28 D. Vitkup, D. Ringe, G. A. Petsko and M. Karplus, Nat. Struct.
New York, 2006. Mol. Biol., 2000, 7, 34–38.
14 O. A. Kucherak, L. Richert, Y. Mely and A. S. Klymchenko, Phys. 29 L. Liao and Y. Pang, J. Mater. Chem., 2001, 11, 3078–3081.
Chem. Chem. Phys., 2012, 14, 2292–2300. 30 R. C. Weast, Handbook of Chemistry and Physics, CRC Press,
15 B. E. Cohen, A. Pralle, X. Yao, G. Swaminath, C. S. Gandhi, Cleveland, Ohio, 1975.
Y. N. Jan, B. K. Kobilka, E. Y. Isacoff and L. Y. Jan, Proc. Natl. 31 T. Ueno, Y. Urano, H. Kojima and T. Nagano, J. Am. Chem. Soc.,
Acad. Sci. U. S. A., 2005, 102, 965–970. 2006, 128, 10640–10641.
16 A. P. Demchenko, Y. Mély, G. Duportail and A. S. Klymchenko, 32 J. B. Birks, Photophysics of Aromatic Molecules, Wiley-
Biophys. J., 2009, 96, 3461–3470. Interscience, London, 1970.
17 V. V. Shvadchak, A. S. Klymchenko, H. de Rocquigny and 33 C. H. Evans and J. C. Scaiano, J. Am. Chem. Soc., 1990, 112,
Y. Mély, Nucleic Acids Res., 2009, 37, e25. 2694–2701.
18 G. S. Loving, M. Sainlos and B. Imperiali, Trends Biotechnol., 34 I. Carmichael, W. P. Helman and G. L. Hug, J. Phys. Chem. Ref.
2010, 28, 73–83. Data, 1987, 16, 239–260.
19 G. Openhaim and E. Grushka, J. Chromatogr., A, 2002, 942, 63–71. 35 I. M. Smallwood, Handbook of Organic Solvent Properties,
20 J. C. Scaiano, CRC Handbook of Organic Photochemistry, CRC Elsevier, 1996.
Press, Boca Raton, 1989. 36 J. A. Riddick, W. B. Bunger and T. K. Sakano, Organic Solvents:
21 I. Carmichael and G. L. Hug, in Handbook of Organic Photo- Physical Properties and Methods of Purification, Wiley, 1986.
chemistry, ed. J. C. Scaiano, CRC Press, Boca Raton, Florida, 37 C. Reichardt, Solvents and solvent effects in organic chemistry,
1989, vol. I, pp. 369–403. Wiley-VCH, Germany, 2003.

6956 Phys. Chem. Chem. Phys., 2012, 14, 6946–6956 This journal is c the Owner Societies 2012

You might also like