Vous êtes sur la page 1sur 20

Encyclopedia of

Nanoscience and
Nanotechnology

www.aspbs.com/enn

Optical Properties of
Oligophenylenevinylenes
Johannes Gierschner, Dieter Oelkrug
University of Tbingen, Tbingen, Germany

CONTENTS
1. Introduction
2. Materials
3. Optical and Photophysical
Properties in Solution
4. Model Compounds for OPVOPV
Interaction Studies
5. Optical and Photophysical Properties
in the Solid State
Glossary
References

1. INTRODUCTION
In the last 20 years, polyene-like organic molecules have
attracted much attention as active components in organic
light-emitting diodes (OLEDs), microlasing cavities, ultrafast photoswitches and detectors, as well as organic eld
effect transistors (FETs) [1]. Among the varieties of materials, polyphenylenevinylene (PPV)-based materials are widely
investigated due to their use in OLEDs [2]. In order to
improve the optical and photophysical properties, that is,
tunable emission wavelengths and high uorescence quantum yields, the interplay of intramolecular effects (the
effective conjugation length and band shifts by chemical substitution) and intermolecular interactions (mutual
geometrical orientations, electronic and vibrational coupling between the chromophores) has to be understood.
Oligophenylenevinylenes (OPVs, see Fig. 1) make possible
the systematic study of these parameters, since the oligomers
provide well-dened conjugation lengths, variable substitution patterns, and a variety of intermolecular orientations in
the condensed phases.
OPVs have been studied since the 1960s as scintillators
[3], laser dyes [4], and optical whiteners [5]. The photochemistry of stilbenoid compounds, especially the transcis
ISBN: 1-58883-064-0/$35.00
Copyright 2004 by American Scientic Publishers
All rights of reproduction in any form reserved.

photo-isomerization of stilbene, has been extensively analyzed [6] and is still under investigation [712]. Since the
early 1990s, with the synthesis of homologous series of soluble t-butyl-substituted OPVs by Schenk and co-workers [13],
the oligomers have become a widely used model system for
PPV. Since then, a great variety of substitution patterns has
been established, allowing the tuning of the emission wavelength over the visible range [14, 15]. Quantum chemical
investigations and theoretical modeling of the OPVs permit
the precise prediction of the electronic transition energies
and spectra of the oligomers in solution and extrapolation
to the ideally conjugated polymer. Experimental and theoretical studies of intermolecular orientations in different
model systems enlightened the correlation of structural and
optical properties in the condensed phases and opened the
opportunity to achieve full control of the photophysics of
these systems. In the recent years, lms, single crystals, dendrimer systems, and hostguest compounds of the OPVs
have attracted growing attention for potential applications
in optoelectronic devices due to their electroluminescent
[1623], lasing [24, 25], (photo) conductive [2631], or photovoltaic properties [32, 33]. Substituted OPVs with large
two-photon absorption cross sections are attractive materials for uorescence microscopy, optical limiting, and optical
data storage [34].
The article focuses on the photophysics of individual and
condensed para-oligophenylenevinylenes and their correlation to intra- and intermolecular structural effects. In Section 2, an overview of the materials is given, including the
variety of substitution patterns and the geometrical structures of the molecules. Section 3 discusses the features of the
molecules in solution (absorption and uorescence spectra,
electronic transition energies, uorescence quantum yields,
and decay times) and the inuence of mesomeric and inductive substituent effects, solvent shifts, and thermal effects.
In Section 4, model compounds for the investigation of
intermolecular interactions and their impact on the optical
properties are discussed. In Section 5, the photophysics of
condensed phases such as lms, nanoparticles, and single
crystals of OPVs are addressed.
Encyclopedia of Nanoscience and Nanotechnology
Edited by H. S. Nalwa
Volume 8: Pages (219238)

220

Optical Properties of Oligophenylenevinylenes

5
4

1
n

Figure 1. Structure and notation of the atoms for oligophenylenevinylenes (nPVs).

2. MATERIALS
2.1. Substitution Patterns
The homologous series of unsubstituted oligophenylenevinylenes, nPV, where n denotes the number of
phenylenevinylene units (see Fig. 1 and Table 1), has been
synthesized up to n = 7 [5, 27, 35, 36]. In order to enlarge
the solubility of the oligomers, alkyl or alkoxy groups
were introduced. The series of t-butyl-substituted oligomers
(BnPV, see Table 2) with n = 16 was synthesized by Schenk
et al. [13], and series of alkoxy-substituted OPVs were synthesized by Stalmach et al. (nPOPVs with n = 111, see
Table 3) [37] and Peeters et al. (nBOPVs, with n = 16, see
Table 4) [38]. Recently, the syntheses of further homologous
series with alkyl [3942] and alkoxy [16, 4043] substituents
was reported. Finally, homologous series of donoracceptor
(DA)-substituted OPVs were synthesized, with D,A substituents either in the terminal 5,5 positions (see Fig. 1, D,

dialkylamino, and A, cyano, carbonyl, nitro groups) [44] or


in the vinylene units (1,1 positions with D, dibutylamino, A,
cyano) [45].
The substitution pattern was widely varied, especially
on para-distyrylbenzene (2PV), including ring substitutions
(Tables 5, 6) with alkyl [13, 26, 36, 39, 4649], alkoxy [16,
26, 3638, 43, 4953], cyano [36, 49, 5457], amino [34, 49,
58], nitro [26, 49], uoro [5961], bromo [49], carbonyl [62],
carboxyl [49], sulfonyl [53], and silyl groups [63, 64] or push
pull systems [44, 53], as well as substitutions in the vinylene
unit (Tables 7, 8) with sulfonyl [50, 65, 66] or cyano groups
[48, 50, 6671]. A systematic variation of cyanosubstitution
was also performed on 4PV (Table 9) [7275]. Thus, different substitution patterns with mesomeric (M) and/or
inductive (I) groups, but also with varying sterical requirements, are now available. Recently, general procedures of
OPV syntheses were reviewed [14].

2.2. Molecular Structure


According to X-ray studies on trans-stilbene (1PV) [7679],
2PV [80], and 4PV [17], the unsubstituted OPVs are only
slightly twisted, with torsional angles around the C1 C2 single bonds (Fig. 1) of  5 . A perfect planar structure (C2h
symmetry) of 1PV is found in the gas phase at low temperatures according to jet-cooled spectra [8184]. The C2h
symmetry is also predicted by density functional (B3LYP) [8,
8588] and semiempirical PM3 [8992] quantum chemical

Table 1. Experimental and calculated adiabatic and vertical electronic transition energies (103 cm1 ) of unsubstituted oligophenylenevinylenes nPV
in vacuo.

1PV

Adiabatic transition energies 00

Experiment
Calculation

In vacuo
Ab initio

Vertical transition energies vert

Experiment e
Calculation

In vacuo
Ab initio

Semiempirical

TDFT
Semiempirical

Fluorescence

F
F / ns

2PV

3PV

4PV

Ref.

315
317
338
300

277
281
299
264

260
263
287
244

251
b
27.4 c
23.3 d

[91]
[91]
[246]
[97]

338
339
418
325
329
338
331
319

298
297
366
261
291
293
287
280

280
282
343
229
269
273
269
262

270
f
33.0 g
21.1 h
25.7 i
26.2 j
25.9 k
25.2 l

[91]
[91]
[246]
[245]
[97]
[91]
[243]
[107]

[148]
[148]

0036 m
009 m

Note: Fluorescence quantum yields (F ) and decay times (F ) in solution are shown.
a
Obtained from experiments in solution by extrapolation to a refractive index n = 1.
b
RCIS/6311G .
c
RCIS/321G.
d
AM1 CAS.
e
Obtained from the experimental adiabatic transition energies by addition of the equilibration energy exp .
f
From 00 (RCIS/6311G ) by addition of the explicitly calculated [91] equilibration energy calc .
g
RCIS/321G (direct calculation).
h
Time-dependent density functional B3LYP/631G .
i
AM1.
j
ZINDO/S-CI, involving (2n + 1) occupied and unoccupied molecular levels.
k
ZINDO/S-CI (full conguration interaction).
l
PPP.
m
[278].

090
120

085
110

221

Optical Properties of Oligophenylenevinylenes

Table 2. Absorption maxima (


max ), oscillator strengths (f ), uorescence subbands (
1
2 ), uorescence quantum yields
(F ), uorescence decay times (F ), and radiative rate constant (kF ) of t-butyl-substituted oligophenylenevinylene BnPVs
in solution (dichloromethane) at T = 293 K. [175].

n1

Absorption

Fluorescence

max /nm
B1PV
B2PV
B3PV
B4PV
B5PV
B6PV

316
360
387
403
410
417

1
2 /nm

F

F /ns

kF /109 s1

344, 361
397, 419
428, 456
448, 478
459, 490
464, 495

010
086
082
075
071
071

110
106
097
073
065

078
077
077
097
109

053
089
121
160
180
188

Table 3. Absorption maxima (


max ), extinction coefcient (), uorescence maxima (
1
2 ), uorescence quantum yields
(F ), uorescence decay times (F ), and radiative rate constant (kF ) of propyloxy-substituted oligophenylenevinylene
nPOPVs in solution at T = 293 K.
OC3H7

OC3H7

Absorptiona
H7C3O

H7C3O

1POPV
2POPV
3POPV
4POPV
6POPV
8POPV
11POPV
a
b

Fluorescenceb

max /nm
[37]

max /103 l mol1 cm1


[37]

1
2 /nm
[283]

F
[283]

F /ns
[283]

kF /109 s1
[283]

354
401
431
450
466
475
481

17
41
59
85
117
146
196

(408), 425
460, 486
500, 529
522, 555
543, 580
549, 586
551, 590

045
082
094
070
043
048
048

189
160
172
110
062
062
064

023
051
055
064
070
077
075

In chloroform.
In dichloromethane.

Table 4. Absorption maxima (


max ), extinction coefcient (), uorescence maxima (
1
2 ), uorescence quantum yields
(F ), uorescence decay times (F ), and radiative rate constant (kF ) of methylbutoxy-substituted oligophenylenevinylene
nBOPVs in chloroform at T = 293 K [173].
O

O
Me
O

1BOPV
2BOPV
3BOPV
4BOPV
5BOPV
6BOPV

Me
n

Absorption

Fluorescence

max /nm

max /103 l mol1 cm1

max /nm

F

F /ns

kF /109 s1

357
407
437
454
464
475

21
44
68
88
107
142

409
466
502
523
534
541

062
076
049
041
025

170
132
073
052
045

037
058
067
080
055

222

Optical Properties of Oligophenylenevinylenes

Table 5. Absorption maxima (


max ), extinction coefcient (), uorescence maxima (
max ), uorescence quantum yields (F ), uorescence decay
tpa
times (F ), two-photon excitation maximum (
max ), and two-photon absorption cross section () of substituted distyrylbenzenes in toluene at T =
293 K.

R
N
R

n-Butyl
n-Butyl
n-Butyl
n-Butyl
n-Butyl
Phenyl
n-Butyl
a

H
F
OMe
SO2 C3 H7
CN
CN
H

R
N
R

One-photon absorption



H
F
H
H
H
H
H

max /nm max /10 l mol

H
H
H
H
H
H
CN

410
430
428
439
490
475
438

Fluorescence

cm

74
77
67
43
66
66
54

max /nm

F

Two-photon absorption

F /ns

455
456
480
528
536
528
504

088
135
058
079
088
138a
031
101
069
130
087
146
0003 <0015

tpa
max

/nm /1050 cm4 /s/photon

730
760
730
816
830
830
790

955
1400
900
4100
1750
1640
890

Ref.
[270, 268]
[271]
[268, 34]
[271]
[270]
[270]
[270]

In acetonitrile.

Table 6. Absorption maxima (


max ), uorescence maxima (
max ), uorescence quantum yields (F ), and uorescence
decay times (F ) of distyrylbenzene (2PV) derivates with substituents in the phenylene moieties.
R

H
OC8 H17
OC8 H17
OC8 H17
OC8 H17
OCH3
OCH3

Absorption

H
H
Ph
OPh
COPh
NPh2
H



H
H
H
H
H
H
OCH3

Fluorescence
a

max /nm

max /nm

355
387
405
392
415
425
393

388, 410
440
460
444
487
480
446, 470

F

F /ns

Solvent

Ref.

0.90
0.83
0.90
0.91
0.70
0.75
0.92

1.29
1.74
1.35
1.49
1.52
1.26
1.55

Dioxane
Dioxane
Dioxane
Dioxane
Dioxane
Dioxane
C6 H12

[48]
[48]
[48]
[48]
[48]
[48]
[69]

Note: Data in solution are shown.


a
For structured spectra the spectral positions of the two high energetic subbands are given.

Table 7. Absorption maxima (


max ), uorescence maxima (
max ), uorescence quantum yields (F ), and uorescence decay times (F ) of distyrylbenzene (2PV) derivates with substituents in the vinylene moieties.

R
R
R

Type
A
B

R
R
R

Absorption






OC3 H7
OC8 H17
OC8 H17
OCH3

H
H
OPh
H

H
H
H
OCH3

H
H
H
H

CN
CN
CN
CN

OCH3

OCH3

C6 H13

C6 H13
C6 H13

H
H

H
C6 H13

Fluorescence

max /nm

max /nm

H
H
H
H

425
422
428
408

CN

CN

H
C6 H13

H
CN

F

F /ns

489
512
511
478

0.36
0.49
0.06

410

445, 475

0.10

375

435, 458

CN
H

304
325

410
441

0.002
050.8
0.001
0.003

290
223
03 25%
10 75%
02 12%
09 88%
<001
15

Note: Data in solution are shown.


a
For structured spectra the spectral positions of the two high energetic subbands are given.

Solvent

Ref.

CHCl3
Dioxane
Dioxane
C6 H12

[50]
[48, 284]
[48, 284]
[69]

CH2 Cl2

[69]

CH2 Cl2
EPA, 77 K
CH2 Cl2
CH2 Cl2

[146]
[237]
[146]

223

Optical Properties of Oligophenylenevinylenes

Table 8. Fluorescence properties of (substituted) distyrylbenzenes (2PV) in solution, thin lms, and nanoparticles (NP): uorescence maxima (
max ;
us, unstructured), quantum yields (F ), decay times (F , ne, nonexponential), natural lifetimes (F0 ), and uorescence anisotropy (rF ).
R

m o Y
p
R

o m

R

State

max /nm

F

F /ns

F0 /ns

rF

H
H

HexO

030

J aggregates

HexO

CN

000

Excimer emission

HexO

CN

HexO

o-CN

HexO

m-CN

HexO

p-CN

p-CN

1.2
2.5 (n.e.)
912 (n.e.)
1.1
0.4
1.6
21 (n.e.)
0.9
2.2 (n.e.)
2.1
30 (n.e.)
2.0
9.5
1.9
2.3
1.8
0.51.0
1.2
1.3
0.3 (n.e.)
12 (n.e.)
20 (n.e.)

H aggregates

0.92
0.050.1
0.30.5
0.82
0.18
0.66
0.15

0.32
0.3
0.65
0.030.06
0.72
0.040.08
0.81
0.70.8
0.80
0.2
0.2
0.40.5
0.4

025

388, 409
(415), 441
(421), 448
403, 424
417, 440
446, 467
583 u.s.
497 u.s.
478 u.s.
520 u.s.
622 u.s.
487 u.s.
580 u.s.
460 u.s.
576 u.s.
500 u.s.
482, 514
409, 433
(430), 460
436, 463
560 u.s.
560 u.s.

1.3
30

Hex

Solution
Film/NP 293 K
15 K
Solution
Film
Solution
Film
Solution
Film
Solution
Film
Solution
Film
Solution
Film
Solution
Film
Solution
NP (methanol/H2 O)
Film, fast deposited
Film, annealed
NP (dioxane/H2 O)

1.3
2.2
2.4
140

6.6
100
3.1
200
2.6
40
2.2
1
1.5
6.5
15
30
50

Remarks

Ref.
[54]
[54]

Excimer emission

001

Excimer emission

025

Excimer emission

020

J aggregates

030
030

023

H aggregates
J aggregates
Excimer emission
Excimer emission

[54]
[54]
[237]
[54, 57]
[237]
[237]
[237]
[237]
[57]
[54, 57]
[57]
[54, 57]
[57]
[54]
[55]
[319]
[55]
[55]
[319]

Table 9. Fluorescence properties of (substituted) distyryl(distyrylbenzene)s (4PV) in solution, thin lms, nanoparticles (NP), and single crystals:
uorescence maxima (
max ; u.s., unstructured), quantum yields (F ), and decay times (F , mono- or biexponential ts).

Name

R

4PV

MEH-4PV

EtHexOa

MetO

Oct-4PV

Oct

OctO-4PV

OctO

OctO-4PV-CNo

OctO

CN

OctO-4PV-CNi
a

OctO

EtHexO is 2-ethyl-hexyloxy.

CN

State

max /nm

F

F /ns

Solution
NP
Solution
Film
Solution
Film
Solution
Film, as deposited
Film, annealed
Single crystal
Solution
Film, as deposited
Film, annealed
Single crystal
Solution
Single crystal

440, 471
(483), 515
480, 510
(500), 530
455, 485
505, 533
483, 515
529
539
539, (563)
537, (570)
595, (620)
615 u.s.
630 u.s.
510 u.s.
560, 593

01

0.15
0.850.9

0.70.9
0.5
0.7
0.5
0.70.9
0.4
0.6
0.5
0.04
0.4

0.65
1.2

1.0
1.1
1.7
1.0
1.4
1.4, 7.0
3.4, 8.0
8.0
0.08
2.2

Remarks
H aggregates

J aggregates
Cofacial x,z-shifted
Excimer emission
Cofacial z-shifted
Cofacial x,z-shifted

Ref.
[91]
[174]
[152, 154, 155]
[153, 320]
[305]
[303]
[25, 305]
[24, 306]
[24, 306]
[306]
[25, 110, 275]
[306, 110]
[306]
[110, 306]
[25]
[306]

224
calculations on 1PV. Ab initio HartreeFock (HF) [86, 87,
9295] and semiempirical AM1 [89, 90, 92, 93, 96, 97] calculations yield a slightly nonplanar structure (C2 symmetry) of 1PV; however, the ring-torsional potential is very at
for  < 30 . The bond lengths and angles of the nPVs, as
determined by X-ray measurements [17, 7680], are quite
well reproduced by theory, resulting in pronounced bond
alterations in the vinylene units [810, 8794, 98103]. In
the gas phase at higher temperatures, the average geometrical structure, according to electron diffraction [104]
and photoelectron studies [105, 106] of 1PV, is nonplanar, due to the low torsional barrier for rotations around
the ethenylphenyl (C1 C2 ) single bonds [8, 85, 8993,
9597, 107].
Deviations from planarity in the solid state are observed
in several substituted OPVs. According to quantum chemical
calculations [69, 92, 93], the deviations from planarity
depend sensitively on the kind and positions of substituents
[93]. However, the torsional angles in most of the reported
X-ray structures [17, 47, 60, 73, 74, 80, 108114] do not
exceed 11 18 . Larger deviations from planarity are found
for 2PV derivatives with cyanosubstituents in the vinylene
moiety and additional alkyl substituents at the central phenyl
ring (see Tables 79) [68].

Optical Properties of Oligophenylenevinylenes


'Shifted' arrangements

CMe3

1a

2a

Me3C

Ref. [118, 119, 121]

'Angled' arrangements
1b

2b

CMe3
'Crossed' arrangements
Me3C

Me3C

Ref. [119, 121]


Me3C

2.3. Bridged Oligophenylenevinylenes


Oligophenylenevinylenes with dened mutual geometric
alignment are represented by (i) hydrogen-bonded OPV
pairs, introduced by Meijer and co-workers [115117], and
by (ii) covalently linked oligomers, which were synthesized
by Schenk et al. [13], Anger et al. [11], Rau and Waldner [12], Bazan [118129], and Song et al. [130133]. The
covalently linked OPVs cover dimeric stilbenoid cyclophanes
with partial  overlap and different intermolecular orientations [13, 118125, 134] (Fig. 2), stilbenophanes with perfect face-to-face alignment [11, 12, 135] (Fig. 2), tetrameric
systems with tetrahedral OPV orientation [126128, 136]
(Fig. 3), and dendrimers [128, 136141]. Dimeric OPVs were
also prepared by Lewis and Letsinger using 1:1 mixtures
of complementary oligonucleotides containing stilbene units
[142145].

2.4. Condensed Phases


Films of OPVs were prepared by vapor deposition from high
vacuum [20, 28, 74, 146155], by spin or drop coating from
solution [21, 23, 156158], and by the LangmuirBlodgett
technique [131, 159170]. Nanoparticles were obtained by
fast precipitation from solutions of the oligomers by addition of a poor solvent (i.e., water) and measured in suspension [38, 54, 116, 146, 148, 170178]. The size of the
nanoparticles (20500 nm) can be controlled by the preparation conditions (concentration, solvent, temperature, mixing
ratio). Nanoparticles were also prepared by cooling saturated solutions of the OPVs [62, 179182]. General procedures for nanoparticle preparation were reviewed recently
[183, 184]. Single crystals were grown from solution [60, 61,
68, 7578, 108112, 163, 185], melt [186], and gas phase
[31, 79, 187, 188].

CMe3

CMe3
'Face-to-Face' arrangements

Ref. [120]
S

S
Ref. [11, 135] S

4a

S
4b

Ref. [12]

Figure 2. Structures of stilbenoid cyclophanes.

Hostguest compounds (HGCs), where the OPVs are


incorporated in an inert host material such as -cyclodextrin
[189] or channel-forming perhydrotriphenylene (PHTP) [55,
190, 191], were prepared by cocrystallization from solution.
The cavities of the cyclodextrin HGCs permit the inclusion
of cofacially oriented stilbene pairs. PHTP HGCs open the
opportunity to investigate specic head-to-tail interactions
of the oligomers.

R
R

AdR4
R
R

CR4
R
R

R
R = t-butyl-2PV [126128]

R = t-butyl-2PV [126, 128]


R = alkoxy-2PV [128]

Figure 3. Structures of tetrahedral oligophenylenevinylenes.

225

Optical Properties of Oligophenylenevinylenes

3. OPTICAL AND PHOTOPHYSICAL


PROPERTIES IN SOLUTION
3.1. Absorption and Fluorescence Spectra
At low temperatures (T < 20 K), the S0 S1 absorption
and S1 S0 uorescence spectra of 1PV [192] and the
longer oligomers [91] show very small 00 bandgaps and
the spectra are almost mirror symmetrical (Fig. 4). Thus,
the frequencies and nuclear displacements of the totally
symmetrical ag vibrational modes which couple to the electronic transition must be very similar in S0 and S1 , in agreement with ab initio quantum chemical calculations [91].
Detailed vibrational analyses of the electronic spectra were
carried out on (substituted) trans-stilbenes under collisionfree conditions [8184, 193196]. The ne structures of the
experimental spectra are dominated by the progressions
and combinations of a few prominent modes, which also
becomes evident for the longer oligomers in solid solution

2PV

a) experiment
(T = 15 K)

fluorescence intensity

b) calculated
in vacuo

c) experiment
(T = 293 K)

d) simulated

[91, 118, 179, 197, 198]. Theoretical calculations, obtained by


a FranckCondon approach [85, 91, 99, 100, 199, 200] from
ab initio quantum chemical calculations, demonstrate that
the nuclear displacements of the prominent modes essentially coincide with the geometrical changes upon the electronic transition. Among these modes, the low-frequency
longitudinal in-plane deformation mode ( = 204 cm1 for
1PV [193]) strongly decreases with n, resulting in  0 for
n [91], whereas the vinylene C1 C1 stretching mode
around 1650 cm1 [201208] decreases only very little with
n [96, 209211]. The optical spectra of the nPVs were calculated at different levels of theory for 1PV [85, 99, 100,
212] and the longer oligomers [91, 200, 213], and good
agreement with the experiment was obtained at the ab initio
HF/restricted conguration interaction singles (RCIS) 6
311G level, including the progressions and combinations of
the complete sets of ag modes (Fig. 4).
With increasing temperature, the mirror symmetry of
absorption and uorescence is lost and a 00 bandgap opens,
which extends to 200 cm1 at 77 K [173, 175, 214] and to
1000 cm1 at room temperature (Fig. 4). The room temperature uorescence spectra still exhibit distinct vibronic
structure, whereas the absorption spectra are only slightly
structured and asymmetrically shifted to the blue [91, 96,
146, 210, 214219] (see Fig. 4). The differences in the spectral characteristics of uorescence and absorption are due
to the different torsional barriers of the ground and excited
state structures. Compared to the S0 state, the C1 C2 bond is
rigidied in the S1 state [8, 9, 8991, 95, 100103, 220], leading to a steeper potential of the torsional modes [8, 90, 91,
97, 212]. The torsional barrier can be investigated in the gas
phase, where the low-frequency torsional mode of 1PV was
determined to  = 8 cm1 in the S0 state and  = 48 cm1 in
the S1 state [81]. Therefore, the room temperature absorption spectra arise from superimposed transitions of different
nonplanar conformers of the molecules [91, 96, 146, 215,
216, 221], and the total band shapes are well described by a
combination of a Gaussian-shaped and an exponential convolution of the low temperature spectra (Fig. 4). The Gaussian part accounts for inhomogeneities of the environment,
and the exponential part accounts for the increase in the torsional potential in S1 against S0 [212]. Since the potential in
S1 is high against the thermal energy, the room temperature
uorescence spectra can be exclusively tted by the Gaussian contribution to line broadening. In rigid environments,
where the potential in S0 is also high against the thermal
energy, the absorption spectrum can also be reasonably tted by a Gaussian convolution [178, 222].

3.2. Electronic Transition Energies


20000

25000

30000

/ cm1
Figure 4. Fluorescence emission (left) and excitation (right) spectra of
distyrylbenzene (2PV). (a) Experimental spectra in tetradecane at T =
15 K. (b) Calculated low temperature spectra (HF/RCIS 6311G ) [91].
(c) Experimental spectra in dioxane at T = 293 K. (d) Simulated spectra, obtained by convolution of the experimental low temperature spectra with a Gaussian and an exponential distribution [91].

The energetic positions of the absorption spectra of the


OPVs are strongly affected by the conjugation length and
to a smaller extent also by the solvent and the temperature (Fig. 4) [36, 223] due to changes in the polarizability of the environment [224226]. In nonpolar solvents, the
polarizability shift becomes a linear function of the Onsager
relation [224], which allows the extrapolation to the spectral
positions in vacuo with a refractive index of nsolv = 1 [91].
The solvent red shift amounts to R = 2500 cm1 compared to the values in vacuo (Fig. 4); thus solvent effects

226
should be considered if experimental transition energies are
compared to calculated ones. Specic solvent interactions in
polar solvents are observed for OPVs with electron withdrawing substituents like cyano, nitro, triyl (SO2 CF3 ), or
uoro groups [59, 75] and for donoracceptor-substituted
OPVs [225227], resulting in unstructured, red-shifted uorescence spectra and large Stokes shifts between the absorption and uorescence maxima due to the contribution of
charge separation not only in the S1 , but also in the S0 state
[228].
Vertical transition energies vert for the S0 S1 transition were calculated at different levels of theory for 1PV
[85, 88, 89, 99, 100, 229233], substituted OPVs [8, 16, 55,
60, 67, 92, 119, 232239], and the homologues series of the
nPVs [91, 97, 101, 103, 107, 218, 234, 240251]. The latter reproduce the experimental result of an inverse chain
length dependence 1/N (with N = 3n + 2) of vert [91, 96,
173, 216, 237, 247, 248]. However, the results obtained by
semiempirical methods depend sensitively on the number of
electron congurations considered in the calculations [89].
Time-dependent density functional (TDFT) methods tend
to underestimate the vertical transition energies (Table 1)
and to overestimate the slope of the 1/N dependence [245].
Reliable results were obtained by the ab initio RCIS/6
311G method, where the vertical transition energies were
calculated from adiabatic transition energies 00 by addition
of the equilibration energy [91] (see Table 1). In contrast,
the direct calculation of the vertical transition energies by
ab initio methods [246] yields unreasonably high values for
vert (see Table 1). For longer oligomers (n > 4), a deviation from the linear 1/N dependence is observed both
experimentally [27, 37, 173, 175] and theoretically [234, 246,
247], which results in a faster convergence for n .
The experimental transition energies can be reasonably tted by empirical functions [44, 246], within the extended
Hckel Molecular Orbital [252] and free electron gas model
[253] including bond alternation, Simpsons exciton model
[254], or the classical oscillator approach by Kuhn [255]. The
extrapolation to idealized PPV yields 00 20,300 cm1 in
solution (23,400 cm1 in vacuo). The equilibration energy
for PPV was calculated to vert 00 1750 cm1 [91].
The introduction of substituents in the terminal phenyl
rings with +M effect (e.g., alkoxy [16, 215, 223, 256] or
amino [34, 49, 58] groups), M effect (e.g., carbonyl [62]
or cyano [36, 49, 5457] groups), or +I effect (e.g., alkyl
groups [39, 96, 107, 175, 216, 218, 223, 257, 258]) shifts the
S0 S1 transitions slightly to lower energies. The additional
introduction of alkoxy substituents [18, 38, 43, 48, 51, 69,
173, 234, 237] in the central phenyl rings (nPOPVs, Table 3,
nBOPVs, Table 4) leads to a splitting of the low energetic
absorption band with a systematic bathochromic shift of the
S0 S1 transition compared to that of the nPVs. According
to quantum chemical calculations, this effect is ascribed to
the interaction between carbon and oxygen  orbitals, destabilizing the highest occupied molecular orbital (HOMO)
more strongly than the lowest unoccupied molecular orbital
(LUMO) [238, 259]. In addition, a blue-shifted RO band
appears, due to strong conguration interactions between
the HOMO and phenyl orbital branches [234]. Extrapolation of the nBOPV spectral positions to n yields

Optical Properties of Oligophenylenevinylenes

00 17,600 cm1 for the idealized polymer (experimentally:


00 (BOPPV) = 18,100 cm1 [260]). The uorescence and
absorption spectra of OPVs with cyanosubstituents in the
vinylene unit are strongly red-shifted against the unsubstituted ones (Tables 79), in agreement with quantum chemical calculations [238, 259, 261]. Terminal donoracceptor
substitution of 1PV, for example, 4-(dialkylamino)-4 -nitrostilbene [122, 225227], causes a large bathochromic shift
in the transition energies due to the intramolecular charge
transfer (CT) character of the transition. These results
are also well reproduced by quantum chemical calculations
[239, 247]. With increasing chain length n, the extent of
the CT character decreases [44, 239, 247]. Thus, for some
donoracceptor-substituted homologous series, the decrease
in the CT character overcompensates the bathochromic shift
induced by the increase in n, thus causing a hypsochromic
shift with n [44].
According to polarized absorption [262] and uorescence
studies [148, 263, 264] in stretched polyethylene lms and to
the rotational resolved uorescence excitation spectrum of
1PV under collision-free conditions [84], the S0 S1 transition moment # of the nPVs is oriented approximately parallel to the long geometrical axes of the molecules (dened
as the distance between the terminal carbon atoms d% % ).
The experimental results are supported by quantum chemical calculations, [9, 218, 229, 265, 266], yielding for 2PV
an angle of approximately 8 between d% % and # [266].
The oscillator strengths increase roughly linearly with the
chain length (Tables 24). The energetic positions, orientations, and oscillator strengths of the higher S0 Sn transitions of the OPVs were assigned by polarized absorption
spectroscopy on 1PV [262, 263] and uorescence excitation
measurements (1PV [265, 267], BnPVs [218]) and correlated
with quantum chemical calculations [9, 218, 229, 265, 272].
Two-photon absorption studies allowed the assignment of
the weak S0 S2 transition in 1PV [267] and substituted
OPVs [270, 268] (Table 5) located about 4000 cm1 above
the S0 S1 transition, in qualitative agreement with quantum chemical calculations [270, 97, 232, 269]. Substituted
nPV derivatives with donordonor, donoracceptordonor,
and acceptordonoracceptor structural motifs exhibit large
two-photon absorption cross sections,  [34, 268, 270,
271], in combination with high uorescence quantum yields
(Table 5). According to quantum chemical calculations,
the magnitude of  can be correlated with the degree
of intramolecular charge transfer from the terminal donor
groups to the  bridge [34, 268] and the calculated  values
are in reasonable agreement with the experimental ones [34,
286, 271].
Triplet T1 Tn transitions of nPVs (n = 1 [273], n =
2 [221]) alkoxy OPVs [173, 274] and differently substituted 4PVs [275] were determined by photoinduced absorption spectroscopy and by pulse radiolysis, respectively. The
transition energies decrease with increasing chain length
[173, 257], in agreement with quantum chemical calculations
[101, 103], whereby the decrease is somewhat stronger than
the one in the S0 S1 transition [173]. The equilibration
energies of the triplet transition decrease rapidly with chain
length n [173] in agreement with theoretical results [101].
The triplet lifetimes at T = 100 K are in the range of 38 ms
[173].

227

Optical Properties of Oligophenylenevinylenes

3.3. Fluorescence Quantum Yields


and Decay Times

4.1. Arrangements with Weak


Dipolar Coupling

The uorescence quantum yields F of the parent nPVs and


their derivatives with substituents at the phenylene moieties
are fairly high (Tables 26) [16, 36, 48, 54, 59, 60, 66, 69,
146, 175, 237, 276, 277]. An exception is 1PV (trans-stilbene)
due to the crossing of the S1 potential surface with the S2
state in the twisted geometry, leading to efcient nonradiative decay and transcis photoisomerization [6, 278]. Consequently, F (1PV) strongly increases in highly viscous [279]
or rigid environments [190], at low temperatures [279281],
and in stiffed trans-stilbenes [7, 279, 282], where the exibility of 1PV is strongly reduced. The radiative rate constants
kr for the S1 deactivation of the OPVs in solution (Tables
24) increase with increasing conjugation length n, saturating at kr = 1 03109 s1 [146, 173]. The increase in the
nonradiative rate constant with n is even stronger, due to
the increase in torsional phonon state density with decreasing S1 S0 transition energy, leading to an overall decrease
in F with n (Tables 24) [16, 173, 175, 283].
Compared to the alkyl- and alkoxy-substituted OPVs,
the uorescence quantum yields are signicantly lowered
for OPVs with cyanosubstituents in the vinylene unit.
The reduction is a sensitive function of the additional
substitution patterns (Table 7). Medium quantum yields
(F = 030.5) are observed if the central phenyl ring of 2PV
is additionally substituted with alkoxy groups (molecules of
Type A, see Table 7) [48, 50, 284]. Further alkoxy substitution at the terminal phenyl rings reduces F to 510%
(Type B) [69, 146]. Finally, the oligomers become practically nonuorescent if the central phenyl ring is substituted
with alkyl groups (Type C) [146, 237]. Fluorescence quenching of these compounds is mainly a sterical effect. While
molecules of Type A are almost planar [74], the structures
of Type C molecules are far from planarity in the S0 state
[68], thus enhancing torsional induced nonradiative deactivation [146]. Rigid environments enforce planarization of
the C-type molecules in the S1 state; thus F increases by 2
orders of magnitude in EPA glass at 77 K [146] (see Table 7).

In order to investigate specic head-to-tail interactions,


HGCs were prepared, where OPVs and their p,p substituted derivatives are incorporated in the channels of
the pseudo-hexagonal host lattice of PHTP [55, 190, 191].
The intrachannel distances of the oligomers are given by the
long axis dimensions and the van der Waals distances. The
interchannel distances are in the order of ra b 15 nm [190,
286], so that side-by-side interactions between the oligomers
are suppressed. The ne structures and spectral positions of
the uorescence spectra are very similar to those observed
in solution [55, 190, 191], thus indicating only small dipolar OPVOPV interactions. The observed red shift of the
spectra R 700 cm1 against solution is partly due to the
high polarizability of the host lattice, and only a portion of
R 300 cm1 can be attributed to OPVOPV interactions
[55, 191]. The intermolecular excitation transfer of energy
(ET) in the HGCs was investigated by time-resolved uorescence on doped systems [191]. The results were interpreted
according to the Frster-type ET mechanism, considering
homo- and heterotransfer as well as intra- and interchannel
transfer steps [191, 287].
The spectral features of the OPV tetramers in Figure 3
are also comparable with those of solutions of the monomers
[126128]. The spectra remain structured and are red-shifted
by R 350 cm1 against the monomers. This also indicates that in these compounds the interactions between the
OPV units are only of weak dipolar character. However,
very fast ET is observed. The monoexponential uorescence
anisotropy decay of AdR4 can be reasonably tted assuming
a Forster-type energy transfer, again indicating a weak coupling of the OPV units [128]. For CR4 a biexponential decay
was observed, which may suggest the presence of short-range
exchange (Dexter)-type interaction [128].

4. MODEL COMPOUNDS FOR OPVOPV


INTERACTION STUDIES
Intermolecular interactions between OPV chromophores
and their impact on optical and photophysical properties
can be studied on model compounds, which are formed
of a small number of oligomers with dened intermolecular geometrical alignment. Simple tetrameric structures
were obtained with 2PV derivatives, where the oligomers
are covalently linked via a central sp3 carbon atom or via
adamantan (Fig. 3) [126128]. One-dimensional head-to-tail
alignments could be realized in channel-forming hostguest
compounds [55, 190, 191, 285]. Since in these structures the
OPVOPV distances are large, the interactions between the
oligomers are only of weak dipolar character. An increase in
interactions is expected for arrangements with  overlap,
represented by stilbenoid cyclophane compounds [1113,
118124, 134] and 1:1 mixtures of complementary oligonucleotides containing stilbene units [142144].

4.2. Arrangements with  Overlap


Stronger interactions are expected in OPV pairs with partial
 electron overlap of the OPV units, as in cyclophanes
[118125]. For the shifted and angled arrangements of
the 2PV cyclophanes (structures 2a, 2b in Fig. 2), a red
shift of R = 1100 cm1 against the spectra of the parent monomer unit (2c in Fig. 5) is observed [119, 121]. The
intermolecular coupling is still of dipolar character, since the
vibronic features of the room temperature spectra as well
as the uorescence quantum yields and decay times [119]
are very similar to those of the monomer. A monomer-like
emitting state was also predicted by theoretical calculations
on the basis of the collective electronic oscillator (CEO)
method [119]. However, differences between 2a and the
monomer become apparent at low temperatures. In contrast
to the monomer, well-resolved phonon lines could not be
observed in the uorescence spectrum of 2a, probably due to
an increase in the FranckCondon activity in low-frequency
chain deformation modes [118].
Signicant changes in the optical properties are observed
if the relative extent of  overlap between the oligomers
is enhanced. Thus, the shifted and angled arrangements of
1PV cyclophanes [119, 134] (structures 1a, 1b in Fig. 2)

228

Optical Properties of Oligophenylenevinylenes

0.3

1 108

0.2
5 107
0.1

400

300

200

100

0
2a
2b
2c

0.8

1.5 109

0.6
1 108
0.4
5 107

0.2
0

Emission Intensity (c.p.s.)

Normalized Absorbance (a.u.)

1.5 109

Absorbance
rel. Intensities

0.4

500

Emission Intensity (c.p.s.)

Normalized Absorbance (a.u.)

2 108
1a
1b
1c

0.5

0
250

300

350

400

450

500

Wavelength (nm)
Figure 5. Absorption and uorescence spectra of stilbenoid cyclophanes. (Top) Spectra of compounds 1a, 1b (see Fig. 2) and of the
parent monomer unit (1c). (Bottom) Spectra of compounds 2a, 2b (see
Fig. 2) and of the parent monomer unit (2c). Reprinted with permission
from [121], W. J. Oldham Jr. et al., J. Am. Chem. Soc. 120, 419 (1998).
1998, American Chemical Society.

both show structureless, strongly red-shifted emission bands


(Fig. 5) with low radiative rate constants [119], reminiscent
of excimer emission. According to CEO calculations the
emission of the 1PV cyclophanes occurs from a low-lying
state with a signicant contribution of the through-space
delocalized paracyclophane core [119]. The main absorption
band of 1b shows a splitting of about 9700 cm1 , in good
agreement with CEO calculations, due to the Davydovlike splitting of the S0 S1 transitions of the monomer
units in the angled arrangement [119]. An equivalent explanation accounts for the properties of the crossed 2PV
cyclophane [120, 134] (structure 3 in Fig. 2), where the 
overlap of the monomer units is increased against 2a and
2b, respectively.
Perfect face-to-face alignment in 1PV cyclophane dimers
was realized in the dimers 4a and 4b of Figure 2. In contrast to the stilbenophane 4a, which can form ZZ, ZE,
and EE isomer pairs [11] due to the exibility of the ethyl
linkers, 4b is unable to undergo EZ isomerization [12].
The main absorption band of 4b shows two components, an
intense one at
1 300 nm, slightly blue-shifted against the
monomer peak at
= 308 nm, and a red-shifted weak one
at
2 350 nm (Fig. 6). Due to the strong  overlap of
the stilbene units, the uorescence mimics excimer emission
with an even stronger red-shifted emission spectrum than 1a
and 1b and a uorescence lifetime of F = 11 ns [12]. Similar

300

400

500

600

Wavelength/nm

Figure 6. Absorption (dots), uorescence excitation (solid line), and


emission (dashed line) spectra of stilbenophane 4b. Reprinted with permission from [12], H. Rau and I. Waldner, Phys. Chem. Chem. Phys. 4,
1776 (2002). 2002, Royal Society of Chemistry.

features are also observed for 1:1 mixtures of complementary oligonucleotides containing stilbene units [142144] and
ternary complexes of 1PV/ -cyclodextrin/cyclohexane [189].
The electronic properties of cofacial nPV dimers were
theoretically treated by Cornil [242, 288290] and by Tretiak
and Mukamel [291] and Tretiak et al. [292] as a function
of the interchain distance. For large interchain separations
(d  5 ), a symmetrical splitting of the rst single chain
transition into a dipole allowed higher energetic components
and a forbidden lower one was calculated (H-type interaction), in agreement with Kashas molecular exciton model
[293, 294]. The excitation energy is still localized on a single chain. Below a critical distance of d 5 the splitting
becomes unsymmetrical due to second-order perturbations,
inducing an additional stabilization of the both states [242].
The lower state is delocalized over both chromophores,
whereas the upper state has small interchromophore coherence [292]. The splitting tends to decrease nonlinearly with
the chain length of the oligomers [242, 288, 295297], in
contrast to the predictions of the molecular exciton model,
which states that there is an increase with chain length [293].
The splitting for a cofacial 1PV dimer [242] at a distance of
d = 35 , which roughly corresponds to the average interchromophore distance in 4b [298], is  = 5800 cm1 at
the INDO/S-CI level, in reasonable agreement with experiments (see Fig. 6). Analogous to considerations on 2,2 paracyclophane [299], the symmetrically forbidden lower
excited electronic Bg state in 4b may borrow intensity from
the higher Bu state via an au vibrational mode. The weakly
allowed Bg state is responsible for the low radiative rate constant of 4b. The large red shift of the uorescence spectra
of 1a and 1b and particularly 4b is mainly due to strong
intermolecular vibronic coupling between the chromophore
units: Quantum chemical calculations of 2,2 -paracyclophane
[300] and 4b [298] predict a strong decrease in the interring
distance after S0 S1 excitation, which results in extremely
high FranckCondon activity of interring breathing modes
[298].

229

Optical Properties of Oligophenylenevinylenes

5. OPTICAL AND PHOTOPHYSICAL


PROPERTIES IN THE SOLID STATE
The optical and photophysical properties of OPVs in the
solid state are extensively studied on vapor-deposited (VD)
lms. Many experimental data are available on the absorption and uorescence spectra [2022, 28, 51, 54, 55, 67, 110,
146149, 153, 158, 237, 258, 301304], uorescence quantum yields [22, 54, 55, 237], and uorescence decay kinetics [22, 54, 55, 110, 158, 237, 258]. However, in only a few
cases were the geometrical arrangements of the oligomers
in the lms determined by atomic force microscopy [149]
or by polarized absorption studies [148, 159, 169, 301]; thus
structureproperty correlations are often missing. On the
other hand, the structures of several single crystals have
been determined by X-ray measurements [17, 25, 47, 60, 61,
68, 7380, 108114, 163, 185, 305, 306], and in some samples the uorescence properties were also investigated [80,
110, 163, 185188, 306]. VD lms, formed from the same
molecules as the corresponding single crystals, are expected
to form similar structures, but substrate-induced orientation
effects and the formation of meta-stable phases cannot be
excluded.
The model compounds of Section 4 demonstrated how
the optical and photophysical properties of the individual
molecules are changed by specic intermolecular electronic
interactions and vibrational coupling. Similar interactions
are expected in the solid state structures, but the extension to three dimensions will induce an increase in the total
intermolecular coupling strength. Moreover, the photophysical properties in the solid state are inuenced by additional effects, such as exciton diffusion, structural defects,
and chemical impurities. However, most of the spectral characteristics in OPV condensed phases can be explained by
(short-range) molecular ordering. A schematic matrix representation of nearest neighbor arrangements is given in
Figure 7.

5.1. Unsubstituted Oligomers


5.1.1. Structure
In 4PV single crystals the molecules are oriented with their
long axes parallel to each other, where the short axes of
adjacent molecules are arranged in a T-shaped manner
(Fig. 7.21) with a short-axis inclination angle of ( 71 [17].
The T-shaped short-axis arrangement of adjacent molecules
is also found for 2PV [80]. The corresponding herringbone
arrangement is also predicted for 7PV by force eld calculations [307]. A different packing was observed for 1PV
single crystals [7679]. The unit cell contains two molecules
that are oriented with their long axes almost perpendicular to each other. Nearest neighbor molecules are cofacially
arranged with a slight shift in the x,z plane (Fig. 7.23).
On the other hand, T-shaped arrangements are also
assumed for small aggregates of phospholipids with 1PV
substituents [130133]. The unit aggregate consists of a
chiral tetramer pinwheel, as determined by induced circular dichroism spectra [131, 133]. Analogous structures are
assumed in self-organized nanoparticles [146, 148, 174179]
or VD lms [20, 21, 28, 51, 146148, 301] of unsubstituted
nPVs (n = 24) as well as in LangmuirBlodgett (LB) lms

x,y-rotated

coplanar

x-shifted

11)

12)

13)

21)

22)

23)

z-shifted

parallel

31)

32)

x,z-rotated

z
y
x

Figure 7. Schematic matrix representation of simple unit cells of


oligophenylenevinylenes in the solid state.

of terminal-substituted nPVs (n = 1 [131, 159164], 2 [165


169]), since the spectral and photophysical features are very
similar for all of these systems. Especially, both LB [159,
169] and VD lms [148, 301] exhibit strong optical linear
dichroism, indicating a preferential parallel alignment of the
transition dipole moments.

5.1.2. Absorption Spectra


The absorption characteristics can be well studied in systems
with low optical density, such as VD lms (lm thickness
<50 nm), small nanoparticles ( < 100 nm), or LB lms. In
thick lms [21, 28, 51, 301, 308] or micro [51, 174, 309] or
single crystals [80] the increase in the optical density masks
most of the spectral characteristics by saturation effects.
The absorption spectra of thin VD lms and nanoparticles
of unsubstituted nPVs consist of some weak bands (A1 , A2 )
followed by the main maximum (H band, e.g., for 2PV [175]
see Fig. 8), which is blue-shifted by  6000 cm1 against
the monomer spectrum. The blue shift is a consequence of
exciton coupling of the transition dipole moments (H-type
aggregation). According to the theoretical treatment by
Spano [266, 310314], the weak low energy peaks (A1 , A2 ) of
the spectrum are ascribed to vibrationally dressed excitons
in pinwheels [266, 312] which may be present in the inhomogeneous distribution of different aggregate sizes [266]. The
noninteracting domains model [313], which accounts for
the distribution of molecular transition frequencies caused
by disorder, reproduces the ne structure of the spectrum.
The deviation from perfect parallel alignment of the transition dipoles was determined from the decrease in the timeresolved uorescence anisotropy from the initial value rF 0 =
04 to the steady-state value of rF = 020.3 [146, 148, 175
178], which yields a mean deviation angle of ) 30 between
the emitting species and the originally excited absorber. The
forbidden lower exciton state (J band) is seen as a very weak
shoulder at the low energy side of the spectrum [146, 313],

230

Optical Properties of Oligophenylenevinylenes

F2

2PV
F3

H
F1
A1

A2

fluorescence intensity

B2PV

Hex-2PV

pCN-2PV

20000

30000

40000

/cm 1
Figure 8. Fluorescence emission (left) and excitation (right) spectra of
(substituted) distyrylbenzene (2PV) nanoparticles (solid lines) and spectra in solution (dashed lines).

where the energetic position in 2PV nanoparticles coincides well with the electronic origin (00 = 23 900 cm1 )
observed in single crystals of 2PV [188]. The 00 position
is red shifted against solution by R = 1800 cm1 and
against vacuum by R = 3800 cm1 due to the high optical polarizability of the solid samples in the direction of the
long molecular axes. The anisotropic polarizability of 2PV
nanoparticles was demonstrated by polarized light scattering
measurements [176] and the spectral positions of the sensitized uorescence of dopants [178].

5.1.3. Fluorescence Properties


The room temperature steady-state uorescence spectra of
the unsubstituted nPVs are very similar for VD lms, LB
lms, and nanoparticles, whereas in thick lms and microcrystals the high energy side of the emission spectra is suppressed by reabsorption [80, 146, 187]. The spectra reveal
some vibronic structure (progression Fi , see Fig. 8) where
the energy spacing of the subbands resembles the one in
solution. Cooling of the lms and nanoparticles does not signicantly enhance the vibrational resolution of the spectra,
which is due to the spatial inhomogeneity of these systems.
Rich vibronic ne structure of the spectra was observed in

2PV single crystals [188] at low temperatures, clearly demonstrating that the vibronic progressions and intercombinations
are very similar to those observed in the spectra of the
isolated molecules. Thus the contribution of intermolecular vibrational modes is small. However, the electronic origins as well as the F1 features (Fig. 8) are only of very
weak intensity, in signicant contrast to the solution spectra.
According to the theoretical investigations by Spano [310
314] and Meskers et al. [158], this effect is ascribed to electronic interactions in the medium coupling regime. In the
nanoparticles, the F1 band shows varying intensities [174],
which are caused by distributions of zero-order excited state
energies [158] and/or distributions of interchain interactions
[158, 310314] due to (point) defects [312, 313]. The uorescence lifetimes (2PV: F 25 ns, Table 8) are considerably
longer than those in solution (2PV: F = 12 ns), whereas
the uorescence quantum yields are reduced from F = 09
in solution to F = 0050.1 in the lms and nanoparticles.
Thus the radiative rate constant kF decreases against solution by 1 order of magnitude. At T = 20 K, F increases by
a factor of 45 [54, 309], similar to PPV [258, 309], reaching the room temperature value of the 2PV single crystal
(F = 065 [80]). The increase in F is accompanied by an
increase in F of the same order [54, 146, 177]; thus in a rst
approximation kF is independent of temperature.
Energy migration of the S1 excitation energy in nPV lms
is due to a diffusion-enhanced Forster transfer process, as
revealed from the bimolecular rate constant of S1 S1 annihilation [308]. The exciton diffusion (diffusion constant D
8 104 cm2 s1 ) [308] allows the deactivation via intrinsic or
extrinsic traps. For 6BOPV (Table 4), which also forms H
aggregates [38, 158, 173], a distribution of disorder-induced
trap sites was deduced from time-resolved and site-selective
uorescence spectra [158]. With increasing excitation wavelength, the uorescence decay becomes exponential on the
time scale of the rst nanoseconds [158]. However, on longer
time scales, up to the ms regime [315], the uorescence
decay traces of the nPVs are strongly nonexponential [146,
148, 177]. This may be due to the recombination of bounded
polaron pairs, whose generation was observed in a photoinduced absorption (PIA) study [308]. In this study evidence
for ssion of vibrationally hot S1 states into triplet pairs
was also found, a process which additionally reduces F .
PIA studies on alkyloxy-substituted 4PV lms (MeH4PV, see
Table 9) [152157, 316, 317] show that at low excitation densities, singlet excitons are formed as the only photoexcitations. At higher excitation densities, the singlet excitons are
subject to bimolecular annihilation. In contrast to unsubstituted OPV, no evidence of exciton diffusion is found [153].
An additional contribution to the PIA at high excitation densities is ascribed to biexciton states that are stable in solution
but dissociate into charged pairs in thin lms [154]. The dissociation of singlet excitons into charged states is found to
be a common process in conjugated polymers [157].

5.1.4. Doping with Acceptor Molecules


for Electronic Excitation Transfer
Doping of the nPVs with suitable acceptor molecules, such
as longer OPVs [146, 177], oligothiophenes [178], or acenes
[186], leads to efcient energy transfer to these extrinsic

231

Optical Properties of Oligophenylenevinylenes

traps, resulting in sensitized uorescence of the dopants. At


doping ratios of xdop > 5 104 the emission is almost completely dominated by the acceptor emission [146, 177, 178].
The appropriate choice of the acceptors allows tuning of
the emission wavelength over the visible range. The acceptor molecules possess the same preferential orientation as
the surrounding nPV molecules, as revealed from the constant uorescence anisotropy over the whole emission range
[177, 178]. The deactivation kinetics depend strongly on the
choice of the dopant. OPVs as acceptors enhance the overall
uorescence quantum yield of doped 2PV nanoparticles up
to F tot = 06 [146, 177]. Doping with oligothiophenes does
not change F tot , which is probably a consequence of the
competition between energy transfer and photoinduced electron transfer [178] from the HOMO of quinquethiophene
to the HOMO of 2PV, which is approximately 1 eV lower in
energy [150, 318]. The strong decrease in donor emission is
not accompanied by a comparable decrease in uorescence
decay, thus indicating spatial inhomogeneities in these systems [146, 178].

5.2. Substituted Oligomers


5.2.1. t-Butyl-Substituted OPVs
Space-consuming substituents like tertiary butyl groups
(BnPVs, see Table 2) hinder the coplanar or side-by-side
alignments of adjacent molecules (Fig. 7.21, 7.22), since
now these structures are energetically less favored than, for
example, pairs with x,z-rotated axes (Fig. 7.31, 7.32). The
lack of short-range parallel intermolecular orientation was
demonstrated by the very fast complete uorescence depolarization, occurring in the rst ps after excitation [148, 176,
177]. Due to the different orientations of neighboring transition dipoles, the absorption spectra are very similar to those
in solution [148, 176, 177] (see Fig. 8) and no indication
of a blue-shifted H band can be observed. Also the energy
spacings and relative intensities of the Fi bands in the uorescence spectra are almost identical to the ones in solution.
The small red shift of R 400 cm1 against solution is
due to the high, approximately isotropic polarizability of the
condensed material. The polarizability in the nonabsorbing
region of B2PV nanoparticles was determined as isotropic
by polarized light scattering measurements [176, 178]. The
calculated isotropic value of the refractive index is in good
agreement with the averaged value obtained from the polarizabilities of the constituting atoms and the one derived from
the spectral positions of the sensitized uorescence spectra
of dopants [178]. The uorescence quantum yields are low at
room temperature (e.g., B2PV: F = 012) [148], increasing
by a factor of three at T = 20 K [258].

5.2.2. OPVs with Alkyl or Alkoxy Side Chains


Alkyl or alkoxy side chains (e.g., nPOPVs, nBOPVs,
Tables 3, 4) destabilize T-shaped arrangements (Fig. 7.21)
and favor partial  overlap. Thus in general cofacial
arrangements, displaced in the x,z plane (Fig. 7.12, 7.13,
7.23), are observed. The 4PV oligomer with octyloxy side
chains in the central phenyl ring (OctO-4PV, see Table 9)
[24, 25, 74, 114, 305, 306] may serve as an example: The
unit cell of the single crystal contains eight molecules, where

the conjugated backbones lie parallel to each other. Neighboring molecules are oriented in a cofacial arrangement,
shifted in the x,z plane [305]. The highly anisotropic character of the cofacial arrangement becomes evident from the
high degree of uorescence polarization [114]. The crystal
shift of the uorescence spectrum against solution yields
R 900 cm1 (Table 9). This value is signicantly higher
than that in isotropic solids (e.g., BnPVs) but much smaller
than that in the herringbone-arranged unsubstituted 4PV
sample (R 1800 cm1 , see Table 9). The smaller shift
is due to the displacement of the OctO-4PV molecules in
the x,z plane, but also to the decrease in the packing density by 11% compared to 4PV [17, 306]. The displacement
gives rise to J aggregation with a well-structured, red-shifted
absorption spectrum and high uorescence quantum yields
and decay rates [24, 306] (see Table 9). Similar considerations also apply for Oct-4PV [303] (Table 9) and Hex-2PV
[54] (Fig. 8, Table 8), which all display well-structured uorescence spectra similar to the ones in solution. No evidence
for contributions of intermolecular vibrational modes to the
uorescence spectra is found for these systems.
On the other hand, microcrystals of hexyloxy-substituted
2PV (HexO-2PV, Table 8) show excimer-like emission [54],
thus indicating a strong  overlap of the emitting adjacent molecules. The uorescence is completely depolarized, which may be due to a random distribution of small
-stacked aggregates or to self-trapping of the excitation
energy in excimer-like states [54].

5.2.3. Cyanosubstituted OPVs


The tendency to arrange in a -stacked structure increases
by introduction of cyanosubstituents. However, the crystal structure of cyanoderivatives depends sensitively on the
substitution position. Thus cyanosubstitution at the inner
position of the central vinyl moieties of octyloxy-4PV (OctO4PV-CNi , see Table 9) results in a cofacial x,z-shifted
arrangement of neighboring molecules (Fig. 7.13) [73],
where the displacement between the molecules is too large
to allow intermolecular vibrational coupling. Concomitantly
the uorescence spectrum is structured and no excimer
emission is observed [306].
On the other hand, cyanosubstitution at the outer positions of the central vinyl linkages (OctO-4PV-CNo , see
Table 9) results in a z shift of neighboring molecules
(Fig. 7.12) by one phenylene unit [306]. The  overlap
is now large and gives rise to high FranckCondon activity of intermolecular vibrational modes, leading to excimer
emission with a long uorescence decay time of 8 ns, with a
considerably high quantum yield of F = 05 [306]. Excimerlike emission spectra are also observed for most of the
cyanosubstituted 2PVs with additional hexyloxy substituents
in the central phenyl ring (Table 8) [54, 57, 237]. However, it
should be emphasized that the preparation conditions play
a crucial role in the resulting arrangement of the molecules
and hence, in the optical and photophysical properties. An
illustrative example is p,p -dicyanosubstituted 2PV (pCN2PV). Depending on the preparation conditions (Table 8), H
aggregates [319], J aggregates with well-structured emission
spectra, or arrangements with excimer-like emission spectra
(Fig. 8) [55] are observed (Table 8). In all these arrangements the molecules show a preferential parallel orientation

232
of the long molecular axes, as revealed from the high values
of the uorescence anisotropy, rF (see Table 8). The differences in the absorption and uorescence spectra of the
systems indicate that small changes in the geometrical alignment may induce strong changes in the optical and photophysical properties of the materials.
Extended  stacks of adjacent molecules were
observed for single crystals of various uorosubstituted distyrylbenzenes [60] and binary cocrystals, containing pairs of
unsubstituted and uorinated nPVs, repectively [61, 109].
Nanoparticles and VD lms of these materials exhibit
H-type absorption spectra and excimer-like emission [321].
Excimer emission is also observed for single crystals of 3,5dichloro-trans-stilbene (Fig. 1), which also crystallizes in a
cofacial arrangement [185].

NOTE
Since the submission of the review chapter a number of new
papers in the eld of oligophenylenevinylenes have been
published [322363], which could not be included in the text,
but should be brought to the attention of the reader. Further
studies on the transcis photoisomerization of trans-stilbene
(1PV) were published by several authors [322324]. Room
temperature Raman studies on 1PV suggest a distorted
structure in solution [325]. In a theoretical paper, MllerPlesset (MP2, MP3) calculations suggest a strictly planar
geometry of 1PV in its absolute energy minimum [326].
Several papers described the preparation and spectroscopic
characterization of new phenyl [327], alkyl [328], alkoxy
[329331], alkoxysilyl [332], amino [333, 334], bromo [335],
and push-pull [336, 337] substituted OPVs. The effects of
conformational [338, 339] and environmental disorder [338]
on the optical properties were investigated both experimentally [338] and theoretically [338, 339]. A new method for
the correlation of molecular shape and polarized luminescence, examined on OPVs, was introduced in Ref. [340]. The
electronic transition energies of (substituted) OPVs were
investigated at different quantum-chemical levels of theory
[341345], which inter alia allow the extrapolation for several
ideally conjugated substituted PPV chains [341]. Two-photon
properties of substituted OPVs were determined both experimentally [346] and theoretically [347, 348]. Molecular interactions in model compounds and their impact on the optical
and photophysical properties were studied in OPV dimers
with biphenyl linkage center [349], complementary oligonucleotides containing 1PV units [350], tetrahedral OPV structures [351], polystyrene bearing stilbenoid side chains [352],
and supramolecular OPV ensembles [353]. Reference [354]
shows that excimer formation of substituted OPVs might
be used as strain sensors in polymer blends. Photo- [355
358] and electroluminescence [359362] of lms formed by
substituted OPVs were investigated by a number of groups,
addressing the inuence of substituents [362], crystal structure [355] and morphology of the sample [356, 360]. OPVs
have also been used as active material for energy transfer
in gelation-assisted light harvesting [363] and for photoinduced electron transfer in TiO2 hybrid materials [364]. Silver
nanoparticles with OPV coating may act as ultrabright twophoton uorescent nanobeacons [365].

Optical Properties of Oligophenylenevinylenes

GLOSSARY
Bridged oligomers Model systems of a dened number of
chromophores whose conformational degrees of freedom
are reduced by hydrogen or covalent bonds between the
chromophores. These systems can be used to understand
intermolecular interactions of adjacent molecules in the
solid state and to build up supramolecular architectures, for
example, dendrimers and helical self-organized structures.
Nanochannel-forming hostguest compounds Hostguest
compounds (HGCs), in which chromophores are incorporated into the channels of inorganic (e.g., zeolites,
nanoporous silica) or organic host materials (e.g., perhydrotriphenylene). HGCs provide ideal systems for studying
weak dipolar interactions of chromophores and open up
the opportunity to overcome the drawbacks of conventional
condensed phase device structures for organic optoelectronic applications: Improved chemical stability by exclusion
of oxygen, high uorescence quantum yields by avoiding
intermolecular aggregation, intrinsic polarization, high NLO
efciencies, and spatially directed energy transfer.
Nanoparticles Nanoparticles of organic chromophores are
obtained by precipitation from solutions by addition of
a poor solvent. Nanoparticles allow the investigation of
substrate-free self-organized systems, easy color tuning by
chemical doping with appropriate acceptor molecules, and
preparation of optoelectronic devices.
Oligomer approach Oligomeric -conjugated organic
molecules provide well-dened conjugation lengths compared to polymers, variable substitution patterns, and
dened intermolecular orientations in the condensed
phase. Hence oligomers with variable conjugation lengths
and substituents make possible the systematic study of
the interplay of intramolecular (optical) properties and
intermolecular interactions as well as the understanding of
the properties of the polymers. These well-dened systems
permit theoretical modeling of solid state properties for
molecular engineering of optoelectronic devices.

REFERENCES
1. M. G. Harrison and R. H. Friend, in Electronic Materials: The

Oligomer Approach (K. Mullen


and G. Wegner, Eds.), p. 515.
Wiley-VCH, Weinheim 1998.
2. R. H. Friend, R. W. Gymer, A. B. Holmes, J. H. Burroughes,
R. N. Marks, C. Taliani, D. D. C. Bradley, D. A. dos Santos, J. L.
Brdas, M. Loglund, and W. R. Salaneck, Nature (London) 397,
121 (1999).
3. L. Pichat, P. Peistel, and J. Clement, J. Chem. Phys. 50, 26 (1953).
4. (a) V. P. Kotsubanov, L. Ya. Malkes, Yu. V. Naboikin, L. A.
Ogurtsova, A. P. Podgornyi, F. S. Potrovskaya, and L. V. Shubina,
Zh. Prik. Spektrosc. 10, 152 (1969); (b) H. Furumato and H. Ceccon, J. Appl. Phys. 40, 4204 (1969); (c) J. T. Warden and L. Gough,
Appl. Phys. Lett. 19, 345 (1971); (d) H. Tell, U. Brinkmann, and
P. Raue, Opt. Commun. 24, 248 (1978).
5. A. E. Siegrist, P. Liechti, H. R. Meyer, and K. Weber, Helv. Chim.
Acta 52, 2521 (1969).
6. (a) D. H. Waldeck, Chem. Rev. 91, 415 (1991) and references cited
therein; (b) T. Arai and K. Tokumaru, Chem. Rev. 93, 23 (1993)
and references cited therein.
7. G. Hohlneicher, R. Wrzal, D. Lenoir, and R. Frank, J. Phys. Chem.
A 103, 8696 (1999).

Optical Properties of Oligophenylenevinylenes


8. W. G. Han, T. Lovell, T. Liu, and L. Noodleman, Chem. Phys.
Chem. 2002, 167 (2002).
9. V. Molina, M. Merchn, and B. O. Roos, J. Phys. Chem. A 101,
3478 (1997).
10. J. Quenneville and T. J. Martnez, J. Phys. Chem. A 107, 829 (2003).
11. I. Anger, K. Sandros, M. Sundahl, and O. Wennerstro m, J. Phys.
Chem. 97, 1920 (1993).
12. H. Rau and I. Waldner, Phys. Chem. Chem. Phys. 4, 1776 (2002).

13. R. Schenk, H. Gregorius, K. Meerholz, J. Heinze, and K. Mullen,


J. Am. Chem. Soc. 113, 2634 (1991).

14. Y. Geerts, G. Klarner, and K. Mullen,


in Electronic Materials:

The Oligomer Approach (K. Mullen


and G. Wegner, Eds.), p. 1.
Wiley-VCH, Weinheim (1998).
15. R. E. Martin and F. Diederich, Angew. Chem. Int. Ed. 38, 1350
(1999).
16. M. S. Wong, Z. H. Li, M. F. Shek, K. H. Chow, Y. Tao, and
M. DIorio, J. Mater. Chem. 10, 1805 (2000).
17. P. F. van Hutten, J. Wildeman, A. Meetsma, and G. Hadziioannou,
J. Am. Chem. Soc. 121, 5910 (1999).
18. V. Gebhardt, A. Bacher, M. Thelakkat, U. Stalmach, H. Meier,
H.-W. Schmidt, and D. Haarer, Synth. Met. 90, 123 (1997).
19. U. Stalmach, H. Detert, H. Meier, V. Gebhardt, D. Haarer,
A. Bacher, and H.-W. Schmidt, Opt. Mater. 9, 77 (1998).
20. F. Meghdadi, G. Leising, W. Fischer, and F. Stelzer, Synth. Met. 76,
113 (1996).
21. H. S. Woo, J. G. Lee, H. K. Min, E. J. Oh, S. J. Park, K. W. Lee,
J. H. Lee, S. H. Cho, T. W. Kim, and C. H. Park, Synth. Met. 71,
2173 (1995).
22. T. Goodson III, W. Li, A. Gharavi, and L. Yu, Adv. Mater. 9, 639
(1997).
23. V. Gebhardt, A. Bacher, M. Thelakkat, U. Stalmach, H. Meier,
H.-W. Schmidt, and D. Haarer, Adv. Mater. 11, 119 (1999).
24. H.-J. Brouwer, V. V. Krasnikov, T.-A. Pham, R. E. Gill, and
G. Hadziioannou, Appl. Phys. Lett. 73, 708 (1998).
25. P. F. van Hutten, V. V. Krasnikov, and G. Hadziioannou, Acc.
Chem. Res. 32, 257 (1999).
26. Z. Yang, H. J. Geise, M. Mehbod, G. Debrue, J. W. Visser, E. J.
Sonneveld, L. Vant Dack, and R. Gijbels, Synth. Met. 39, 137
(1990).

27. G. Drefahl, R. Kuhmstedt,


H. Oswald, and H.-H. Horhold, Makromol. Chem. 131, 89 (1970).
28. H.-J. Egelhaaf, M. Brun, S. Reich, and D. Oelkrug, J. Mol. Struct.
267, 297 (1992).

29. L. Luer,
H.-J. Egelhaaf, and D. Oelkrug, Synth. Met. 119, 621
(2001).

30. L. Luer,
H.-J. Egelhaaf, and D. Oelkrug, Opt. Mater. 9, 454 (1998).
31. J. H. Schon, C. Kloc, J. Wildeman, and G. Hadziioannou, Adv.
Mater. 13, 1273 (2001).
32. B. de Boer, U. Stalmach, C. Melzer, and G. Hadziioannou, Synth.
Met. 121, 1541 (2001).
33. B. de Boer, U. Stalmach, H. Nijland, and G. Hadziioannou, Adv.
Mater. 12, 1581 (2000).
34. M. Albota, D. Beljonne, J. L. Brdas, J. E. Ehrlich, J.-Y. Fu,
A. A. Heikal, S. E. Hess, T. Kogej, M. D. Levin, S. R. Marder,
D. McCord-Maughon, J. W. Perry, H. Rockel, M. Rumi, G. Subramaniam, W. W. Webb, X.-L. Wu, and C. Xu, Science 281, 1653
(1998).
35. (a) S. Misumi, M. Kuwana, K. Murashima, and M. Nakagawa,
Bull. Chim. Soc. Jpn. 34, 1833 (1961); (b) S. Misumi, M. Kuwana,
and M. Nakagawa, Bull. Chim. Soc. Jpn. 35, 135 (1962).

36. R. Erckel and H. Fruhbeis,


Z. Naturforsch. 37b, 1472 (1982).
37. U. Stalmach, H. Kolshorn, I. Brehm, and H. Meier, Liebigs Ann.
1996, 1449 (1996).
38. E. Peeters, R. A. J. Janssen, S. C. J. Meskers, and E. W. Meijer,
Polym. Prepr. (Am. Chem. Soc. Div. Polym. Chem.) 40, 519 (1999).
39. R. Reetz, O. Narwark, O. Herzog, S. Brocke, and E. Thorn-Csnyi,
Synth. Met. 119, 539 (2001).

233
40. O. Narwark, O. Herzog, and E. Thorn-Csnyi, Synth. Met. 121,
1375 (2001).
41. O. Herzog, O. Narwark, and E. Thorn-Csnyi, Synth. Met. 119, 141
(2001).
42. E. Thorn-Csnyi and P. Kraxner, Macromol. Chem. Phys. 198, 3827
(1997).
43. A. P. H. J. Schenning, A. El-ghayoury, E. Peeters, and E. W. Meijer, Synth. Met. 121, 1253 (2001).
44. H. Meier, J. Gerold, H. Kolshorn, W. Baumann, and M. Bletz,
Angew. Chem. Int. Ed. 41, 292 (2002).

45. G. Klarner, C. Former, X. Yan, R. Richert, and K. Mullen,


Adv.
Mater. 11, 932 (1996).
46. I. Anger, M. Sundahl, O. Wennerstrom, K. Sandros, T. Arai, and
K. Tokumaru, J. Phys. Chem. 96, 7027 (1992).
47. H. Irngartinger, J. Lichtenthaler, and R. Herpich, Struct. Chem. 5,
283 (1994).
48. E. Birckner, U.-W. Grummt, H. Rost, A. Hartmann, S. Pfeiffer,
H. Tillmann, and H.-H. Horhold, J. Fluoresc. 8, 73 (1998).
49. S. Nakatsuji, K. Matsuda, Y. Uesugi, K. Nakashima, S. Akiyama,
G. Katzer, and W. Fabian, J. Chem. Soc. Perkin Trans. 2, 6, 861
(1991).
50. U. Stalmach and H. Detert, J. Prakt. Chem. 342, 10 (2000).
51. N. N. Barashkov, D. J. Guerrero, H. J. Olivos, and J. P. Ferraris,
Synth. Met. 75, 153 (1995).
52. F. Babudri, G. M. Farinola, L. C. Lopez, M. G. Martinelli, and
F. Naso, J. Org. Chem. 66, 3878 (2001).
53. M. S. Wong, M. Samoc, A. Samoc, B. Luther-Davies, and M. G.
Humphrey, J. Mater. Chem. 8, 2005 (1998).
54. K.-H. Schweikhart, M. Hohloch, E. Steinhuber, M. Hanack,
J. Gierschner, H.-J. Egelhaaf, and D. Oelkrug, Synth. Met.
L. Luer,
121, 1641 (2001).
55. J. Gierschner, H.-J. Egelhaaf, H.-G. Mack, D. Oelkrug,
R. Martinez-Alvarez, and M. Hanack, Synth. Met. 137, 1449 (2003).
56. M. S. Liu, X. Jiang, S. Liu, P. Herguth, and A. K.-Y. Jen, Macromolecules 35, 3532 (2002).
and D. Oelkrug, Eur. J. Org.
57. K.-H. Schweikart, M. Hanack, L. Luer,
Chem. 2001, 293 (2001).
58. H. Detert and E. Sugiono, Synth. Met. 127, 233 (2002).
59. B. Strehmel, A. M. Sarker, J. H. Malpert, V. Strehmel, H. Seifert,
and D. C. Neckers, J. Am. Chem. Soc. 121, 1226 (1999).
60. M. L. Renak, G. P. Bartholomew, S. Wang, P. J. Ricotto, R. J.
Lachicotte, and G. C. Bazan, J. Am. Chem. Soc. 121, 7787 (1999).
61. G. W. Coates, A. R. Dunn, L. M. Henling, J. W. Ziller, E. B.
Lobkovsky, and R. H. Grubbs, J. Am. Chem. Soc. 120, 3641 (1998).
62. K. Sandros and M. Sundahl, J. Phys. Chem. 98, 5705 (1994).
63. E. Sugiono, T. Metzroth, and H. Detert, Adv. Synth. Catal. 343,
351 (2001).
64. H. Detert and E. Sugiono, Synth. Met. 127, 237 (2002).
65. A. C. Freydank, M. G. Humphrey, R. W. Friedrich, and B. LutherDavies, Tetrahedron 58, 1425 (2002).
66. S. E. Dottinger, M. Hohloch, J. L. Segura, E. Steinhuber,
M. Hanack, A. Tompert, and D. Oelkrug, Adv. Mater. 9, 233 (1997).
67. P. Martinez-Ruiz, B. Behnisch, K.-H. Schweikhart, M. Hanack,

L. Luer,
and D. Oelkrug, Chem. Eur. J. 2000, 1294 (2000).
68. M. Hohloch, C. Maichle-Mossmer, and M. Hanack, Chem. Mater.
10, 1327 (1998).
69. M. M. de Souza, G. Rumbles, I. R. Gould, H. Amer, I. D.
W. Samuel, S. C. Moratti, and A. B. Holmes, Synth. Met. 111112,
539 (2000).
70. M. Leuze, M. Hohloch, and M. Hanack, Chem. Mater. 14, 3339
(2002).
71. M. M. de Souza, G. Rumbles, D. L. Russel, I. D. W. Samuel, S.
C. Moratti, A. B. Holmes, and P. L. Burn, Synth. Met. 119, 635
(2001).
72. H. Detert and E. Sugiono, J. Phys. Org. Chem. 13, 587 (2000).
73. R. E. Gill, P. F. van Hutten, A. Meetsema, and G. Hadziioannou,
Chem. Mater. 8, 1341 (1996).

234
74. R. E. Gill, A. Hilberer, P. F. van Hutten, G. Berentschott, M. P.
L. Werts, A. Meetsema, J.-C. Wittmann, and G. Hadziioannou,
Synth. Met. 84, 637 (1996).
75. H. Detert, D. Schollmeyer, and E. Sugiono, Eur. J. Org. Chem.
2001, 2927 (2001).
76. J. Bernstein, Acta Crystallogr. B 31, 1268 (1975).
77. C. J. Finder, M. G. Newton, and N. L. Allinger, Acta Crystallogr.
B 30, 411 (1974).
78. A. Hoeckstra, P. Meertens, and A. Voss, Acta Crystallogr. B 31,
2813 (1975).
79. J. A. Bouwstra, A. Schouten, and J. Kroon, Acta Crystallogr. C 40,
428 (1984).
80. C. C. Wu, M. C. DeLong, Z. V. Vardeny, and J. P. Ferraris, Synth.
Met. 137, 939 (2003).
81. T. Suzuki, N. Mikami, and M. Ito, J. Phys. Chem. 90, 6431 (1986).
82. (a) L. H. Spangler, R. van Zee, S. C. Blankespoor, and T. S. Zwier,
J. Phys. Chem. 91, 6077 (1987); (b) L. H. Spangler, R. van Zee,
and T. S. Zwier, J. Phys. Chem. 91, 2782 (1987).
83. W.-Y. Chiang and J. Laane, J. Chem. Phys. 100, 8755 (1994).
84. B. B. Champagne, J. F. Pfanstiel, D. F. Plusquellic, D. W. Pratt,
W. M. van Herpen, and W. L. Meerts, J. Phys. Chem. 96, 6 (1990).
85. S. P. Kwasniewski, M. S. Deleuze, and J. P. Francois, Int. J. Quantum Chem. 85, 557 (2001).
86. L. Claes, S. Kwasniewski, M. S. Deleuze, and J. P. Franois, J. Mol.
Struct. (Theochem) 549, 63 (2001).
87. C. H. Choi and M. Kertesz, J. Phys. Chem. A 101, 3823 (1997).
88. S. P. Kwasniewski, M. S. Deleuze, and J. P. Francois, Int. J. Quantum Chem. 80, 672 (2000).
89. Z. G. Soos, S. Ramasesha, D. S. Galvo, and S. Etemad, Phys. Rev.
B 47, 1742 (1993).
90. D. S. Galvo, Z. G. Soos, S. Ramasesha, and S. Etemad, J. Chem.
Phys. 98, 3016 (1993).

91. J. Gierschner, H.-G. Mack, L. Luer,


and D. Oelkrug, J. Chem.
Phys. 116, 8596 (2002).
92. C.-J. Kim, Bull. Korean. Chem. Soc. 23, 330 (2002).
93. O. Lhost and J. L. Brdas, J. Chem. Phys. 96, 5279 (1992).
94. G. Baranovic, Z. Meic, and A. H. Maulitz, Spectrochim. Acta A
54, 1017 (1998).
95. G. C. Claudio and E. R. Bittner, Chem. Phys. 276, 81 (2002).

96. a) B. Tian, G. Zerbi, R. Schenk, and K. Mullen,


J. Chem. Phys. 95,

3191 (1991); (b) B. Tian, G. Zerbi, and K. Mullen,


J. Chem. Phys.
95, 3198 (1991).

97. S. Karabunarliev, M. Baumgarten, and K. Mullen,


J. Phys. Chem.
A 104, 8236 (2000).
98. J. F. Arenas, I. L. Tocn, J. C. Otero, and J. I. Marcos, J. Phys.
Chem. 99, 11392 (1995).
99. L. Gagliardi, G. Orlandi, V. Molina, P.-A. Malmqvist, and B. Roos,
J. Phys. Chem. A 106, 7355 (2002).
100. F. Negri, G. Orlandi, and F. Zerbetto, J. Phys. Chem. 93, 5124
(1989).
101. D. Beljonne, Z. Shuai, R. H. Friend, and J. L. Brdas, J. Chem.
Phys. 102, 2042 (1995).
102. H. Watanabe, Y. Okamoto, K. Furuya, A. Sakamoto, and
M. Tasumi, J. Phys. Chem. A 106, 3318 (2002).
103. D. Beljonne, J. Cornil, J. L. Brdas, and R. H. Friend, Synth. Met.
76, 61 (1996).
104. M. Trtteberg, E. B. Frantsen, F. C. Mijehoff, and A. Hoeckstra,
J. Mol. Struct. 26, 57 (1975).
105. T. Kobayashi, H. Suzuki, and K. Ogawa, Bull. Chem. Soc. Jpn. 55,
1734 (1982).
106. P. Rademacher, A. L. Marzinzik, K. Kowski, and M. E. Wei, Eur.
J. Org. Chem. 2001, 121 (2001).

107. S. Karabunarliev, M. Baumgarten, N. Tyutyulkov, and K. Mullen,


J. Phys. Chem. 98, 11892 (1994).
108. U. Stalmach, D. Schollmeyer, and H. Meier, Chem. Mater. 11, 2103
(1999).

Optical Properties of Oligophenylenevinylenes


109. G. P. Bartholomew, X. Bu, and G. C. Bazan, Chem. Mater. 12, 2311
(2000).
110. P. F. van Hutten, H.-J. Brouwer, V. V. Krasnikov, L. Ouali, U. Stalmach, and G. Hadziioannou, Synth. Met. 102, 1443 (1999).
111. M. Verbruggen, Yang Zhou, A. T. H. Lenstra, and H. J. Geise,
Acta Crystallogr. C 44, 2120 (1988).
112. K. Ogawa, H. Suzuki, T. Sakurai, K. Kobayashi, A. Kira, and
A. Toriumi, Acta Crystallogr. C 44, 505 (1988).
113. M. Hakansson, S. Jagner, M. Sundahl, and O. Wennerstro m, Acta
Chem. Scand. 46, 1160 (1992).
114. H. J. Brouwer, V. V. Krasnikov, T. A. Pham, R. E. Gill, P. F. van
Hutten, and G. Hadziioannou, Chem. Phys. 227, 65 (1998).
115. A. El-ghayoury, E. Peeters, A. P. H. J. Schenning, and E. W. Meijer, Chem. Commun. 2000, 1969 (2000).
116. A. P. H. J. Schenning, P. Jonkheijm, E. Peeters, and E. W. Meijer,
J. Am. Chem. Soc. 123, 409 (2001).
117. A. El-ghayoury, A. P. H. J. Schenning, P. A. van Hal, K. J. van
Duren, R. A. Janssen, and E. W. Meijer, Angew. Chem. Int. Ed.
40, 3660 (2001).
118. N. Verdal, J. T. Godbout, T. L. Perkins, G. P. Bartholomew, G. C.
Bazan, and A. Myers Kelley, Chem. Phys. Lett. 320, 95 (2000).
119. G. C. Bazan, W. J. Oldham Jr., R. J. Lachicotte, S. Tretiak,
V. Chernyak, and S. Mukamel, J. Am. Chem. Soc. 120, 9188 (1998).
120. S. Wang, G. C. Bazan, S. Tretiak, and S. Mukamel, J. Am. Chem.
Soc. 122, 1289 (2000).
121. W. J. Oldham Jr., Y.-J. Mio, R. J. Lachicotte, and G. C. Bazan,
J. Am. Chem. Soc. 120, 419 (1998).
122. A. M. Moran, G. P. Bartholomew, G. C. Bazan, and A. Myers
Kelley, J. Phys. Chem. A 106, 4928 (2002).
123. J. Zyss, I. Ledoux, S. Volkov, V. Chernyak, S. Mukamel, G. P.
Bartholomew, and G. C. Bazan, J. Am. Chem. Soc. 122, 11956
(2000).
124. G. P. Bartholomew, I. Ledoux, S. Mukamel, G. C. Bazan, and
J. Zyss, J. Am. Chem. 124, 13480 (2002).
125. J. W. Hong, B. S. Gaylord, and G. C. Bazan, J. Am. Chem. Soc.
124, 11868 (2002).
126. S. Wang, W. J. Oldham, Jr., R. A. Hudack, Jr., and G. C. Bazan,
J. Am. Chem. Soc. 122, 5695 (2000).
127. M. R. Robinson, S. Wang, G. C. Bazan, and Y. Cao, Adv. Mater.
12, 1701 (2000).
128. O. P. Varnavski, J. C. Ostrowski, L. Sukhomlinova, R. J. Twieg,
G. C. Bazan, and T. Goodson III, J. Am. Chem. Soc. 124, 1736
(2002).
129. M. A. Summers, M. R. Robinson, G. C. Bazan, and S. K. Buratto,
Chem. Phys. Lett. 364, 542 (2002).
130. X. Song, C. Geiger, U. Leinhos, J. Perlstein, and D. G. Whitten,
J. Am. Chem. Soc. 116, 10340 (1994).
131. X. Song, C. Geiger, M. Farahat, J. Perlstein, and D. G. Whitten,
J. Am. Chem. Soc. 119, 12481 (1997).
132. X. Song, C. Geiger, I. Furman, and D. G. Whitten, J. Am. Chem.
Soc. 116, 4103 (1994).
133. D. G. Whitten, L. Chen, C. Geiger, J. Perlstein, and X. Song,
J. Phys. Chem. B 102, 10098 (1998).
134. B. Konig, B. Knieriem, and A. de Meijere, Chem. Ber. 126, 1643
(1997).
135. Y. Ito, S. Miyata, M. Nakatsuka, T. Saegusa, M. Takamoto, and
Y. Wada, J. Chem. Soc., Chem. Commun. 375 (1982).
136. O. Varnawski, I. D. W. Samuel, L.-O. Plson, R. Beavington, P. L.
Burn, and T. Goodson III, J. Chem. Phys. 116, 8893 (2002).
137. C. C. Kwok and M. S. Wong, Macromolecules 34, 6821 (2001).

138. S. C. J. Meskers, M. Bender, J. Hubner,


Y. V. Romanovskii,
M. Oestreich, A. P. H. J. Schenning, E. W. Meijer, and H. Bassler,
J. Phys. Chem. A 105, 10220 (2001).
139. D. M. Guldi, A. Swartz, C. Luo, R. Gmez, J. L. Segura, and
N. Martin, J. Am. Chem. Soc. 124, 10875 (2002).
140. J. M. Lupton, I. D. W. Samuel, and P. L. Burn, Phys. Rev. B 66,
155206 (2002).

Optical Properties of Oligophenylenevinylenes


141. L.-O. Plson, R. Beavington, M. J. Frampton, J. M. Lupton, S. W.
Magennis, J. P. J. Markham, J. N. G. Pillow, P. L. Burn, and
I. D. W. Samuel, Macromolecules 35, 7891 (2002).
142. R. L. Letsinger and T. Wo, J. Am. Chem. Soc. 116, 811 (1994).
143. F. D. Lewis, T. Wu, E. L. Burch, D. M. Bassani, Y.-S. Yang,
S. Schneider, W. Jager, and R. L. Letsinger, J. Am. Chem. Soc.
117, 8785 (1995).
144. R. L. Letsinger and T. Wu, J. Am. Chem. Soc. 117, 7323 (1995).
145. F. D. Lewis, Y.-S. Yang, and C. L. Stern, J. Am. Chem. Soc. 118,
2772 (1996).
146. D. Oelkrug, A. Tompert, J. Gierschner, H.-J. Egelhaaf,
M. Hanack, M. Hohloch, and E. Steinhuber, J. Phys. Chem. B 102,
1902 (1998).
147. D. Oelkrug, J. Haiber, R. Lege, H. Stauch, and H.-J. Egelhaaf,
Thin Solid Films 284285, 581 (1996).
148. H.-J. Egelhaaf, J. Gierschner, and D. Oelkrug, Synth. Met. 83, 221
(1996).
149. J. P. Ni, Y. Ueda, T. Hanada, N. Takada, Y. Ichino, Y. Yoshida,
N. Tanigaki, K. Yase, D. K. Wang, and F. S. Wang, J. Appl. Phys.
86, 6150 (1999).
150. A. Schmidt, M. L. Anderson, D. Dunphy, T. Wehrmeister,

K. Mullen,
and N. R. Armstrong, Adv. Mater. 7, 722 (1995).
151. S. C. Veenstra, U. Stalmach, V. V. Krasnikov, G. Hadziioannou,
H. T. Jonkman, A. Heeres, and G. A. Sawatzky, Appl. Phys. Lett.
76, 2253 (2000).
152. V. Klimov, D. McBranch, N. Barashkov, and J. Ferraris, SPIE
3145, 58 (1997).
153. E. S. Maniloff, V. I. Klimov, and D. W. McBranch, Phys. Rev. B
56, 1876 (1997).
154. V. I. Klimov, D. W. McBranch, N. Barashkov, and J. Ferraris, Phys.
Rev. B 58, 7654 (1998).
155. V. I. Klimov, D. W. McBranch, N. N. Barashkov, and J. P. Ferraris,
Chem. Phys. Lett. 277, 109 (1997).
156. D. W. McBranch, B. Kraabel, S. Xu, R. S. Kohlman, V. I. Klimov,
D. D. C. Bradley, B. R. Hsieh, and M. Rubner, Synth. Met. 101,
291 (1999).
157. B. Kraabel, V. I. Klimov, R. Kohlman, S. Xu, H.-L. Wang, and
D. W. McBranch, Phys. Rev. B 61, 8501 (2000).
158. S. C. J. Meskers, R. A. J. Janssen, J. E. M. Haverkort, and J. H.
Wolter, Chem. Phys. 260, 415 (2000).
159. W. F. Mooney III, P. E. Brown, J. C. Russell, S. B. Costa, L. G.
Pedersen, and D. G. Whitten, J. Am. Chem. Soc. 106, 5659 (1984).
160. W. F. Mooney and D. G. Whitten, J. Am. Chem. Soc. 108, 5712
(1986).
161. S. P. Spooner and D. G. Whitten, J. Am. Chem. Soc. 116, 1240
(1994).
162. J. Heesemann, J. Am. Chem. Soc. 102, 2167 (1980).
163. S. Vaday, H. C. Geiger, B. Cleary, J. Perlstein, and D. G. Whitten,
J. Phys. Chem. B 101, 321 (1997).
164. D. G. Whitten, Acc. Chem. Res. 26, 502 (1993).
165. M. Era, J. Koganemaru, T. Tsutsui, A. Watakabe, and T. Kunitake,
Synth. Met. 91, 83 (1997).
166. A. Watakabe and T. Kunitake, J. Colloid Interface Sci. 145, 90
(1991).
167. A. Watakabe and T. Kunitake, Chem. Lett. 1991, 905 (1991).
168. A. Watakabe and T. Kunitake, Thin Solid Films 186, L21 (1990).
169. A. Watakabe, H. Okoda, and T. Kunitake, Langmuir 10, 2722
(1994).
170. P. Jonkheijm, M. Fransen, A. P. H. J. Schenning, and E. W. Meijer,
J. Chem. Soc., Perkin Trans. 2 2001, 1280 (2001).
171. C. J. Collison, V. Treemaneekarn, W. J. Oldham, Jr., J. H. Hsu,
and L. J. Rothberg, Synth. Met. 119, 515 (2001).
172. B.-K. An, S.-K. Kwon, S.-D. Jung, and S. Y. Park, J. Am. Chem.
Soc. 124, 14410 (2002).
173. E. Peeters, A. M. Ramos, S. C. J. Meskers, and R. A. J. Janssen,
J. Chem. Phys. 112, 9445 (2000).

174. J. Gierschner, Dissertation, Tubingen,


2000.

235
175. D. Oelkrug, H.-J. Egelhaaf, J. Gierschner, and A. Tompert, Synth.
Met. 76, 249 (1996).
176. J. Gierschner, H.-J. Egelhaaf, and D. Oelkrug, Synth. Met. 84, 529
(1997).

177. J. Gierschner, H.-J. Egelhaaf, D. Oelkrug, and K. Mullen,


J. Fluoresc. 8, 37 (1998).
178. H.-J. Egelhaaf, J. Gierschner, and D. Oelkrug, Synth. Met. 127,
221 (2002).
179. I. A. Vasileva, M. D. Galanin, A. N. Nikitina, and Z. A.
Chizhikova, Opt. Spektrosk. 60, 976 (1986).
180. B. Brocklehurst, D. C. Bull, M. Evans, P. M. Scott, and G. Stanney,
J. Am. Chem. Soc. 97, 2977 (1975).
181. K. Henderson, A. B. Dalton, G. Chambers, A. Drury, S. Maier,
A. G. Ryder, W. Blau, and H. J. Byrne, Synth. Met. 119, 555 (2001).
182. J. Cataln, L. Zimnyi, and J. Saltiel, J. Am. Chem. Soc. 122, 2377
(2000).
183. D. Horn and J. Rieger, Angew. Chem. Int. Ed. 40, 4330 (2001).
184. H. Kasai, H. S. Nalwa, S. Okada, H. Oikawa, and H. Nakanish,
in Handbook of Nanostructured Materials and Nanotechnology
(H. S. Nalwa, Ed.), Vol. 5, Chap. 8, pp. 43304361. Academic
Press, New York, 2000.
185. M. D. Cohen, B. S. Green, Z. Ludmer, and G. M. J. Schmidt,
Chem. Phys. Lett. 7, 486 (1979).
186. R. M. Hochstrasser, J. Mol. Spectrosc. 8, 485 (1962).
187. A. F. Prikhotko and I. Fugol, Opt. Spektrosc. 7, 19 (1959).
188. C. C. Wu, O. J. Korovyanko, M. C. DeLong, Z. V. Vardeny, and
J. P. Ferraris, Synth. Met. 139, 735 (2003).
189. R. A. Agbaria, E. Roberts, and I. M. Warner, J. Phys. Chem. 99,
10056 (1995).

190. J. Gierschner, L. Luer,


D. Oelkrug, E. Musluoglu, B. Behnisch,
and M. Hanack, Adv. Mater. 12, 757 (2000).

191. J. Gierschner, L. Luer,


D. Oelkrug, E. Musluoglu, B. Behnisch,
and M. Hanack, Synth. Met. 121, 1695 (2001).
192. R. H. Dyck and D. S. McClure, J. Chem. Phys. 36, 2326 (1962).
193. J. A. Syage, P. M. Felker, and A. H. Zewail, J. Chem. Phys. 81,
4685 (1984).
194. Z. Arp, W.-Y. Chiang, A. Sakamoto, and M. Tasumi, J. Phys. Chem.
A 106, 3479 (2002).
195. R. A. Rijkenberg, D. Bebelaar, W. J. Buma, and J. W. Hofstraat,
J. Phys. Chem. A 106, 2446 (2002).
196. T. Urano, H. Hamaguchi, M. Tasumi, K. Yamanouchi, S. Tsuchiya,
and T. L. Gustafson, J. Chem. Phys. 91, 3884 (1989).
197. (a) R. Mahrt, J. Yang, A. Greiner, and H. Bassler, Makromol.
Chem. 11, 415 (1990); (b) S. Heun, R. F. Mahrt, A. Greiner,
U. Lemmer, H. Bassler, D. A. Halliday, D. D. C. Bradley, P. L.
Burns, and A. B. Holmes, J. Phys. Condens. Matter 5, 247 (1993).

198. T. Pauck, H. Bassler, J. Grimme, U. Scherf, and K. Mullen,


Chem.
Phys. 210, 219 (1996).
199. F. Negri and M. Z. Zgierski, J. Chem. Phys. 100, 2571 (1994).
200. S. Karabunarliev and E. R. Bittner, J. Chem. Phys. 118, 4291
(2003).

201. G. Baranovic, Z. Meic, H. Gusten,


J. Mink, and G. Keresztury,
J. Phys. Chem. 94, 2833 (1990).

202. Z. Meic and H. Gusten,


Spectrochim. Acta 34A, 101 (1978).
203. A. Bree and R. Zwarich, Spectrochim. Acta 38A, 719 (1982).
204. X. Ci and A. B. Myers, Chem. Phys. Lett. 158, 263 (1989).
205. S. L. Schultz, J. Qian, and J. M. Jean, J. Phys. Chem. A 101, 1000
(1997).
206. K. Palmo, Spectrochim. Acta 44A, 341 (1988).
207. T. Nakabayashi, H. Okomoto, and M. Tasumi, J. Phys. Chem. A
102, 9686 (1998).
208. T. Nakabayashi, H. Okomoto, and M. Tasumi, J. Phys. Chem. A
101, 7189 (1997).
209. I. Orion, J. P. Buisson, and S. Lefrant, Phys. Rev. B 57, 7050 (1998).
210. T. P. Nguyen, V. H. Tran, P. Destruel, and D. Oelkrug, Synth. Met.
101, 633 (1999).

236
211. V. Hernandez, C. Castiglioni, M. Del Zoppo, and G. Zerbi, Phys.
Rev. B 50, 9815 (1994).
212. A. B. Myers, M. O. Trulson, and R. A. Mathies, J. Chem. Phys. 83,
5000 (1985).

213. S. Karabunarliev, M. Baumgarten, E. R. Bittner, and K. Mullen,


J. Chem. Phys. 113, 11372 (2000).
214. E. Mulazzi, A. Ripamonti, J. Wery, B. Dulieu, and S. Lefrant,
Phys. Rev. B 60, 16519 (1999).
215. Y. Ichino, J. P. Ni, Y. Ueda, and D. K. Wang, Synth. Met. 116, 223
(2001).
216. J. Cornil, D. Beljonne, C. M. Heller, I. Campbell, B. K. Laurich,

D. L. Smith, D. D. C. Bradley, K. Mullen,


and J. L. Brdas, Chem.
Phys. Lett. 278, 139 (1997).
217. J. Cornil, D. Beljonne, Z. Shuai, T. W. Hagler, I. Campbell,

D. D. C. Bradley, J. L. Brdas, C. W. Spangler, and K. Mullen,


Chem. Phys. Lett. 247, 425 (1995).

218. H.-J. Egelhaaf, L. Luer,


A. Tompert, P. Bauerle, K. Mullen,
and
D. Oelkrug, Synth. Met. 115, 63 (2000).
219. J. Dale, Acta Chem. Scand. 11, 971 (1957).
220. A. Warshel, J. Chem. Phys. 62, 214 (1975).
221. D. Oelkrug, S. Reich, F. Wilkinson, and P. A. Leicester, J. Phys.
Chem. 95, 269 (1991).
222. J. Gierschner, H.-G. Mack, H.-J. Egelhaaf, S. Schweizer, B. Doser,
and D. Oelkrug, Synth. Met. 138, 311 (2003).
223. T. E. Bush and G. W. Scott, J. Phys. Chem. 85, 146 (1981).
224. L. Onsager, J. Am. Chem. Soc. 58, 1486 (1936).
225. E. Lippert, Z. Elektrochem. 61, 962 (1957).
226. W. Liptay, Z. Naturforsch. 20a, 1441 (1965).
227. H. Gruen and H. Gorner, J. Phys. Chem. 93, 7144 (1989).
228. B. Boldrini, E. Cavalli, A. Painelli, and F. Terenziani, J. Phys.
Chem. A 106, 6286 (2002).
229. C. W. M. Castleton and W. Barford, J. Chem. Phys. 117, 3570
(2002).
230. W. G. Han, T. Lovell, T. Liu, and L. Noodleman, Chem. Phys.
Chem. 2002, 167 (2002).
231. J. Fabian, L. A. Diaz, G. Seifert, and T. Niehaus, J. Mol. Struct.
(Theochem.) 594, 41 (2002).
232. G. P. Das, A. T. Yeates, and D. S. Dudis, Chem. Phys. Lett. 361, 71
(2002).
233. E. Zojer, D. Beljonne, T. Kogej, H. Vogel, S. R. Marder, J. W.
Perry, and J. L. Brdas, J. Chem. Phys. 116, 3646 (2002).

234. D. Oelkrug, J. Gierschner, H.-J. Egelhaaf, L. Luer,


A. Tompert,

K. Mullen,
U. Stalmach, and H. Meier, Synth. Met. 121, 1693
(2001).
235. X. Zhou, A.-M. Ren, J.-K. Feng, and X.-J. Liu, Chem. Phys. Lett.
362, 541 (2002).
236. F. Guo, Chem. Phys. Lett. 355, 89 (2002).
237. D. Oelkrug, A. Tompert, H.-J. Egelhaaf, M. Hanack, E. Steinhuber, H. Meier, and U. Stalmach, Synth. Met. 83, 231 (1996).
238. J. Cornil, D. A. dos Santos, D. Beljonne, and J. L. Brdas, Synth.
Met. 99, 5604 (1995).
239. Y. Shuto, Int. J. Quantum Chem. 58, 407 (1996).
240. J. Obrzut and F. E. Karasz, J. Chem. Phys. 87, 2349 (1987).
241. J. Obrzut and F. E. Karasz, J. Chem. Phys. 87, 6178 (1987).
242. J. Cornil, D. A. dos Santos, X. Crispin, R. Silbey, and J. L. Brdas,
J. Am. Chem. Soc. 120, 1289 (1998).
243. J. Cornil, D. Beljonne, and J. L. Brdas, J. Chem. Phys. 103, 834
(1995).
244. M. Chandross, S. Mazumdar, M. Liess, P. A. Lane, Z. V. Vardeny,
M. Hamaguchi, and K. Yoshino, Phys. Rev. B 55, 1486 (1997).
245. A. Pogantsch, G. Heimel, and E. Zojer, J. Chem. Phys. 117, 5921
(2002).
246. J. B. Lagowski, J. Mol. Struct. (Theochem.) 589590, 125 (2002).
247. W. B. Davis, M. R. Wasielewski, and M. A. Ratner, Int. J. Quantum. Chem. 72, 463 (1999).
248. J. Yu, W. S. Fann, F. J. Kao, D. Y. Yang, and S. H. Lin, Synth.
Met. 66, 143 (1994).

Optical Properties of Oligophenylenevinylenes


249. W. Barford and R. J. Bursill, Synth. Met. 89, 155 (1997).
250. T. Wagenstreiter and S. Mukamel, J. Chem. Phys. 104, 7086 (1996).
251. J. L. Brdas, J. Cornil, D. Beljonne, D. A. dos Santos, and
Z. Shuai, Acc. Chem. Res. 32, 267 (1999).
252. (a) B. E. Kohler, J. Chem. Phys. 93, 5838 (1990); (b) J. E. LennardJones, Proc. R. Soc. London, Ser. A 158, 280 (1937).
253. (a) H. Kuhn, Fortschr. Chem. Org. Naturst. 16, 169 (1958);
(b) H. Kuhn, Fortschr. Chem. Org. Naturst. 17, 404 (1959).
254. W. T. Simpson, J. Am. Chem. Soc. 77, 6164 (1955).
255. W. Kuhn, Helv. Chim. Acta 31, 1780 (1948).
256. H.-K. Ryu, W. Y. Kim, K. S. Nahm, Y. B. Hahn, Y.-S. Lee, and
C. Lee, Synth. Met. 128, 21 (2002).
257. H. S. Woo, O. Lhost, S. C. Graham, D. D. C. Bradley, R. H.

Friend, C. Quattrocchi, J. L. Brdas,, R. Schenk, and K. Mullen,


Synth. Met. 59, 13 (1993).
258. C. M. Heller, I. H. Campbell, B. K. Laurich, D. L. Smith, D. D. C.

Bradley, P. L. Burn, J. P. Ferraris, and K. Mullen,


Phys. Rev. B 54,
5516 (1996).
259. J. Cornil, D. Beljonne, D. A. dos Santos, and J. L. Brdas, Synth.
Met. 76, 101 (1996).
260. L. Chiavarona, M. Di Terlizzi, G. Scamarcio, F. Babudri, G. M.
Farinola, and F. Naso, Appl. Phys. Lett. 75, 2053 (1999).
261. D. A. dos Santos, D. Beljonne, J. Cornil, and J. L. Brdas, Chem.
Phys. 227, 1 (1998).
262. P. Usnanski, M. Kryszewski, and E. W. Thulstrup, Spectrochim.
Acta 46A, 23 (1990).
263. A. Yogev and L. Margulies, Isr. J. Chem. 16, 258 (1977).
264. (a) T. Damerau and M. Hennecke, J. Chem. Phys. 103, 6232 (1995);

(b) M. Hennecke, T. Damerau, and K. Mullen,


Macromolecules 26,
3411 (1993).
265. M. S. Gudipati, M. Maus, J. Daverkausen, and G. Hohlneicher,
Chem. Phys. 192, 37 (1995).
266. F. C. Spano and S. Siddiqui, Chem. Phys. Lett. 314, 481 (1999).
267. G. Hohlneicher and B. Dick, J. Photochem. 27, 215 (1984).
268. M. Rumi, J. E. Ehrlich, A. A. Heikal, J. W. Perry, S. Barlow, Z. Hu,
D. McCord-Moughon, T. C. Parker, H. Rockel, S. Thayumanavan,
S. R. Marder, D. Beljonne, and J.-L. Brdas, J. Am. Chem. Soc.
112, 9500 (2000).
269. D. Beljonne, Z. Shuai, J. Cornil, D. A. dos Santos, and J. L. Brdas, J. Chem. Phys. 111, 2829 (1999).
270. S. J. K. Pond, M. Rumi, M. D. Levin, T. C. Parker, D. Beljonne,
M. W. Day, J.-L. Brdas, S. R. Marder, and J. W. Perry, J. Phys.
Chem. A 106, 11470 (2002).
271. B. Strehmel, A. M. Sarker, and H. Detert, Chem. Phys. Chem. 4,
249 (2003).
272. S. Mukamel, S. Tretiak, T. Wagersreiter, and V. Chernyak, Science
277, 781 (1997).
273. F. S. Dainton, C. T. Peng, and G. A. Salmon, J. Phys. Chem. 72,
3801 (1968).
274. L. P. Candeias, J. Wildeman, G. Hadziioannou, and J. M. Warman,
J. Phys. Chem. B 104, 8366 (2000).
275. L. P. Candeias, G. H. Gelinck, J. J. Piet, J. Piris, B. Wegewijs,

E. Peeters, J. Wildeman, G. Hadziioannou, and K. Mullen,


Synth.
Met. 119, 339 (2001).
276. D. Oelkrug, K. Rempfer, E. Prass, and H. Meier, Z. Naturforsch.
43a, 583 (1988).
277. M. M. de Souza, G. Rumbles, I. D. W. Samuel, S. C. Moratti, and
A. B. Holmes, Synth. Met. 101, 631 (1999).
278. U. Mazzucato, Pure Appl. Chem. 54, 1705 (1982).
279. J. Saltiel and J. T. DAgostino, J. Am. Chem. Soc. 94, 6445
(1972).
280. S. Sharafy and K. A. Muszkat, J. Am. Chem. Soc. 93, 4119
(1971).
281. (a) J. L. Charlton and J. Saltiel, J. Phys. Chem. 81, 1940 (1977); (b)
J. Saltiel, A. S. Waller, D. F. Sears Jr., and C. Z. Garett, J. Phys.
Chem. 97, 2516 (1993); (c) J. Saltiel, A. S. Waller, D. F. Sears, Jr.,

Optical Properties of Oligophenylenevinylenes

282.
283.
284.
285.

286.
287.
288.
289.
290.
291.
292.
293.

294.

295.
296.
297.
298.
299.
300.
301.
302.
303.
304.
305.
306.
307.
308.
309.
310.
311.

E. A. Hoburg, D. M. Zeglinski, and D. H. Waldeck, J. Phys. Chem.


98, 10689 (1994).
J. Saltiel, O. C. Zariou, E. D. Megarity, and A. A. Lamola, J. Am.
Chem. Soc. 90, 4759 (1968).
D. Oelkrug, A. Tompert, H. Maier, and U. Stalmach, unpublished
results.
H. Tillmann and H.-H. Horhold, Synth. Met. 101, 138
(1999).
(a) A. Colombo and G. Allegra, Macromolecules 5, 579 (1971); (b)

O. Konig, H.-B. Burgi,


T. Armbruster, J. Hulliger, and T. Weber,
J. Am. Chem. Soc. 119, 10632 (1997); (c) G. Bongiovanni, C. Botta,
J. L. Brdas, J. Cornil, D. R. Ferro, A. Mura, A. Piaggi, and
R. Tubino, Chem. Phys. Lett. 278, 146 (1997).
M. Farina, G. de Silvestro, and P. Sozzani, in Comprehensive Supramolecular Chemistry (D. D. McNicol, F. Toda, and
R. Bishop, Eds.), Vol. 6. Elsevier, Oxford, 1996.
N. Gfeller and G. Calzaferri, J. Phys. Chem. B 101, 1396
(1997).
D. Beljonne, J. Cornil, R. Silbey, P. Milli, and J. L. Brdas,
J. Chem. Phys. 112, 4749 (2000).
J. Cornil, D. Beljonne, D. A. dos Santos, J. Ph. Calbert, and J. L.
Brdas, Thin Solid Films 363, 72 (2000).
J. Cornil, D. Beljonne, J.-P. Calbert, and J. L. Brdas, Adv. Mater.
13, 1053 (2001).
S. Tretiak and S. Mukamel, Chem. Rev. 102, 3171 (2002).
S. Tretiak, A. Saxena, R. L. Martin, and A. R. Bishop, J. Phys.
Chem. B 104, 7029 (2000).
(a) M. Kasha, Rev. Mod. Phys. 31, 162 (1959); (b) R. M.
Hochstrasser and M. Kasha, Photochem. Photobiol. 3, 317 (1964);
(c) E. G. McRae and M. Kasha, J. Chem. Phys. 28, 721 (1958);
(d) M. Kasha, H. R. Rawls, and M. A. El-Bayoumi, Pure Appl.
Chem. 11, 371 (1965).
(a) V. Czikkely, H. D. Forsterling, and H. Kuhn, Chem. Phys. Lett.
6, 11 (1970); (b) V. Czikkely, H. D. Forsterling, and H. Kuhn,
Chem. Phys. Lett. 6, 207 (1970); (c) H. Nolte and V. Buss, Chem.
Phys. Lett. 19, 395 (1973); (d) H. Nakahara, K. Fukuda, D. Mo bius,
and H. Kuhn, J. Phys. Chem. 90, 6144 (1986); (e) C. E. Evans and
P. W. Bohn, J. Phys. Chem. 97, 12302 (1993).
M. J. McIntire, E. S. Manas, and F. C. Spano, J. Chem. Phys. 107,
8152 (1997).
E. S. Manas and F. C. Spano, J. Chem. Phys. 109, 8087 (1999).
Z. G. Soos, G. W. Hayden, P. C. M. McWilliams, and S. Etemad,
J. Chem. Phys. 93, 7439 (1990).
J. Gierschner, H.-G. Mack, D. Oelkrug, I. Waldner, and H. Rau,
J. Phys. Chem. A, in press.
S. Canuto and M. C. Zerner, J. Am. Chem. Soc. 112, 2114 (1990).
S. Canuto and M. C. Zerner, Chem. Phys. Lett. 157, 353 (1989).

D. Oelkrug, H.-J. Egelhaaf, B. Lehr, J. Gierschner, L. Luer,


P. Matousek, and M. Towrie, Tech. Rep.Counc. Cent. Lab. Res.
Counc. 1999, 1 (1999).
T. Damerau and M. Hennecke, J. Polym. Sci. 33, 2219 (1995).
R. E. Gill, G. Hadziioannou, P. Lang, F. Garnier, and J. C.
Wittmann, Adv. Mater. 9, 331 (1997).
A. Menon, H. Dong, Z. I. Niazimbetova, L. J. Rothberg, and M. E.
Galvin, Chem. Mater. 14, 3668 (2002).
R. E. Gill, A. Meetsma, and G. Hadziioannou, Adv. Mater. 8, 212
(1996).
P. F. van Hutten, V. V. Krasnikov, H.-J. Brouwer, and G. Hadziioannou, Chem. Phys. 241, 139 (1999).
L. Claes, J.-P. Franois, and M. S. Deleuze, Chem. Phys. Lett. 339,
216 (2001).

G. Cerullo, G. Lanzani, S. De Silvestri, H.-J. Egelhaaf, L. Luer,


and D. Oelkrug, Phys. Rev. B 62, 2429 (2000).
N. F. Colaneri, D. D. C. Bradley, R. H. Friend, P. L. Burn, A. B.
Holmes, and C. W. Spangler, Phys. Rev. B 42, 11670 (1990).
F. C. Spano, Chem. Phys. Lett. 331, 7 (2000).
F. C. Spano, Synth. Met. 116, 339 (2001).

237
312. (a) F. C. Spano, J. Chem. Phys. 114, 5376 (2001); (b) F. C. Spano,
J. Chem. Phys. 117, 9961 (2002).
313. F. C. Spano, J. Chem. Phys. 116, 5877 (2002).
314. F. C. Spano, J. Chem. Phys. 118, 981 (2003).

315. L. Luer,
H.-J. Egelhaaf, and D. Olekrug, in preparation.
316. S. Brazovskii, A. Kirova, A. R. Bishop, V. Klimov, D. McBranch,
N. N. Barashkov, and J. P. Ferraris, Opt. Mater. 9, 472 (1998).
317. N. Kirova, S. Brazovskii, A. R. Bishop, D. McBranch, and
V. Klimov, Synth. Met. 101, 188 (1999).
318. D. Jones, M. Guerra, L. Favaretto, A.Modelli, M. Fabrizio, and
G. Distefano, J. Phys. Chem. 94, 5761 (1990).
319. J. Gierschner, H.-J. Egelhaaf, D. Oelkrug, R. Martinez Alvarez,
M. Hanack, D. Nadele, and J. Strahle, in preparation.
320. M. D. Joswick, I. H. Campbell, N. N. Barashkov, and J. P. Ferraris,
J. Appl. Phys. 80, 2883 (1996).
321. J. Gierschner, H.-J. Egelhaaf, D. Oelkrug, H. Benmansour, and
G. Bazan, in preparation.
322. K. Iwata, R. Ozawa, and H. Hamaguchi, J. Phys. Chem. A 106,
3614 (2002).
323. (a) Y. Dou and R. E. Allen, Chem. Phys. Lett. 378, 323 (2003), (b)
Y. Dou and R. E. Allen, J. Chem. Phys. 119, 10658 (2003).
324. J.-H. Perng, J. Fluorescence 12, 311 (2002).
325. K. Furuya, K. Kawato, H. Yokoyama, A. Sakamato, and
M. Tasumi, J. Phys. Chem. A 107, 8251 (2003).
326. S. P. Kwasniewski, L. Claes, J.-P. Francois, and M. S. Deleuze,
J. Chem. Phys. 118, 7823 (2003).
327. Q.-L. Fan, S. Lu, Y.-H. Lai, X.-Y. Hou, and W. Huang, Macromolecules, 36, 6976 (2003).
328. O. Narwark, A. Gerhard, S. C. J. Meskers, S. Brocke, E. ThornCsnyi, and H. Bssler, Chem. Phys. 294, 17 (2003).
329. O. Narwark, S. C. J. Meskers, R. Peetz, E. Thorn-Csnyi, and
H. Bssler, Chem. Phys. 294, 1 (2003).
330. Z. I. Niazimbetova, A. Menon, M. E. Galvin, and D. H. Evans,
J. Electroanalyt. Chem. 529, 43 (2002).
331. Y. Chen and S.-P. Lai, J. Polym. Sci. A 39, 2571 (2001).
332. H. Detert and E. Sugiono, Synth. Met. 138, 181 (2003).
333. H. Detert and O. Sadovski, Synth. Met. 138, 185 (2003).
334. C.-L. Li, S.-J. Shieh, S.-C. Lin, and R.-S. Liu, Org. Letters 5, 1131
(2003).
335. A. M. Sarker, Y. Kaneko, P. M. Lahti, and F. E. Karasz, J. Phys.
Chem. A 107, 6533 (2003).
336. M. S. Wong, Z. H. Li, Y. Tao, and M. DIorio, Chem. Mater. 15,
1198 (2003).
337. H. Meier, J. Gerold, and D. Jacob, Tetrahedron Letters 44, 1915
(2003).
338. S. Wachsmann-Hogiu, L. A. Peteanu, L. A. Liu, D. J. Yaron, and
J. Wildeman, J. Phys. Chem. B 107, 5133 (2003).
339. S. P. Kwasniewski, J. P. Francois, and M. S. Deleuze, J. Phys. Chem.
A 107, 5168 (2003).
340. M. A. Summers, P. R. Kemper, J. E. Bushnell, M. R. Robinson,
G. C. Bazan, M. T. Bowers, and S. K. Buratto, J. Am. Chem. Soc.
125, 5199 (2003).
341. J.-S. K. Yu, W.-C. Chen, and C.-H. Yu, J. Phys. Chem. A 107, 4268
(2003).
342. F. C. Grozema, R. Telesca, J. G. Snijders, and L. D. A. Siebbeles,
J. Chem. Phys. 118, 9441 (2003).
343. P. C. Chen and Y. C. Chieh, J. Mol. Struct. Theochem. 624, 191
(2003).
344. B.-C. Wang, J.-C. Chang, J.-H. Pan, C. Xue, and F.-T. Luo, J. Mol.
Struct. Theochem. 636, 81 (2003).
345. J. B. Lagowski, J. Mol. Struct. Theochem. 634, 243 (2003).
346. B. J. Zhang and S.-J. Jeon, Chem. Phys. Lett. 377, 210 (2003).
347. F. Guo and Z. Y. Shih, Chem. Phys. Lett. 370, 572 (2003).
348. W. Bartkowiak, R. Zalesny, and J. Leszczynski, Chem. Phys. 287,
103 (2003).
349. F. He, G. Cheng, H. Zhang, Y. Zheng, Z. Xie, B. Yang, Y. Ma,
S. Liu, and J. Shen, Chem. Commun., 2206 (2003).

238
350. M. Egli, V. Tereshko, G. N. Mushudov, R. Sanishvili, X. Liu, and
F. D. Lewis, J. Am. Chem. Soc. 125, 10842 (2003).
351. Y. Wang, M. I. Ranasinghe, and T. Goodson, III, J. Am. Chem.
Soc. 125, 9562 (2003).
352. W. Dermaut, C. Wuyts, G. Aerts, E. Goovaerts, and H. J. Geise,
Synth. Met. 135136, 249 (2003).
353. (a) L. M. Herz, C. Daniel, C. Silva, F. J. M. Hoeben, A. P. H. J.
Schenning, E. W. Meijer, R. H. Friend, and R. T. Phillips, Synth.
Met. 139, 839 (2003); (b) L. M. Herz, C. Daniel, C. Silva, F. J. M.
Hoeben, A. P. H. J. Schenning, E. W. Meijer, R. H. Friend, and
R. T. Phillips, Phys. Rev. B 64, 045203 (2003).
354. (a) B. R. Crenshaw and C. Weder, Chem. Mater. (2003) in press;
(b) C. Lwe and C. Weder, Adv. Mater. 14, 1625 (2003).
355. R. Capelli, M. A. Loi, C. Taliani, H. B. Hansen, M. Murgia,
G. Ruani, M. Muccini, P. W. Lovenich, and W. J. Feast, Synth. Met.
139, 909 (2003).
356. M. A. Summers, M. R. Robinson, G. C. Bazan, and S. K. Buratto,
Synth. Met. 137, 957 (2003).

Optical Properties of Oligophenylenevinylenes


357. J. De Ceuster, E. Goovaerts, A. Bouwen, and V. Dyakonov, Phys.
Rev. B 68, 125202 (2003).
358. J. E. Wong, M. S. Weaver, T. Richardson, D. D. C. Bradley, and
D. W. Bruce, Mater. Sci. Engineer. C 22, 393 (2002).
359. J. E. Wong, S. Schrader, H. Detert, S. Katholy, and L. Brehmer,
Mater. Sci. Engineer. C 22, 413 (2002).
360. C. Melzer, V. V. Krasnikov, and G. Hadziioannou, J. Polym. Sci. B
41, 2665 (2003).
361. A. Yoshiki, N. Matsuoka, M. Kondo, and H. Yanagi, Thin Solid
Films 438439, 308 (2003).
362. M. Hanack, B. Behnisch, H. Hckl, P. Martinez-Ruiz, and K.-H.
Schweikart, Thin Solid Films 417, 26 (2002).
363. A. Ajayaghosh, S. J. George, and Y. K. Praveen, Angew. Chem.
Int. Ed. 42, 332 (2003).
364. P. A. van Hal, M. M. Wienk, J. M. Kroon, and R. A. J. Janssen,
J. Mater. Chem. 13, 1054 (2003).
365. F. Stellacci, C. A. Bauer, T. Meyer-Friedrichsen, W. Wenseleers,
S. R. Marder, and J. W. Perry. J. Am. Chem. Soc. 125, 328 (2003).

Vous aimerez peut-être aussi