Você está na página 1de 7

4596

J. Phys. Chem. B 2001, 105, 4596-4602

Time-Resolved Fluorescence in 3-Dimensional Ordered Columnar Discotic Materials


A. Bayer,* J. Hu1 bner, J. Kopitzke, M. Oestreich, W. Ru1 hle, and J. H. Wendorff*
Department of Chemistry, Department of Physics, Center of Material Science, Philipps UniVersity,
Marburg, Germany
ReceiVed: September 27, 2000; In Final Form: February 26, 2001

Absorption and fluorescence were investigated for liquid-crystalline discotics, which are characterized by the
spontaneous formation of one-dimensional columnar structures in the fluid phase. Such materials have been
considered for applications in organic light-emitting diodes and as photoconductors. We investigated materials
based on asymmetrically substituted triphenylenes displaying a novel highly ordered plastic columnar state.
These materials show an unexpected time dependence of the fluorescence spectrum during irradiation apparently
because of their specific spatial structure. Transfer of energy from a high-energy excited state to a newly
developing lower-energy state takes place. We attribute the evolution of this state to the particular spatial
arrangement of the molecules within the columns in the plastic columnar state. This causes the photoinduced
formation of dimers, a process that is absent in solutions and in polymer-dispersed systems of discotic materials
and that has, so far, not been documented in the literature.

I. Introduction
The electronic states of a given organic molecule are known
to depend strongly on the molecules environment. Prominent
examples are the effect of the polarity of the solvent in the case
of solutions or of the structural disorder in the case of solid
guest-host systems on absorption and emission.1-4 The effect
of disorder, characteristic for the amorphous state, on electronic
excited states has been the subject of intense investigations.2-4
Molecular assemblies occurring either in solution or in the
condensed state are, in this respect, of considerable significance
because, in general, they give rise to strong modifications of
the electronic states and, thus, of both the absorption and the
emission spectra.5-7
The spontaneous formation of molecular assemblies extending
over several tens of nanometers is a characteristic feature of
liquid crystals. Thermotropic liquid-crystalline materials display
both an orientational order within the fluid state and various
types of intermediate positional order, differing from the threedimensional crystalline order.8,9 Of particular interest, as far as
electronic excitations are concerned, are discotic liquid-crystalline phases displayed by flat disk-shaped molecules. Such
materials have been considered for use in organic light-emitting
diodes, sensors, and field-effect transistors and as photoconductors.10-16 The disk-shaped molecules organize spontaneously in one-dimensional stacks within the fluid state, and the
stacks form a two-dimensional lattice.8,9 Figure 1 displays
characteristic examples for such columnar discotic structures.
Columnar phases with various degrees of order can be induced
depending on the choice of chemical structure and the temperature. Homogeneous films can be prepared via spin-coating in
which all columnar axes are oriented either within the plane of
the film or, after annealing, parallel to the film normal.17 The
ordered molecular and supermolecular state can often be frozen
into a glassy ordered state.
Such self-organization is known to influence electronic
properties significantly. It has been reported, for instance, that
the mobility of holes increases by several orders of magnitude
because of the formation of columns.10-12 Investigations have

also revealed that discotics are characterized by a very rapid


one-dimensional transport of excitons and charge carriers along
the columns, whereas transport perpendicular to this direction
is more than 4 orders of magnitude slower.10-12 Previous optical
studies on similar substances have revealed broad absorption
and photoluminescence bands with sidebands resulting from the
coupling of electronic transitions to molecular vibrations.18-20
We discuss here the absorption and emission properties of
two substances that show a novel highly ordered discotic state,
the plastic columnar phase (colhp) in which the columns are
ordered on a two-dimensional hexagonal lattice. It is characterized by a three-dimensional regular arrangement of the centers
of the molecules as in a crystalline state.21 Yet, the molecules
are able to rotate around the columnar axis, giving rise to a
rotational disorder. The repeat distance of the disklike molecules
along the columns amounts to about 0.35 nm, whereas the
distance between the columns is on the order of 2 nm. This
state is not solid but can be frozen into a highly ordered glassy
state.22 Time-of-flight investigations have revealed that the
transition from a lower ordered to this more highly ordered
columnar state causes the hole mobility to increase by at least
1 order of magnitude.12 A particular feature that will be shown
to be of considerable importance for optical properties is that
the molecules have a tendency toward dimer formation within
the columns in this phase, as is apparent from X-ray analysis.21
Our aim was to analyze the effect of the enhanced order on
the optical excitation and relaxation processes. We performed
absorption and fluorescence experiments in solution and in the
condensed state, including time-resolved fluorescence investigations, and we found that, in fact, the optical properties are
influenced significantly by the specific discotic order studied
here.
II. Experimental Section
We synthesized asymmetrically substituted triphenylene
compounds that are characterized by a reduced tendency toward
crystallization as compared to that of the symmetric derivatives.22 This allows us to study solid films retaining the specific

10.1021/jp003505k CCC: $20.00 2001 American Chemical Society


Published on Web 04/26/2001

Time-Resolved Fluorescence in Discotic Materials

J. Phys. Chem. B, Vol. 105, No. 20, 2001 4597

Figure 1. Columnar structures displayed by discotics: orientational order of the disks, different intracolumnar distance distributions, various
arrangements of the columns.

SCHEME 1: Chemical Structure of the Triphenylene


Derivatives

order characteristics of the discotic columnar phases. In the


following, we consider the adamantanoate derivatives 4 and 5
(Scheme 1), which display the plastic columnar phase. Derivative 4 shows an additional superstructure because of the spatial
separation of the substituents, as shown schematically in Figure
2. This is apparent from X-ray investigations. The synthesis of
compound 5 and the characterization of its spatial structure and
molecular mobility have been reported previously;17,21 compound 4 has been synthesized equivalently. Thermal properties
of both substances are given in Table 1. It is apparent that the
more highly ordered compound 4 has a higher glass transition
temperature as well. This substance is thus in a glassy state
under experimental conditions (room temperature), whereas the
derivative 5 is in the fluid colhp phase.
The substances were synthesized with a purity of 99.9 wt %,
as is apparent from NMR and chromatographic analyses.
Solutions used for optical investigations had a concentration of
10-5 mol/L. Corresponding films were prepared from solutions
of 2 wt % via spin-coating, the residual solvent was removed
during 0.5 h of annealing at 65 C in a vacuum.
Absorption studies were carried out using a Perkin-Elmer
Lambda 9 spectrometer with a spectral resolution of about 2
nm. Steady-state fluorescence and excitation spectra were
recorded with a Shimadzu RF-1502 spectrofluorometer (light
source ) XBO lamp) with a resolution of about 15 nm. The
light power at the excitation wavelength was in the range below
1 mW.

Figure 2. Possible models of the superstructure formed by derivative


4 determined from X-ray investigations.

TABLE 1: Phase-Transition Temperatures and Exposed


Structures of the Investigated Substances
compound

phase transitions
7080 C

168 C

188 C

adamantanoate 4

glass 98 colhp 98 colho 98 isotropic

adamantanoate 5

glass 98 colhp 98 colho 98 isotropic

20 C

135 C

186 C

The fluorescence decay behavior of the solutions was


analyzed with a gated optical multichannel analyzer (oma) using
10-ns pulses of a Neodym-YAG laser and an optical parametrical oscillator (opo). The repetition rate of the pulses was 10
Hz. The accessible time range of this experimental setup extends
from about 10 ns to a few hundreds of microseconds, depending
on the shift and size of the gate window. The energy per pulse
was 70 J, which is equal to a light power of 0.7 mW. Time-

4598 J. Phys. Chem. B, Vol. 105, No. 20, 2001

Bayer et al.

Figure 3. Absorption spectrum of adamantanoate 5 (;) in solution


(trichloroethane, TCE) in comparison with that of the unsubstituted
triphenylene core ().

Figure 4. Absorption spectra of adamantanoates 4 (- - -) and 5


() in the condensed columnar state, in comparison with the solution
data (;).

resolved spectra for investigations of the condensed state were


obtained using 1.5-ps pulses of a mode-locked frequency-tripled
Ti:sapphire laser with a repetition rate of 80 MHz for excitation
and a light power of 12 mW. The fluorescence was spectrally
dispersed in a 0.32 m spectrometer and temporally resolved with
a streak camera with a two-dimensional readout and a time
resolution of 2 ps. All experiments were carried out at room
temperature with the samples exposed to the atmosphere, with
the exception of the streak-camera experiment where the films
were placed in an evacuated cryostat (p 5 10-5 mbar).

exciton interactions. The shift depends on the state of organization, in particular, the number of interacting molecules within
the columns during the excitation.23,24 On the basis of the results
of the quantum chemical calculations for the symmetric species,
we expect that the number is probably above 10. In fact, the
correlation length, as determined by X-ray investigations, is on
the order of 15-20 nm along the columns, corresponding to a
number of structurally correlated molecules on the order of 4060.25 A prominent feature of the spectrum obtained for the
condensed state is the occurrence of a broad high-energy band
at about 5.9 eV that is not present in the solution spectra. Its
origin is still an unsolved question.
Photoluminescence. The photoluminescence spectra observed
for solutions of component 5 [solvents tetrahydrofurane (THF),
chloroform (Chl), and trichloroethane (TCE)] are displayed in
Figure 5a. Again, the solutions of adamantanoate 4 show a very
similar behavior. The emission probably results from the S1S0 transition, as judged from the results of quantum chemical
calculations on the symmetric model compounds.23 The various
weak emission maxima that are apparent in Figure 5a are
vibronic fine structures. Time-resolved fluorescence studies
revealed decay times of about 16 ns in THF solution, which is
in the neighborhood of the values (about 7.1 ns in 1,2dichloroethane solution) reported for the symmetrically substituted hexakis(n-alkoxy)triphenylenes having different lateral
alkoxy chains.26 We observed a particular solvent effect in the
case of chloroform leading to time dependent changes in the
emission, but no solvent effects for other solvents. It seems that
the chloroform effect arises from a photochemical reaction with
chlorine radicals or phosgene.
The emission spectra of both compounds in the condensed
state are displayed in Figure 5b. In general, the two substances
show very similar emission behaviors. We observe a red shift
compared to the solution spectra. It amounts to about 0.06 eV
for derivative adamantanoate 4 and 0.13 eV for compound 5.
This difference in the shift apparently results from the spatial
superstructure of derivative 4, which thus has an influence on
the optoelectronic properties of the substance; however, the
influence is weak. A similar red shift has been reported for a
symmetrical hexa-alkylthiotriphenylene in its neat phases, and
it was explained in terms of an emission from weakly bound
excimers.23 Yet, no such red shift has been reported for the
symmetrically substituted hexakis(n-alkoxy)triphenylenes.20
Our interpretation is that the newly formed plastic columnar
phase characterized by a strong tendency toward dimer formation is responsible for the observed shift in the case of the

III. Results and Discussion


Absorption Spectra. The UV-vis absorption spectrum of
adamantanoate derivative 5 as obtained in solution (solid line)
is displayed in Figure 3; compound 4 has a very similar
spectrum. Figure 3 also contains the absorption spectrum of the
symmetric unsubstituted triphenylene compound for comparison
(dashed line). It is obvious that the absorption spectra of
symmetrically substituted triphenylene compounds are very
similar to those described here.23 This similarity can be taken
as an indication that the asymmetrical chemical substitution
constitutes only a small perturbation as far as the optical
properties are concerned. The absorption bands are broad in all
cases; some weak maxima are apparent. In the spirit of quantum
chemical calculations performed on model compounds (symmetrical triphenylene derivatives), we tentatively attribute the
two lowest, weak-energy maxima to two weak symmetryforbidden transitions: S0-S1 and S0-S2.23 The other maxima
would correspond in this case to the transitions S0-S3 and S0S4 (maximum of the absorption band), and they are characterized
by vibronic progressions. The transitions are expected to be
polarized within the planes of the triphenylene cores. The
adamantanoate absorption bands are only slightly red-shifted
compared to the one of the model compound, and the vibrational
features are less pronounced. This small red shift is probably
due to the side groups of the triphenylene core which contain
oxygen and thus increase the delocalized -electron system
slightly. The signal decrease at the higher-energy end of both
spectra is due to the strong absorption of the solvent.
The absorption spectra (Figure 4) displayed by the condensed
ordered state (dashed and dotted lines) roughly resembles that
found for the solution (solid line). A closer inspection reveals
that the S0-S4 transition is blue-shifted by about 0.3 eV. Such
a blue shift has already been described in the literature for
symmetric triphenylene derivatives. In this case, the blue shift
is attributed to collective excited states originating from strong

Time-Resolved Fluorescence in Discotic Materials

J. Phys. Chem. B, Vol. 105, No. 20, 2001 4599

Figure 6. Fluorescence spectra of adamantanoate 4 in the condensed


columnar state, displaying the time dependence of the spectra during
cw excitation (stepwise enhancement of the irradiation time, each step
) 24.4 s)

Figure 5. (a) Fluorescence spectra of adamantanoate 5 in solutions


with different solvents (;, chloroform; - - - , 1,1,1-trichloroethane;
, THF). (b) Steady-state photoluminescence spectra of compounds
4 (;) and 5 (- - -) in the condensed state.

systems studied here. This interpretation is based on the results


obtained for polymer-dispersed systems in which the discotic
compounds were finely dispersed in a solid polymer matrix.24
X-ray and dielectric relaxation studies revealed in this case that,
because of geometric confinement effects, the plastic columnar
state is transformed into the hexagonal ordered phase, which
does not display a tendency toward dimer formation. This, in
turn, suppresses the red shift of the emission spectra. Thus, the
conclusion is that the colhp phase favors the formation of such
excimers.
Formation of a Low-Energy Emission Band. An unexpected observation is the formation of a low-energy emission
band. First, we consider the case of cw excitation. The
interesting finding is that a low-energy emission band grows
as a function of the irradiation time on a time scale of several
100 s, while the magnitude of the high-energy band decreases
during this time. This is shown in Figure 6. The band maximum
occurs at about 2.34 eV (2.45 eV for compound 5). A prominent
feature is the presence of an isosbestic point at 2.55 eV (2.52
eV for compound 5) on the energy coordinate. The isosbestic
point indicates a transition on this time scale between two states
that are both able to emit light. We have to point out that this
effect is not related to solvent effects originating from the spincoating conditions, as the effect occurs independent of the
solvent used for the preparation of the films. The decrease of
the high-energy band with time during irradiation can be
represented by a biexponential decay. The corresponding decay
times are on the order of 20 and 150 s for all samples. The
low-energy band, on the other hand, approaches saturation after
about 150 s.

Figure 7. Photoluminescence excitation spectra obtained for compound


5, detected within the high- (3.10 eV, 9) and low- (2.43 eV, O) energy
emission bands; also displayed is the absorption spectrum (s).

Photoluminescence excitation investigations (Figure 7) revealed, first, that the PL excitation spectra agree rather well
with the absorption spectra. Second, the excitation spectra are
very similar if detected in the wavelength ranges of the lowerand the higher-energy emission bands, respectively. The conclusion is that the excitation mechanism is apparently identical for
the two emission bands. Differences between the PLE and UVvis spectra in the high-energy regime are probably due to the
limited resolution of the spectrofluorometer.
Next, we discuss the time-resolved fluorescence (Figure 8).
Visual inspection actually shows a transition from a blue to a
greenish-blue emission during the first few seconds of irradiation, which indicates the growth of the lower-energy emission
band. The time scale of this growth differs strongly from that
discussed above; it obviously results from the higher excitation
intensity used here. Directly after excitation, the two emission
bands at 3.05 and 2.34 eV show rise times on the order of 101
ps (the growth of the lower-energy band is a factor of 2 smaller),
which can be attributed to vibronic relaxation from higher states
to the ground S1 state. The following decay of the high-energy
band can be approximated by a biexponential law with relaxation
times of 200 and 850 ps when the sample is spot excited for a
short time, typically only 1 min. These decay times are much
shorter than those reported for the symmetrically substituted
hexakis(n-alkoxy)triphenylenes in the columnar state (which
amounted to about 10 ns), indicating additional decay channels.
The low-energy band is long-lived, with a lifetime of about 5
ns (estimated). The time evolution is depicted in Figure 8, where

4600 J. Phys. Chem. B, Vol. 105, No. 20, 2001

Figure 8. Time evolution of the fluorescence spectra obtained by


streak-camera experiments. (a) Transient spectra at three time intervals
after excitation (1, 100 ps; 2, 700 ps; 3, 6.5 ns). (b) Time decay
behavior at different photon energies.

Figure 9. Dependence of the fluorescence decay on the irradiation


time.

part a shows the transient spectra at three time intervals after


excitation and part b depicts the decay behavior at different
photon energies. The differences in the relative intensities of
the two emission bands are clearly observed in both presentations, the low-energy band being the more persistent; after nearly
1 ns, there is almost no decrease of the intensity.
Next, we discuss the variation of the decay times of the highenergy emission with increasing irradiation time. Figure 9
depicts the changes after 10 and 30 min of excitation. Clearly,
both decay times of the emission at about 3 eV become shorter
with increasing irradiation time. The fast (slow) decay varies
from 200 (850) to 70 (400) to 40 (230) ps after 12-mW
irradiations of 0, 10, and 30 min, respectively. Obviously, the
decay channels become more effective with increasing irradiation.

Bayer et al.
A reasonable explanation is that energy transfer to the lowenergy state becomes more effective with increasing irradiation
time because of an increased concentration of centers involved
in the low-energy emission. We conclude that the energetic
modification, which has to be linked to some structural changes
within the columnar structure, is apparently caused by photochemical or photophysical reactions.
We summarize: all three observations, the change in cw
fluorescence, the luminescence excitation spectra, and the
change in time-resolved fluorescence, can consistently be
explained by a concentration increase of fluorescence centers
involved in the low-energy emission. The formation of excimers,
i.e., dimers in the excited state, and of aggregates that are also
stable in the ground state, has been discussed before as origin
of a low-energy emission band.27 Our time-resolved fluorescence
experiments, however, reveal for the first time the dynamics of
the energy transfer from the monomers/weakly bound excimers
to some kind of more strongly interacting aggregate (built up
through photophysics or photochemistry). No such effect has
been reported for the symmetrically substituted hexakis-(nalkoxy)triphenylenes which normally show the less-ordered colho
phase. We can rule out specific chemical degradation such as a
reaction with oxygen, for instance. The reasons are, first, that
studies on chemiluminescence have shown that such effects
occur only at elevated temperatures;28 second, that the streakcamera investigations were performed in the absence of oxygen,
in high vacuum; and third, that annealing of the samples in the
presence of oxygen for a long time at elevated temperatures
would lead to changes of the optical properties that are not
consistent with the changes reported here (namely, a very broad
emission band at about 2.58 eV). It is important to point out
that no such effects occur in solutions (chloroform solutions
display very specific effects) where oxygen is present.
Our investigations seem to indicate that the effect is strongly
controlled by the specific arrangement of the molecules within
the plastic columnar state. The plastic columnar state is known
to display a tendency toward dimer formation, as discussed
above. This is apparent from X-ray investigations.21 The
transition from the usually observed discotic columnar hexagonal
ordered state (colho) to the more highly ordered plastic columnar
state (colhp) was found to be accompanied with a doubling of
the repeat unit along the columnar axes. In addition, the
observation of a reentrant behavior, i.e., the plastic columnar
phase is transformed into the ordered phase both on heating
and on cooling, can also be taken as a strong indication of the
formation of dimers.29 Thus, our interpretation for the particular
case of the plastic columnar state is that a photoinduced
formation of specific lower-energy configurations takes place,
which can be directly linked to the specific spatial structure of
the plastic columnar state.
In fact, the fine dispersion of the discotic compound 4 in a
polymer matrix mentioned above, which is known to result in
a disruption of this specific spatial structure, causes a suppression of the formation of the low-energy band.24 This is apparent
from Figure 10, which shows the emission spectra for the
polymer-dispersed discotics for various concentrations of the
discotics. The findings of the structural analysis are that the
lower the concentration, the finer the dispersion and the stronger
the suppression of the plastic columnar state. This, in turn,
reduces the formation of the low-energy band. It is worth
mentioning that the energetic position of the lower-energy band
is, in this case, located at 2.43 eV, which is identical to that of
compound 5 showing no superstructure. This is a strong
indication that it is the mutual arrangement of the molecules in

Time-Resolved Fluorescence in Discotic Materials

Figure 10. Photoluminescence spectra (after about 2 min of irradiation)


of decreasing concentrations of compound 4 in PMMA matrix: 25 wt
% (;), 5 wt % (- - -), 0.3 wt % (); leads to suppression of the
lower-energy emission.

J. Phys. Chem. B, Vol. 105, No. 20, 2001 4601


To obtain additional results concerning the special nature of
the fluorescence centers, we tried to perform annealing experiments. In the case of photophysical relaxation, heating to the
isotropic phase should lead to the destruction of the favored
spatial arrangement, so the emission properties of a nonirradiated
sample should be restored. Unfortunatelly, such annealing gives
rise to structural modifications in the bulk material: dewetting
takes place, and the molecules are partially reoriented toward a
homeotropic state. Because of the corresponding changes in the
transition probabilities and because of the onset of scattering
effects, we were not able to distinguish beyond any doubt
between photophysical and photochemical effects. Yet, the
modification of the spectra observed seems to suggest a
photochemical effect of still unknown nature.
In any case, it seems to be the peculiar spatial structure within
the plastic columnar state that is the origin of this effect.
Acknowledgment. We gratefully acknowledge the financial
support of the Deutsche Forschungsgemeinschaft (SFB 383).
References and Notes

Figure 11. Working model: the monomer exciton is transferred along


the column to the photoinduced dimers, which can emit light of a lower
energy.

the columns already in the ground state that gives rise to the
optical effect described here.
We thus speculate that the presence of excited monomers
induces structural relaxations within the plastic columnar state,
leading to a more ordered mutual configuration or a photochemical linkage of possibly two monomers, which, in turn,
can serve as nuclei for excimer formation. The isosbestic point
discussed above indicates a transformation of more or less
uncoupled monomers to such specific pairs. An excited monomer will transfer energy to the dimer, which, in turn, emits light
(Figure 11). Because the decrease with time of the intensity of
the higher-energy band reflects the decrease of the occupation
of its excited state, the expectation is that the lower-energy
emission grows on the same time scale. In our case, we have to
take into account the facts that, within the columns, very fast
singlet exciton hopping processes (30 fs-1 ps range)26,30 occur
and that the lifetime of the lower-energy emission is at least a
factor of 10 higher than that of the other emission. This means
that, as long as there are monomer excitons (indicated by the
existence of the higher-energy emission), the lower-energy
excited states are permanently pumped, causing a constant
occupation of all of those excited states and, therefore, a constant
emission intensity in time as obvious from our observation
(Figure 8).
It is furthermore apparent that the additional supermolecular
structure displayed by derivative 4 does not affect the energy
transfer.

(1) Richert, R. In Disorder Effects On Relaxation Properties; Richert,


R., Blumen, A., Eds.; Springer-Verlag: Berlin, 1994; p 333.
(2) Mott, N. F.; Davies, E. A. Electronic Processes in Non-Crystalline
Materials; Clarendon Press: Oxford, U.K., 1971.
(3) Bassler, H. Phys. ReV. Lett. 1987, 58, 767.
(4) Bassler, H. In Disorder Effects On Relaxation Properties; Richert,
R., Blumen, A., Eds.; Springer-Verlag: Berlin, 1994; p 485.
(5) Pope, M., Swenberg, C. E., Eds. Electronic Processes in Organic
Crystals; Clarendon Press: Oxford, U. K., 1982.
(6) Stevens, B.; Hutton, E. Nature 1960, 186, 1045.
(7) Lemmer, U.; Heun, S.; Mahrt, R. F.; Scherf, U.; Hopmeier, M.;
Siegner, U.; Gobel, E. O.; Mullen, K.; Bassler, H. Chem. Phys. Lett. 1995,
240, 373.
(8) Chandrasekhar, S. Liquid Crystals, 2nd ed.; Cambridge University
Press: Cambridge, U.K., 1992.
(9) Stegemeyer, H. Liquid Crystals; Springer-Verlag: New York, 1994.
(10) Adam, D.; Schuhmacher, P.; Simmerer, J.; Haussling, L.; Siemensmeyer, K.; Etzbach, K. H.; Ringsdorf, H.; Haarer, D. Nature 1994, 371,
141.
(11) Adam, D.; Closs, F.; Frey, T.; Funhoff, D.; Haarer, D.; Ringsdorf,
H.; Schuhmacher, P.; Siemensmeyer, K. Phys. ReV. Lett. 1993, 70, 457;
AdV. Mater. 1993, 7, 276.
(12) Simmerer, J.; Glusen, B.; Paulus, W.; Kettner, A.; Schuhmacher,
P.; Adam, D.; Etzbach, K. H.; Siemensmeyer, K.; Wendorff, J. H.;
Ringsdorf, H.; Haarer, D. AdV. Mater. 1996, 8, 815.
(13) Christ, Th.; Glusen, B.; Greiner, A.; Kettner, A.; Sander, R.;
Stumpflen, V.; Trukruk, V.; Wendorff, J. H. AdV. Mater. 1997, 9, 48.
(14) Christ, Th.; Stumpflen, V.; Wendorff, J. H. Macromol. Rapid
Commun. 1997, 18, 93.
(15) Stapff, I.; Stumpflen, V.; Wendorff, J. H.; Spohn, D. B.; Mobius,
D. Liq. Cryst. 1997, 23, 4, 613-617.
(16) Boden, N.; Movaghar, B. In Handbook of Liquid Crystals; WileyVCH: Weinheim, Germany, 1998.
(17) Kettner, A. Ph.D. Thesis, Philipps University, Marburg, Germany,
1998.
(18) Markovitsi, D.; Lecuyer, I.; Simon, J. J. Phys. Chem. 1991, 95,
3620-3626.
(19) Boden, N.; Bushby, R. J.; Clements, J.; Luo, R. J. Mater. Chem.
1995, 5, 1741.
(20) Markovitsi, D.; Rigaut, F.; Mouallem, M.; Malthete, J. Chem. Phys.
Lett. 1987, 135, 236.
(21) Glusen, B.; Heitz, W.; Kettner, A.; Wendorff, J. H. Liq. Cryst. 1996,
20, 627.
(22) Glusen, B.; Kettner, A.; Kopitzke, J.; Wendorff, J. H. J. NonCrystalline Solids 1998, 241, 113.
(23) Marguet, S.; Markovitsi, D.; Millie, Ph.; Sigal, H.; Kumar S. J.
Phys. Chem. B 1998, 102, 4697-4710.
(24) Bayer, A.; Kopitzke, J.; Noll, F.; Wendorff, J. H. Macromolecules,
manuscript accepted.
(25) Fimmen, W.; Kettner, A.; Kopitzke, J.; Ruland, W.; Wendorff, J.
H. Tagungsband der 25. Freiburger Arbeitstagung: Freiburg, Germany,
1999.
(26) Markovitsi, D.; Lecuyer, I.; Lianos, P.; Malthete, J. J. Chem. Soc.,
Faraday Trans. 1991, 87 (11), 1785-1790.

4602 J. Phys. Chem. B, Vol. 105, No. 20, 2001


(27) Klessinger, M.; Michl, J. In Excited States and Photochemistry of
Organic Molecules; VCH: New York, 1995.
(28) Schartel, B.; Kettner, A.; Kunze, R.; Wendorff, J. H. AdV. Mater.
Opt. Electron. 1999, 9, 55-64.

Bayer et al.
(29) Fimmen, W.; Glusen, B.; Kettner, A.; Wittenberg, M.; Wendorff,
J. H. Liq. Cryst. 1997, 23, 569.
(30) Sigal, H.; Markovitsi, D.; Gallos, L. K.; Argyrakis, P. J. Phys. Chem.
1996, 100, 10999-11004.

Você também pode gostar