0% found this document useful (0 votes)
166 views27 pages

Spano Silva 2014 H and J Aggregate Behavior in Polymeric Semiconductors

This document summarizes research on H- and J-aggregate behavior in polymeric semiconductors. It discusses how intrachain and interchain electronic interactions in conjugated polymer films can be understood using concepts of J- and H-aggregation originally developed for small molecule aggregates. Intrachain interactions lead to J-aggregate behavior and favor photoluminescence, whereas interchain interactions lead to H-aggregate behavior. The document reviews formalisms for describing absorption and emission lineshapes based on intra- and intermolecular excitonic coupling, electron-vibrational coupling, and energetic disorder in polymers.

Uploaded by

cyunqi926
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
166 views27 pages

Spano Silva 2014 H and J Aggregate Behavior in Polymeric Semiconductors

This document summarizes research on H- and J-aggregate behavior in polymeric semiconductors. It discusses how intrachain and interchain electronic interactions in conjugated polymer films can be understood using concepts of J- and H-aggregation originally developed for small molecule aggregates. Intrachain interactions lead to J-aggregate behavior and favor photoluminescence, whereas interchain interactions lead to H-aggregate behavior. The document reviews formalisms for describing absorption and emission lineshapes based on intra- and intermolecular excitonic coupling, electron-vibrational coupling, and energetic disorder in polymers.

Uploaded by

cyunqi926
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

PC65CH21-Silva ARI 10 March 2014 15:47

ANNUAL
REVIEWS Further H- and J-Aggregate Behavior
Click here for quick links to
Annual Reviews content online,
including:
in Polymeric Semiconductors
• Other articles in this volume
• Top cited articles Frank C. Spano1 and Carlos Silva2,3
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

• Top downloaded articles


• Our comprehensive search 1
Department of Chemistry, Temple University, Philadelphia, Pennsylvania 19122;
email: [email protected]
2
Département de physique et Regroupement québécois sur les matériaux de pointe,
Université de Montréal, Montréal H3C 3J7, Canada; email: [email protected]
3
Department of Physics and Center for Plastic Electronics, Imperial College London,
London SW7 2AZ, United Kingdom

Annu. Rev. Phys. Chem. 2014. 65:477–500 Keywords


First published online as a Review in Advance on conjugated polymers, H-aggregates, J-aggregates
January 9, 2014

The Annual Review of Physical Chemistry is online at Abstract


physchem.annualreviews.org
Aggregates of conjugated polymers exhibit two classes of fundamental
This article’s doi: electronic interactions: those occurring within a given chain and those
10.1146/annurev-physchem-040513-103639
occurring between chains. The impact of such excitonic interactions on the
Copyright  c 2014 by Annual Reviews. photophysics of polymer films can be understood using concepts of J- and
All rights reserved
H-aggregation originally developed by Kasha and coworkers to treat aggre-
gates of small molecules. In polymer assemblies, intrachain through-bond
interactions lead to J-aggregate behavior, whereas interchain Coulombic
interactions lead to H-aggregate behavior. The photophysics of common
emissive conjugated polymer films are determined by a competition between
intrachain, J-favoring interactions and interchain, H-favoring interactions.
We review formalisms describing absorption and photoluminescence
lineshapes, based on intra- and intermolecular excitonic coupling, electron-
vibrational coupling, and correlated energetic disorder. Examples include
regioregular polythiophenes, pheneylene-vinylenes, and polydiacetylene.

477
PC65CH21-Silva ARI 10 March 2014 15:47

1. INTRODUCTION
The promise of highly efficient solar cells (1–18) and light-emitting diodes (19–23) based on
Exciton: collective semiconductor π -conjugated polymers has driven the need for a more fundamental understand-
electronic excited ing of charge and energy transport in soft organic assemblies, characterized by significant coupling
states in an assembly of between the delocalized neutral (or charged) electronic excitations and primarily intramolecular
interacting
vibrations. The coupling results in electronic excited states surrounded by a field of vibrationally
chromophores
distorted chromophores (in polymers, the chromophores can be monomeric repeat units or groups
J-aggregate:
thereof ). Exciton-vibrational coupling increases the exciton effective mass and enhances exciton
assembly of
chromophores in localization due to the wide range of defects normally encountered in organic materials. A rigorous
which the optically description of the nature of this primary photoexcitation in polymeric-semiconductor architec-
allowed exciton resides tures is essential for the development of applications in optoelectronics and is fundamentally
at the bottom of the important in their materials physics. Here, our objective is to review recent advances in the under-
band
Access provided by National Cheng Kung University on 12/18/23. For personal use only.

standing of steady-state optical signatures of semicrystalline polymeric semiconductors in terms


Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

H-aggregate: of photophysical aggregate models, focusing on the spectral signatures of intra- and interchain
assembly of
excitonic coupling, exciton-vibrational coupling, and energetic disorder.
chromophores in
which the optically Relatively recent research (24–28) has shown that the photophysics of conjugated polymer
allowed exciton resides assemblies can be understood using the ideas developed decades ago by Kasha and coworkers
at the top of the band (29–31) for understanding absorption and fluorescence in aggregates of much smaller molecular
chromophores. The classification of H- versus J-aggregation was a key development in unra-
veling the relationship between morphology and photophysical properties. The two aggregate
types are depicted in Figure 1. In J-aggregates, neighboring chromophores are oriented in a
head-to-tail fashion, resulting in a negative excitonic coupling and the placement of the optically
allowed (k = 0) Frenkel exciton at the bottom of the exciton band. Conversely, in H-aggregates,
nearest-neighbor chromophores are oriented in a more side-by-side manner, resulting in a positive
excitonic coupling and the placement of the k = 0 exciton at the top of the exciton band. The
energetic ordering of the excitons has profound effects on the photophysical response: In J- (H-)
aggregates, the main S0 → S1 absorption peak undergoes a red (blue) shift compared to the
spectrum of an isolated monomer. Moreover, the excited-state radiative decay rate is enhanced in
a J-aggregate and suppressed in an H-aggregate.
Unlike in small molecules, where the lowest electronically excited state S1 is typically separated
by ∼0.1–1 eV from the next higher electronic excited state, the lowest electronic states in poly-
mers are organized into a band of intrachain excitonic states, in which the energetic separation
between neighboring states ideally tends toward zero as the polymer length increases. Hence, in
describing the photophysics of polymer assemblies—for example, the lamellar π -stacks depicted in
Figure 2a—one must treat not only interchain coupling, as is the usual consideration for small
molecular aggregates within the Kasha formalism, but also intrachain coupling. In other words,
polymer π -stacks are inherently two-dimensional (2D) excitonic systems with electronic excita-
tions delocalized along the polymer chain as well as between chains. Such excitons possess a dual
nature; within polymer chains, excitons are of the Wannier-Mott type, where the electron and hole
can readily separate over several repeat units, whereas across chains, excitons are more Frenkel-
like as charge separation is less likely and is limited to at most neighboring chains, resulting in the
so-called interchain polarons (32, 33).
Despite the complex nature of excitons in polymer assemblies, we demonstrate below that the
related photophysics can be understood using the concepts of J- and H-aggregation. In doing so,
we employ a new set of spectral signatures with which to distinguish J- and H-aggregation based
on the strong vibronic coupling between the S0 → S1 optical transition and the ≈ 1,400 cm−1
vinyl/ring stretching modes found in virtually all π -conjugated molecules. Such coupling results

478 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

a J-aggregate b H-aggregate c HJ-aggregate


Jinter > 0

J0 > 0
Jintra < 0
J0 < 0

+
ΔE
Energy


Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

X 0-0
0-2
0-2 0-2
Two-phonon
ħω0 0-1 One-phonon 0-1 0-1

|G> 0-0
–π 0 π – 0 – 0
k,q k,q k,q

Figure 1
(a,b) Molecular orientations within conventional J- and H-aggregates. The sign of the nearest-neighbor
coupling J0 is determined by the through-space Coulombic coupling. Generally, head-to-tail orientations
lead to J0 < 0 and J-aggregation, whereas side-by-side orientations lead to J0 > 0 and H-aggregation.
(c) In polymer HJ-aggregates, Coulombic interchain coupling is positive (Jinter > 0), whereas the effective
intrachain coupling between adjacent repeat units is negative (Jintra < 0) owing to through-bond interactions
in 1D direct band-gap semiconductors. Also shown is the energy dispersion, E(k), corresponding to the lowest
vibronic band in each aggregate (higher bands are omitted for clarity). The band curvature at k = 0 is positive
(negative) in J- (H-)aggregates. The red dot indicates the (k = 0) exciton that is optically allowed from the
ground state, |G> (black dot). The energies of the one- and two-phonon states within the electronic ground
state are also indicated. The dispersionless (Einstein) phonons of wave vector q derive from the intramolecular
vibrations with frequency ω0 . Arrows indicate emission pathways at low temperatures, such that emission
originates primarily from the lowest-energy exciton. In J-aggregates, 0-0 emission is strongly allowed,
leading to superradiance. In contrast, in H-aggregates, rapid intraband relaxation subsequent to absorption
populates the lowest-energy k = π exciton, which cannot radiatively couple to |G>, thereby preventing
0-0 emission (assuming no disorder). In the HJ-dimer, the J-like intrachain band in each polymer is split
into symmetric (+) and antisymmetric (−) bands by interchain interactions. Owing to selection rules, only
the k = 0 symmetric state can radiatively couple to |G>. Hence, in HJ-aggregates, as well as H-aggregates,
0-0 emission is thermally activated. 0-0 emission is also made allowed by symmetry-breaking disorder.

in well-pronounced vibronic progressions in the absorption and photoluminescence (PL) spec-


tra of isolated chromophores, or repeat units in the case of single polymer chains. When such
chromophores interact electronically, the progression is distorted, and the relative peak intensi-
PL:
ties no longer follow the Poissonian distribution characteristic of an isolated chromophore. The photoluminescence
interaction-induced changes in the vibronic line strengths reveal information about the exciton
bandwidth, the nature of disorder, and the exciton coherence length (24, 34–36). Importantly, such

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 479


PC65CH21-Silva ARI 10 March 2014 15:47

a b
|A2>
J-like

|A1>
|em>
A2
16 Å A1

0-2
Normalized spectral intensity

H-like
3.8 Å
0-1

0-0
Access provided by National Cheng Kung University on 12/18/23. For personal use only.

|G>
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

S
0-1 π-π stacking direction
A2 C

A1 c 0-0

0-1
Intensity

0-2 0-0 Absorption

PL Absorption
T = 10 K
PL
T = 300 K

1.2 1.6 2.0 2.4 2.8 1.6 1.8 2.0 2.2 2.4 2.6 2.8
Photon energy (eV) Energy (eV)

Figure 2
(a) Absorption and photoluminescence (PL) spectra of a P3HT film cast from a chloroform solution. Panel a reprinted with permission
from Reference 27. Copyright 2009, AIP Publishing LLC. (Insets) P3HT π -stack and Jablonski diagram corresponding to the weakly
coupled H-aggregate model where 0-0 emission is allowed by disorder (118). (b) Graphical depiction of P3HT J- and H-aggregates.
(c) Absorption and PL spectra of P3HT nanofibers grown in toluene. Panel c reprinted with permission from Reference 56. Copyright
2012 American Chemical Society.

changes are diametrically opposed in J- versus H-aggregates, allowing for more effective ways with
which to distinguish the two aggregate types beyond the usual spectral shifts. [The latter can be
misleading as there can be several sources of spectral shifting; for example, enhanced intrachain
planarization in poly(3-hexylthiophene) (P3HT) aggregates leads to a significant red shift (37),
despite their H-aggregate nature.] As outlined in Reference 35, in J- (H-) aggregates, the ratio of
the oscillator strengths of the first two vibronic peaks in the absorption spectrum, I A0-0 /I A0-1 , in-
creases (decreases) with exciton bandwidth. Furthermore, in the PL spectrum, the analogous ratio,
0-0 0- 1
IPL /IPL , decreases in J-aggregates (and increases in H-aggregates) with increasing temperature
or increasing disorder. Because increasing disorder and temperature also localize excitations, the
PL ratio in J-aggregates probes directly the exciton coherence length, Lcoh . The scaling is linear,
0- 0 0-1
with IPL /IPL ∝ Lcoh , as shown in Reference 36. In marked contrast, in H-aggregates, the PL
P3HT: poly(3- ratio diminishes as Lcoh increases because of the more efficient destructive interference among
hexylthiophene)
the emitting chromophores (27). Hence, unlike the absorption ratio, which is directly responsive
to the exciton bandwidth, the PL ratio is largely sensitive to the exciton coherence (including

480 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

phase). The different origins of the absorption and PL spectral ratios generally invalidate the
mirror-image relationship in both J- and H-aggregates.
A single polymer chain can be viewed as an unconventional J-aggregate consisting of elec-
PDA: polydiacetylene
tronically coupled monomer repeat units. The J-like behavior derives from the direct band-gap
nature of 1D semiconductors, for which it can be shown, using second-order perturbation
theory, that the effective coupling between adjacent monomers originating from electron and
hole transfer is in fact negative (28, 38). Interestingly, such a through-bond mechanism for
J-like behavior is entirely different from the through-space Coulombic coupling operating in
conventional J-aggregates (although Coulombic coupling may also enhance the J-like behavior
in polymer chains owing to the head-to-tail orientation of monomer transition dipoles). The
deep relationship between emissive conjugated polymer chains and J-aggregates is evident in
the work of Schott and coworkers (39–46) on isolated red-phase chains of the polydiacetylene
(PDA) derivatives referred to as 3BCMU and 4BCMU. The micrometer-long straight chains,
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

or quantum wires, are polymerized via ultraviolet (UV) excitation of the monomer (diacetylene)
crystal and, because of steric restrictions, are virtually devoid of torsional defects. Such chains
have been argued to support enormous exciton coherence lengths of the order of 1–10 μm
(42). They can also be prepared at low concentrations, allowing for the analysis of isolated,
practically defect-free conjugated polymer chains. The red phase of such chains behaves just
like linear J-aggregates with respect to absorption and PL (28). In addition to the conventional
red-shifted absorption spectrum and enhanced radiative decay rates expected of J-aggregates, the
red PDA chains also abide by the new set of spectral signatures based on the vibronic coupling
involving the vinyl and acetyl stretching modes at approximately 1,400 cm−1 and 2,000 cm−1 ,
respectively. Most notably, the PL ratio, I P0-L0 /I P0-L1 , involving either side band is very large, scaling
with temperature as T−1/2 , in excellent agreement with that expected in linear J-aggregates
(36).
A π -stack of such J-like polymer chains presents an interesting photophysical dilemma—does
the stack behave like an H-aggregate, as dictated by the side-by-side arrangement of chains, or
like a J-aggregate, as dictated by the intrachain head-to-tail arrangement of repeat units within
a chain? As we have found, the π -stack displays a unique set of hybrid photophysical properties
depending mainly on the competition between intrachain electronic coupling, which favors J-like
behavior, and interchain coupling, which favors H-like behavior. Hence, we refer to a polymer
π -stack as an HJ-aggregate (47). An HJ-dimer is depicted in Figure 1. The delicate interplay
of interactions, as manifest in changes in the absorption and PL lineshapes, is evident in many
luminescent conjugated polymers. For example, spin-cast films of P3HT (with chloroform as the
solvent) behave remarkably like H-aggregates, likely because of the predominance of aggregates
comprising chains with relatively short conjugation lengths, owing to the high level of disorder
produced by spin casting from a low-boiling-point solvent (24–26). Several groups have deter-
mined that the excitonic interaction between two neighboring polymer chains diminishes with
chain length (48–52); hence, short conjugation lengths lead to stronger interchain interactions
and H-aggregate behavior (24–27, 34). In such P3HT films, the 0-0 peak in the steady-state PL
spectrum is substantially attenuated compared with solution, as expected for H-aggregates. Also
consistent with H-aggregation is the increase in the relative 0-0/0-1 vibronic peak ratio in the PL
spectrum with increasing temperature (25, 27) and the decrease of the PL ratio following impulsive
excitation higher into the exciton band (25, 53–55). All these observations can be accounted for
with the H-aggregate model (24–27), which derives directly from Kasha’s model: Each polymer
(or polymer segment) in the π -stack is treated as a single chromophore coupled via Coulombic
interactions to neighboring chromophores within the stack. The model is 1D, as it ignores exciton
motion along the polymer chain.

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 481


PC65CH21-Silva ARI 10 March 2014 15:47

In a rather dramatic recent development, P3HT nanofibers formed by slowly cooling the
polymer in toluene more strongly resembled J-aggregates, with dominant 0-0 PL and absorp-
tion peaks, presumably because of increased intrachain order, which leads to stronger intrachain
PPV: poly-p-
phenylenevinylene interactions, larger conjugation lengths, and therefore weaker interchain interactions (56). For
such architectures, the H-aggregate model breaks down, as the intrachain coupling and exciton
MEH-PPV:
poly[2-methoxy-5-(2- motion along the polymer chain can no longer be ignored. J- and H-type behaviors have also
ethylhexyloxy)-1,4- been observed in P3HT nanofibers prepared from mixed solvents (119). Very recently, Barnes
phenylenevinylene] and coworkers (120) showed that cross-linking in P3HT copolymers can be used to control the
interchain interactions. Hence, it appears possible to tune between J- and H-dominated behaviors
by controlling microstructure in P3HT.
In poly-p-phenylenevinylene (PPV) and poly[2-methoxy-5-(2-ethylhexyloxy)-1,4-pheny-
lenevinylene] (MEH-PPV) films, the 0-0/0-1 PL intensity ratio decreases with increasing temper-
ature (57–61) and disorder (57) in direct opposition to the increase measured in P3HT spin-cast
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

films and consistent with J-aggregation. Hagler et al. (62) observed some time ago that the vibronic
origin in the absorption spectrum of oriented MEH-PPV films is substantially enhanced as tem-
perature decreases (or orientation increases), unlike what happens in P3HT films spun cast from
volatile solvents, in which only a slight red shift is observed (27). In addition, the PL polarization
anisotropy of PPV films peaks in the vicinity of the 0-0 transition (63–66), an observation that is
also consistent with J-aggregate behavior. All the J-like behaviors listed above can be attributed
to improved intrachain ordering and the resultant increase in intrachain coupling.
Very recently, researchers demonstrated that MEH-PPV dissolved in methyltetrahydrofuran
undergoes a phase transition from a disordered blue phase to an aggregated red phase below a
critical temperature of approximately 200 K (67). Such red-phase aggregates spectrally resemble
J-aggregates with dominant 0-0 peaks in the PL and absorption spectra. In fact, the PL lineshape
is quite similar to the red single-chromophore emitters observed in single-molecule spectroscopy
(68, 69). J-dominant effects may also be observed in polyfluorenes, where the β-phase is charac-
0−0
terized by a substantial increase in IPL /IPL
0−1
as the temperature is lowered (70, 71). The ratio also
increases when comparing (at 10 K) the disordered phase of polyfluorene with β-phase polyfluo-
rene (71).
To account for the H- and J-like photophysical properties exhibited by many emissive conju-
gated polymers, we developed the HJ-aggregate model (47), which accounts for both inter- and
intramolecular degrees of freedom, incorporating the H-like behavior induced by interchain cou-
pling and the J-like behavior induced by intrachain coupling. Such a model can potentially reconcile
the diverse range of photophysical behaviors exhibited by thiophene-, phenylene vinylene-, and
fluorene-based conjugated polymers. The inter- versus intrachain competition is quite sensitive
to the nature and magnitude of disorder and can lead to exotic photophysics, such as disorder-
induced crossover from H- to J-aggregate behavior. Moreover, one can generally deduce the 2D
extent of the exciton coherence in π -stacks from the structure of the PL spectral lineshape. Exciton
coherence is strongly dependent on morphology (72–76) and has recently been analyzed in detail
in MEH-PPV solutions and nanoparticles using ultrafast polarization decay (73). Recent work on
P3HT suggests that exciton coherence in P3HT π -stacks is spread approximately isotropically
along the polymer backbone and along the stacking axis in a manner that depends on molecular
weight (76).
Below we review the experimental evidence for H- and J-aggregate behavior in several emis-
sive conjugated polymers, beginning with P3HT spin-cast films under the H-aggregate model in
Section 2, followed by the seminal work by Schott and coworkers on single PDA chains described
within the J-aggregate model in Section 3. We next consider P3HT whiskers and MEH-PPV
red-phase aggregates within the HJ-aggregate model in Section 4. Interestingly, the PPV-based

482 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

polymers are more J-like than is P3HT, suggesting an increased intrachain bandwidth or a de-
creased interchain bandwidth. Both J- and H-aggregate influences are readily apparent in the
2D exciton coherence function considered in Section 5. Section 6 summarizes our findings and
outlines important future directions.

2. H-AGGREGATES IN P3HT FILMS


The nature of excitons in regioregular P3HT was the subject of some debate over the first part
of the past decade, sparked by the report by Brown et al. (37) that the absorption and PL spectra
of solution-cast P3HT films cannot be described by a simple Franck-Condon progression. These
authors proposed two quasi-resonant emissive states, an intra- and an interchain state, both con-
tributing to the absorption and emission spectra. Kobayashi et al. (77) proposed two intrachain
emissive states, one with allowed transition to the ground state and one generally forbidden but
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

weakly allowed owing to vibronic coupling. Others proposed that the first excited state is a single
interchain H-aggregate state (78, 79). Spano (24) developed that idea with a theoretical model of
weakly interacting H-aggregates. In this section, we discuss how this single-emitter, H-aggregate
model describes much of the dominant photophysics of P3HT films when cast from common
organic solvents. The photophysics of more complex microstructures obtained by more involved
processing conditions is discussed in Section 4.
The H-aggregate model is 1D, accounting for interchain (Frenkel) exciton delocalization be-
tween the polymer chromophores within π -stacks (see the inset in Figure 2a) but neglecting ex-
citon motion along each polymer chain. Vibronic coupling involving a symmetric ring-stretching
mode with energy ω0 = 180 meV is included in a Holstein-like Hamiltonian (80) to which
correlated site disorder is also added.
In the weak coupling limit, the model successfully describes the essential features of the spectral
lineshapes in P3HT films. Figure 2a displays the absorption spectrum of a P3HT film spun from
chloroform (27). The film absorption spectrum comprises two parts: a lower-energy, dominant
part from regions of the film that form weakly interacting H-aggregate states and a higher-energy
part due to more disordered chains that form intrachain states (25). Owing to interband mixing,
the ratio of line strengths of the 0-0 and 0-1 vibronic lines (labeled A1 and A2 in Figure 2a)
is related to the nearest-neighbor interchain Coulombic coupling J0 (>0) via the free-exciton
bandwidth of the aggregates, W = 4J0 , and the energy of the main intramolecular vibration, ω0 ,
coupled to the electronic transition. Assuming a Huang-Rhys factor of unity (25), one finds that
the ratio is given by (24, 25, 34)
 
I A0-0 1 − 0.24W /ω0 2
≈ . (1)
I A0-1 1 + 0.073W /ω0
The ratio decreases with increasing W and has been shown to display a marked dependence on the
solid-state microstructure and average molecular conformation in P3HT films cast from different
solvents (26). Assuming similar interchain order in the films, one finds that W is related to the
conjugation length. An increase in conjugation length and order will lead to a decrease in W (49–52,
81–83). The latter is not intuitive and follows from numerous studies on cofacial oligomer dimers,
showing that once the chain length L approximately exceeds the intermolecular separation, the
interchain Coulombic coupling decreases with L (49–52). In the limit of perfectly ordered polymer
chains, the interchain coupling practically vanishes. The absorbance spectrum is thus sensitive to
interchain excitonic coupling, which in turn is governed largely by intrachain order.
The PL spectral lineshape also reflects the H-aggregate nature of the photophysics in P3HT
films. In an H-aggregate at low temperature, emission occurs from the lowest excited state

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 483


PC65CH21-Silva ARI 10 March 2014 15:47

DISORDER PARAMETERS IN H-AGGREGATE MODEL

The disorder parameters are defined through the correlation function i  j  = σ 2 exp(−|i − j |/l 0 ), where i is
the transition energy offset for the i-th chromophore. Here, σ 2 is the variance of the Gaussian inhomogeneous
distribution, and l0 is in units of d, where d is the nearest-neighbor separation between chains.

according to Kasha’s rule (see Figure 1 and the inset in Figure 2a) with zero oscillator strength for
the 0-0 transition in the absence of disorder. In contrast, side-band transitions are allowed whether
disorder is present or not. In the weak coupling limit, the intensities of the 0-1 and subsequent
replicas are somewhat diminished from their single-molecule values (24, 26, 27, 84), consistent
with the aggregation-induced reduction in quantum yield. However, the side bands retain the
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

same relative intensities as the isolated molecule spectrum (24, 27). Hence, the aggregate emission
spectrum shows a weak, disorder-allowed 0-0 component followed by a Franck-Condon progres-
sion. Accordingly, the PL spectrum can be modeled as a modified Franck-Condon progression
with a variable 0-0 amplitude, which is decoupled from the rest of the progression (25).
Figure 2a also shows the PL spectra at 10 K and room temperature. The 0-0 line is weakly
allowed owing to the presence of disorder. Upon increasing temperature, the PL spectra blue-shift
and broaden because of increasing thermal disorder, accompanied by a systematic relative increase
of the 0-0 peak intensity due primarily to thermal excitation to states with greater oscillator strength
in the lowest vibronic band. Thermal disorder may also relax suppression of the 0-0 intensity. The
weakly coupled H-aggregate model predicts a spectral blue shift and 0-0 peak enhancement with
temperature (25, 27), in good qualitative agreement with experiment. Furthermore, a dynamic red
shift and the loss of 0-0 peak intensity with time indicate energy diffusion to more ordered domains,
which carry lower 0-0 intensity (25, 27). From combined PL measurements and modeling, we
concluded that H-aggregates are the only emissive species in P3HT (25).
Although both absorption and PL spectra reflect the electronic structure predicted by an
H-aggregate model in the weak coupling limit, the PL spectrum reflects the quality of the dis-
ordered energy landscape and reveals insight on the exciton coherence length—the spatial extent
of the vibronic excitation along the π -stacking direction. In particular, it is sensitive to correlated
disorder, characterized by the width of the Gaussian distribution of site energies σ , and a spatial
correlation length l0 (see the sidebar, Disorder Parameters in H-Aggregate Model) (85).
The influence of l0 on the PL lineshape is profound (Figure 3). When intra-aggregate disorder
is absent entirely (l 0 = ∞), the 0-0 peak vanishes rigorously, as expected in H-aggregates, in which
the wave function of the band-bottom exciton changes phase between neighboring chromophores,
making the transition to the vibrationless ground state optically forbidden. In Figure 3, the 0-0
peak is already effectively absent when l0 = 100. Introducing disorder via a reduction in l0 breaks
the symmetry, thereby allowing 0-0 emission. The relative 0-0 intensity increases monotonically
with disorder. Figure 3 also shows that the prominent 0-1 peak substantially red-shifts and narrows
relative to the 0-1 peak in more ordered aggregates.
From the perturbative expression derived in Reference 27, the ratio of the 0-0 and 0-1 PL line
strengths scales as

I 0-0 (1 − β)σ 2
- ∼ (1 − β)σ W ω0 , (2)
I 0 1 (1 + β)W 2
where β ≡ exp(−1/l 0 ). Equation 2 is consistent with the vanishing of the 0-0 line strength
in the disorder-free limit (l 0 = ∞ or σ = 0), as well as the increase in the 0-0 intensity

484 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

0.4 2
0 l0 = 0
l0 = 0 0-1 0.2 –2
l0 = 3.5

C(r)
l0 = 10 0.0
l0 = 100
Normalized emission intensity

0-0
–0.2
Lcoh = 0.57
–0.4

0.4 2
0 l0 = 3.5
0.2 –2

C(r)
0.0
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

–0.2
Lcoh = 1.55
–0.4

2
0.4 0 l0 = 10
–2
0.2
1.2 1.4 1.6 1.8 2.0
C(r)
Photon energy (eV) 0.0

–0.2
Lcoh = 2.35
–0.4

2
0.4 0 l0 = 100
–2
0.2
C(r)

0.0

–0.2
Lcoh = 4.66
–0.4
–10 –5 0 5 10
r

Figure 3
(a) Calculated photoluminescence spectra based on the H-aggregate model for P3HT π -stacks containing N = 100 chromophores
normalized to the 0-1 intensity. Spectra are shown for several values of l0 with σ = 0.08 eV after averaging over 104 disorder
configurations. (b) The corresponding configuration-averaged coherence function. (Insets) Representative disorder distributions. Figure
adapted with permission from Reference 27. Copyright 2009, AIP Publishing LLC.

with increasing disorder (decreasing l0 ). When intra-aggregate disorder is greatest (l0 = β =


0), the 0-0 peak is maximized for a given σ . In Figure 3, the 0-0 intensity is nearly equal to
the 0-1 peak intensity when l0 = 0 because we have chosen the Huang-Rhys parameter to be
unity. In this case, an isolated molecule would exhibit identical 0-0 and 0-1 line strengths. The
single-molecule-like behavior of the l0 = 0 spectrum in Figure 3 results from an enhanced like-
lihood that the offset difference between neighboring molecules is much larger than the excitonic
coupling, J0 . In marked contrast, the relative intensities of the side bands are practically unchanged
with disorder. Even when disorder is absent, there is no selection rule requiring the side-band
transitions to be zero, as exists for the 0-0 transition, because the electronic ground state can
conserve momentum in the presence of one or more phonons of the appropriate wave vector

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 485


PC65CH21-Silva ARI 10 March 2014 15:47

(see Figure 1). Within an H-aggregate model in the weak coupling limit, the relative side-band
intensities conform to what is expected for an isolated chain.
Figure 3b shows the configurationally averaged coherence function for π -stacks with 100
chains. For a thermal distribution of low-energy excitons responsible for emission and transport,
all information regarding exciton coherence is contained in the ensemble-averaged coherence
function, given by
 
 
 
C̄(r) ≡  (em)  BR† BR+r  (em) , (3)
R C,T

where BR†
≡ |R; vac  g; vac| creates an exciton at site R with no vibrational quanta (vac) relative
to the ground-state unshifted potential well (27, 86). Here, the vector R locates the position of a
given chain in the stack. ...C,T represents a dual configurational and thermal average, the former
taking place over the various realizations of site-energy disorder and the latter taking place over a
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

Boltzmann distribution of emitting excitons. The coherence function in Equation 3 is similar to


that used by Mukamel and coworkers (87–89) and Kuhn & Sundström (90).
As the H-aggregate model is 1D, we considered only the coherence along the π -stacking
direction in Reference 27, showing that the coherence function oscillates along the aggregate axis
in linear H-aggregates, changing sign as (−1)n , where n identifies the n-th chain in the stack.
The oscillation reflects the dominant admixture of the high-wave-vector (k = π ) exciton in the
band-bottom excitons, as characteristic of disordered H-aggregates (27). By contrast, in linear
J-aggregates, the coherence function is nodeless, reflecting the dominant admixture of the k =
0 exciton in the band-bottom excitons (36) (see Section 4). In Reference 27, we reproduced the
experimental temperature-dependent PL spectra shown in Figure 2a with l0 = 3.5 (β = 0.75).
With this value and a Huang-Rhys parameter of 1, the calculated emission spectra agreed with
the measured spectra throughout the entire range of temperatures, from 10 to 300 K.
Based on the spatial extent of the coherence function in Equation 3, one can approximate the
total number of coherently connected chains, Ncoh :
  
N coh = C̄(0)−1 C̄(r), (4)
r

where the dimensionless vector r runs over all site-separation vectors within the π -stack (27, 36).
The absolute value dependence on C(r) eliminates the phase oscillations, as Ncoh is determined by
the envelope of the coherence function. We note that in the limit of strong delocalization where
C̄(r) = C̄(0)δr,0 , Ncoh properly tends toward 1. The coherence length, Lcoh , is simply N coh − 1 in
units of interchain spacing (∼4 Å) and is shown in the insets of Figure 3.
In P3HT films cast from common organic solvents, linear absorption and steady-state PL
spectra are well described by the H-aggregate model developed in References 24, 25, 27, and 34.
This is the case in which interchain excitonic coupling dominates the photophysics, which appears
to be generally the case in spin-cast P3HT films. Nevertheless, there are important examples in
which the pure H-aggregate model breaks down, which we outline in Section 4. Before doing so,
we introduce the most compelling example of pure J-like photophysical behavior, red-phase PDA.

3. SINGLE-CHAIN PDA: A J-AGGREGATE?


Schott and coworkers (40–46) have conducted extensive studies on the PDA derivatives referred
to as 3BCMU and 4BCMU, which are prepared by UV polymerization of the monomer (diacety-
lene) crystal, forming dilute distributions of essentially defect-free, straight polymers or quantum
wires. The red phase of such wires is strongly emissive, with the low-temperature PL spectrum
shown in Figure 4 characterized by a dominant 0-0 peak at 2.28 eV and much smaller vibronic

486 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

1,400

1,200 0-0

1,000

ID/I 0-0
Intensity (a.u.)

800
0-1
600
T D
0
400 0 2 4 6 8
T 1/2 (K 1/2) ×60
200

0
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

1.6 1.7 1.8 1.9 2.0 2.1 2.2 2.3


E (eV)
Figure 4
Photoluminescence spectrum for red-phase 3BCMU PDA at T = 15 K. D and T represent the double-bond
and triple-bond stretching modes, respectively. (Inset) The ratio of the side-band 0-1 (D) and 0-0 spectral
areas, I D /I 0-0 , as a function of T1/2 . Hence, the 0-0/0-1 ratio scales as the inverse square root of temperature.
Figure reprinted with permission from Reference 44. Copyright 2002 by the American Physical Society.

side bands due to the double-bond and triple-bond stretching modes approximately 0.18 eV and
0.25 eV below the 0-0 peak, respectively (44). The 0-0 peak is shifted by approximately 1.3 eV
to the red of the corresponding peak in small oligomers containing only three repeat units (91),
consistent with the usual narrowing of the optical gap as a function of chain length exhibited by
conjugated polymers. The vibronic lines in Figure 4 are only 20–30 cm−1 wide, almost two orders
of magnitude narrower than the corresponding spectral lines measured in P3HT and MEH-PPV
films at low temperatures, reflecting the near-total lack of disorder. The dominance of the PL
0-0 peak, which dwarfs the much smaller side bands by about a factor of 100 at T = 15 K, speaks
to an exceptionally large exciton coherence length (see below), consistent with an associated re-
duction in vibronic coupling. The 0-0 peak in the absorption spectrum is also dominant (not
shown), but is only a factor of ten or so larger than a vibronic side band (46). Hence, there is no
mirror-image symmetry between the absorption and PL lineshapes, which is often expected in
single-chromophore systems. Schott and coworkers went on to show that the ratio of the 0-0 to
0-1 intensities (for either the double- or triple-bond stretching modes) in the PL spectrum scales
as the inverse square root of temperature (see the inset of Figure 4) and that the radiative decay
rate scales similarly with temperature (not shown), as characteristic of 1D semiconductors (92).
Using an elaborate microfluorescence detection apparatus, Dubin et al. (42) argued that within a
single PDA wire, exciton coherence extends over a range of 1–10 μm, a truly amazing distance
that is orders of magnitude larger than typical exciton diffusion lengths measured in films of the
more common emissive polymers such as P3HT.
All the behaviors recounted above are consistent with the fundamental photophysical properties
of ordered linear J-aggregates as demonstrated in Reference 28. These include the conventional
signatures (red-shifted absorption, enhanced radiative decay rates) as well as the vibronic sig-
natures. The J-like behavior of PDA chains arises from an intimate connection between linear
J-aggregates and 1D direct band-gap semiconductors; in both systems, the exciton band exhibits
positive curvature at k = 0 (see Figure 1), leading, for example, to the T −1/2 scaling of the 0-
0/0-1 PL ratio (28, 36). Based on a 1D Wannier exciton Hamiltonian that includes local vibronic

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 487


PC65CH21-Silva ARI 10 March 2014 15:47

coupling à la Holstein (80), the ratio of the 0-0 and 0-1 line strengths in the PL spectrum was
shown to obey

0- 0 0-1 κ 4π ωc
IPL /IPL ≈ 2 , (5)
λ0 kb T
where ωc is the band curvature, ωc ≡ (1/2)d 2 E/d k2 , evaluated at the bottom of the exciton
band where the wave vector is k = 0 (28) (see Figure 1). Here, k is taken to be dimensionless,
ranging from −π to π . λ20 in Equation 5 is the Huang-Rhys factor for a given repeat unit, and κ is a
dimensionless factor that deviates from unity as the electron and hole separate to form a Wannier-
like exciton. The calculations in Reference 28 for red-phase PDA showed that ωc ≈ 0.5 e V/ for
the 11 Bu exciton with κ ≈ 0.84, reflecting an exciton with a Bohr radius of approximately 12 Å
and a binding energy of 0.8 eV, in good agreement with the estimates of Horvath et al. (43). The
expression in Equation 5 assumes a parabolic exciton band approximation as well as sufficiently
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

large values of N to ensure the thermodynamic limit, N > 4π ωc /kb T . Equation 5 is also useful
in that it allows an independent means of measuring the band curvature and hence the exciton’s
effective mass directly from the temperature dependence of the PL ratio (28).
When κ = 1, Equation 5 reduces exactly to the expression derived for a linear J-aggregate
(36). In this limit, the electron and hole are bound together within a given repeat unit, thereby
resembling Frenkel excitons in conventional J-aggregates. The demise of the PL ratio for either
1D semiconductors or linear J-aggregates with increasing temperature reflects a diminishing co-
herence size, localized by thermal fluctuations. For the parameters that best describe PDA chains,
we obtained a coherence length of approximately 50 nm at T = 15 K (28).
Similar to the PL spectrum, the absorption spectrum of a PDA chain is also dominated by the
0-0 peak (28), which is approximately 5 to 10 times more intense than the first side band (0-1)
in red and blue phases (46). By contrast, in small oligomers, the 0-0 and 0-1 peaks are similarly
intense (91). These observations are consistent with calculations based on the 1D semiconductor
model, in which the ratio of the 0-0 to 0-1 oscillator strengths increases with the number of repeat
units, converging to a value of approximately five in the polymer limit (28). The ratio also increases
with the exciton bandwidth, consistent with the J-aggregate version of Equation 1, obtained by
replacing W with −W to reflect the sign change in the excitonic coupling (35). The absorption
ratio is more stable with temperature as it mainly reflects the exciton bandwidth, unlike the PL
ratio, which mainly reflects the exciton coherence size. The decrease of the radiative decay rate of
the red-phase PDA with increasing temperature can also be viewed as superradiance, which is a
well-known signature of molecular J-aggregates (93, 94). In superradiance, the radiative decay rate
scales as the number of coherently connected emitters. Such an effect has also been observed in the
phosphorescence of platinum-containing conjugated polymers (95). The theory and calculations
from Reference 28 show that the number of coherently coupled monomers within the PDA chain
scales as the inverse square root of temperature, thereby reproducing the temperature dependence
of the radiative decay rate. Hence, essentially all the photophysical properties measured by Schott
and coworkers on red-phase PDA mimic those of linear J-aggregates. This is ultimately due to the
fact that, like conventional J-aggregates, the k = 0 exciton in direct band-gap semiconductors lies
at the bottom of the exciton band (see Figure 1). The direct band-gap nature of ideal conjugated
polymer chains is already evident at the level of simple Huckel theory, at which the valence and
conduction bands in dimerized polyenes have opposite curvature. (Interestingly, indirect band-gap
π -systems would behave as linear H-aggregates and would be weakly emissive.) The mechanism
for J-like behavior in 1D semiconductors is therefore established at the noninteracting electron
level through electron and hole transfer, as opposed to conventional J-aggregates in which the
mechanism relies on Coulombic coupling. The J-aggregate analogy in conjugated polymers does,

488 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

however, require that the optically allowed 1Bu exciton lies energetically below the 2Ag state.
Accordingly, J-like behavior does not occur in nonemissive blue-phase PDA chains, in which
electron correlation effects likely place the 2Ag exciton below the emissive 1Bu exciton. (With this
condition, polyacetylene, for example, would also not behave like a J-aggregate.)

4. HJ-AGGREGATES IN P3HT AND MEH-PPV FILMS


In a polymer π -stack, typical of the packing present in P3HT films, there exists competition
between the interchain (H-favoring) interactions considered in the H-aggregate model and intra-
chain ( J-favoring) interactions. As discussed in Section 2, the H-aggregate model well describes the
photophysical properties of P3HT π -stacks in spin-cast films. However, in a recent investigation,
P3HT whiskers formed by slowly cooling a P3HT/toluene solution showed dominant J-aggregate
Access provided by National Cheng Kung University on 12/18/23. For personal use only.

photophysical behavior (56). As shown in Figure 2c, the absorption and PL spectra of such whiskers
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

are dominated by the 0-0 vibronic component, unlike the case for the spin-cast H-aggregates
shown in Figure 2a. The ability of one polymer to assume both H- and J-aggregate forms most
likely results from differing morphologies (see Figure 2b): In whiskers, superior ordering along
the chains promotes stronger intrachain interactions and weaker interchain interactions (see the
discussion following Equation 1). The mostly H-like behavior occurring in P3HT films cast from
the lowest-boiling-point solvents, such as chloroform, arises from greater disorder from rapid
solvent evaporation and hence shorter conjugation lengths (and greater interchain interactions).
The HJ-aggregate model outlined in detail in Reference 47 considers excitons delocalized both
along and across polymer chains within a π -stack and is therefore able to unravel the competitive
effects of intrachain ( J-favoring) versus interchain (H-favoring) interactions and their impact on
the photophysical response. In Reference 47, a pair of cofacial polymer chains was considered with
Coulombic interactions between adjacent repeat units on neighboring chains. When disorder is
ignored, as may occur between two PDA chains prepared in situ from the monomer crystal, the
interchain interaction leads to a symmetric and antisymmetric version of each intrachain exciton
of wave vector k, with a splitting, E, independent of k, as depicted in Figure 1. Here the
symmetry refers to a reflection plane bisecting the dimer pair. Interestingly, only the k = 0,
symmetric exciton, which is of higher energy, can provide 0-0 emission; in other words, it is the
only state that couples radiatively to the vibrationless ground state. Hence, 0-0 emission must be
thermally activated, as in an H-aggregate. Under the parabolic band approximation and within the
thermodynamic limit, the complete temperature dependence of the PL ratio takes the form (47)

0-0 0-1 κ 2e −E/kb T 4π ωc


IPL /IPL ≈ 2 . (6)
λ0 1 + e −E/kb T kb T

When the splitting vanishes, as in two noninteracting chains, Equation 6 reduces to the single-
chain result in Equation 5. As demonstrated in Reference 47, when kb T is approximately E,
the PL ratio peaks and thereafter decreases with increasing temperature, just like a J-aggregate.

The maximum PL ratio scales as ωc /E, directly demonstrating the competitive influences
of intrachain (ωc ) and interchain (E) interactions. The overall temperature dependence shows
that, with respect to the PL, the dimer displays an H-to-J transition upon increasing temperature.
The dependence of the PL ratio as a function of disorder has a similar form (N.J. Hestand,
H. Yamagata & F.C. Spano, unpublished data). For example, for diagonal energetic disorder,
in which each repeat unit has a randomly chosen excitation energy taken from a Gaussian
distribution of width, σ , one can show that at low temperatures (kb T E), the PL ratio initially
increases with σ , like an H-aggregate; peaks when σ is of the order of the splitting E; and

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 489


PC65CH21-Silva ARI 10 March 2014 15:47

0-0

Standard

0-1
Improved
300 K

15 K
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

1.8 2.0 2.2 2.4 2.6


Energy (eV)
Figure 5
Photoluminescence spectra for standard and improved (96) PPV films at 15 K (solid line) and 300 K (dotted line).
For both films, a decrease in temperature results in a substantial increase in the integrated 0-0/0-1 line strength
ratio, I0-0 /I0-1 . A similar increase is observed when the intrachain order is increased in going from the standard
to improved samples. These are clear signatures of J-aggregate behavior owing to dominant intrachain
interactions. Figure reproduced with permission from Reference 57,  c IOP Publishing. All rights reserved.

then decreases with further increases in σ , as is characteristic of a J-aggregate. (This assumes a


constant interchain interaction throughout.)
Figure 5 shows the temperature-dependent PL spectra of standard and improved PPV (57), the
latter prepared utilizing a synthetic route that limits disorder as compared to the former (96). For
both temperatures reported, the improved PL spectrum demonstrates significant enhancement of
the PL vibronic ratio, a J-like response most likely resulting from the enhanced intrachain order
in the improved samples. The 0-0 peak in the absorption spectrum (not shown) is also enhanced
relative to the first side band in the improved samples. The temperature dependence of the PL
ratio is also J-like; in either sample, increasing temperature leads to a significant drop in the PL
ratio. It would be interesting to demonstrate the existence of an HJ transition in such samples by
observing the complete temperature dependence of the PL ratio.
Aggregate effects in MEH-PPV have been investigated rather intensively over the past several
years (33, 97–103). Recently, Köhler et al. (67) showed that MEH-PPV in methyltetrahydrofuran
undergoes a phase transition from a disordered blue phase (Figure 6a) to an ordered red phase
(Figure 6b) at approximately 200 K. In addition to the large red shift and line narrowing
representative of the red phase at 110 K, there is also a pronounced increase in the 0-0/0-1
PL ratio. Such attributes indicate J-like behavior brought on by the chain elongation expected
upon aggregation. Chain elongation enhances the alignment of the repeat unit transition
dipoles, thereby increasing the PL 0-0/0-1 ratio relative to the blue phase. However, there is
also substantial disorder present as well as destructive interference between adjacent chains (or
segments) within the aggregate, two effects that limit the PL ratio. In applying the HJ-aggregate
model, it was recently found that the overall properties of the aggregate are nearly the same as a
single chain (104), consistent with the earlier findings of Collison et al. (99), making MEH-PPV
aggregates similar to disordered J-aggregates. The interchain interactions serve to reduce the PL
ratio by approximately 30% with a much smaller impact on the absorption spectral lineshape.
Moreover, the red-phase PL spectrum in Figure 6b closely resembles the spectrum of the

490 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

1.0
a 290 K

0.8
PL

PL/abs (a.u.)
0.6
Ab

0.4

0.2

0.0
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

1.0
b 110 K

0.8
PL/abs (a.u.)

0.6 Ab

PL
0.4

0.2

0.0
2.0 2.5 3.0 3.5
Energy (eV)
Figure 6
(a) The blue-phase absorption and photoluminescence (PL) spectra of MEH-PPV in methyltetrahydrofuran
at a temperature above the phase transition temperature (205 K). (b) The spectra of red-phase MEH-PPV
obtained below the phase transition temperature. Figure adapted with permission from Reference 67.
Copyright 2012 American Chemical Society.

red single-chromophore emitter identified in single-molecule spectroscopy (68, 69). Hence,


the photophysical response of the red-phase MEH-PPV aggregates is dictated by dominant
intrachain coupling, in marked contrast to the H-aggregates in P3HT films discussed in Section
2. Similar intrachain dominant behavior is found in polyfluorenes (105), in which the so-called
β-phase displays a dominant 0-0 peak in absorption and emission (70, 71).

5. EXCITON COHERENCE IN TWO DIMENSIONS


In polymer π -stacks, exciton coherence is spread over two main dimensions: along the polymer
backbone and across polymer chains within the stack. In this section, we discuss the exciton
coherence function in two dimensions and show that it displays characteristics of both J- and
H-type coupling. The relationship between the coherence function and the shape of the PL
spectrum is also considered.
Figure 7 shows an example of a 2D coherence function. The coherence function corresponds
to films of low-molecular-weight P3HT (12 kg/mol) as detailed in Reference 76 and is evaluated
from Equation 3, where R now identifies the location of a given thiophene ring within the stack.

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 491


PC65CH21-Silva ARI 10 March 2014 15:47

0.3

0.25

0.2

0.15

0.1

0.05

–0.05
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

–0.1

–0.15

–0.2

20
10 15
10
0 5
Separ –5 0
ation a –10
lo –10 ain (Å)
ng cha –20 –15
n across ch
in (Å) Separatio

Figure 7
2D exciton coherence function calculated using the HJ-aggregate model under an effective Frenkel exciton
approximation as outlined in Reference 76. Figure reprinted with permission from Reference 76. Copyright
2013 by the American Physical Society.

The emitting exciton wave functions were obtained from an effective Frenkel exciton Hamiltonian
with the disorder correlation parameter set to β = 0.6. The remaining parameters defining the
Hamiltonian were chosen to reproduce the PL spectral lineshape as well as the absorption spectral
lineshape of the low-molecular-weight P3HT films (not shown). Despite both spectra showing
strong H-like characteristics, similar to the spectra in Figure 2a, the coherence function corre-
sponding to the band-bottom (emitting) exciton exhibits properties of both H- and J-aggregation:
The oscillations along the π -stacking direction result from the positive sign of the interchain
coupling and signal H-aggregation, as discussed in Section 2, whereas the uniform phase of C̄(r)
along the chain direction results from the negative sign of the intrachain interactions and signals
J-aggregation. Based on the coherence function in Figure 7, one can also approximate the total
number of thiophene rings, Ncoh , within the coherence area defined by the spatial extent of the
envelope of the coherence function. Using Equation 4, one finds that the value of Ncoh is approx-
imately 16 thiophene rings. Because the temperature is low (T = 10 K), static disorder is mainly
responsible for localizing the exciton.
Within the π -stack, one can also define the coherence lengths along the polymer chain direction
(L|| ) and along the π -stacking axis (L⊥ ) from the coherence function via


L|| ≡ C̄(0)−1 |C̄(r)| − 1, (7a)
r∈chain

492 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

⎧ ⎫

⎨  ⎪

L⊥ ≡ C̄(0)−1 |C̄(r)| − 1. (7b)

⎩ ⎪

r∈chain
normal

The dimensionless coherence length L|| (L⊥ ) is reported in units of d (d ⊥ ), the nearest-neighbor
distance between adjacent thiophene rings (polymer chains). Both d and d ⊥ are approximately
4.0 Å in P3HT stacks. From the coherence function in Figure 7, the use of Equation 7a,b gives
L|| ≈ 3.3 thiophene rings and L⊥ ≈ 2.6 polymer chains. The coherence length between chains is
only slightly smaller than the value of ≈ 3 obtained using the H-aggregate model in Reference
27. A more detailed investigation of the coherence lengths as a function of molecular weights is
reported in Reference 76.
The 2D coherence function directly impacts the PL spectrum, mainly through the ratio
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

0-0 0-1
IPL /IPL , which is driven mainly by the uniquely coherent nature of the 0-0 peak within the
PL vibronic progression (26, 35, 36). For any aggregate type, the dimensionless 0-0 line strength
is directly related to the coherence function through

0-0
IPL = C̄(r). (8)
r

In the case of linear J-aggregates, where C̄(r) is uniformly positive, the 0-0 peak is a direct measure
of the coherence size and is the source of superradiance [see Equation 4 with |C̄(r)| = C̄(r)].
0-0 0-1
Because the side-band line strengths are largely incoherent, the ratio IPL /IPL becomes a useful
probe of coherence. In Reference 36, we obtained the simple relationship

0- 0 0-1
IPL /IPL ≈ N coh /λ2 . (9)

Equation 9 provides a simple means of extracting Ncoh directly from the PL spectrum in linear
J-aggregates (28, 36).
In marked contrast, in H-aggregates, the 0-0 intensity is not directly proportional to the co-
0-0
herence number because the phase oscillations in C(r) lead to destructive interferences in IPL (see
Equation 8) but not in Ncoh (see Equation 4). As a result, the PL ratio decreases with increasing
Ncoh as the destructive interference between chains becomes more effective (27). For a fully co-
herent exciton with Ncoh = N, the PL ratio vanishes because the 0-0 peak is symmetry forbidden
in ordered H-aggregates at T = 0 K. Hence, there is no simple relationship relating the PL ratio
to Ncoh in H-aggregates. Expressions such as Equation 3 show that Ncoh is a complex function of
the exciton bandwidth, the nature of the disorder, and the vibronic coupling. However, once a
model for disorder is assumed, one can determine Ncoh numerically from the measured PL ratio,
as was done in Reference 27.
In the π -stacks of interest here, the PL ratio is enhanced by the coherence along the polymer
chain, as the transition dipoles of the repeat units are aligned in phase (see Figure 7). However,
between chains, there is a phase shift, which causes destructive interference between the chains.
0-0 0- 1
Hence, the H-like interchain character leads to attenuation of IPL /IPL . In the presence of dis-
order and thermal fluctuations, the overall PL ratio therefore results from competition between
intrachain, J-favoring interactions and interchain, H-favoring interactions. As shown here, the
competition in P3HT films is also a function of the chain conformation dictated by the mate-
rial’s molecular weight, with the most influential factor being the enhanced intrachain coupling
(and attenuated interchain coupling) experienced by the more planar (torsionally less disordered)
macromolecules of P3HT of higher molecular weight (76).

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 493


PC65CH21-Silva ARI 10 March 2014 15:47

6. SUMMARY AND OUTLOOK


This review provides evidence that an isolated (emissive) polymer chain behaves photophysically
pBTTT: like a J-aggregate, with the most striking example provided by red-phase PDA. In aggregated
poly[2,5-bis(3- conformations, the more common emissive polymers, such as P3HT and MEH-PPV, display
tetradecylthiophen-2- a microstructure-dependent interplay of interchain (H-favoring) and intrachain ( J-favoring)
yl)thieno[3,2-
excitonic coupling, which is observed through linear absorption and steady-state PL spectroscopy
b]thiophene]
and described by effective models based on disordered Holstein Hamiltonians. We consider
F8TBT: poly[(9,9-
that this microstructure-dependent electronic structure is general in polymeric semiconductors.
dioctylfluorenyl-2,7-
diyl)-alt-(4,7-bis(3- Polymers can generally adopt a semicrystalline microstructure, which results in conjugated poly-
hexylthiophen-5-yl)- mer films with significant intra- and interchain electronic dispersion, with profound influence,
2,1,3- for example, on electronic transport properties (106–112). To underscore the generality of the
benzothiadiazole)- existence of H- and J-aggregate signatures in these materials, Hellmann et al. (113) demonstrated
2 ,2 -diyl]
Access provided by National Cheng Kung University on 12/18/23. For personal use only.

that the transition from H- to J-aggregate behavior—as displayed by an increase of the 0-0/0-1
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

MDMO-PPV: absorbance ratio with respect to the isolated molecule value—is not only achieved in P3HT
poly[2-methoxy-5-
but is also observed in films of pBTTT (poly[2,5-bis(3-tetradecylthiophen-2-yl)thieno[3,2-b]
(3 ,7 -
dimethyloctyloxy)-1,4- thiophene]), F8TBT (poly[(9,9-dioctylfluorenyl-2,7-diyl)-alt-(4,7-bis(3-hexylthiophen-5-yl)-
phenylenevinylene] 2,1,3-benzothiadiazole)-2 ,2 - diyl]), MDMO-PPV (poly[2-methoxy-5-(3 ,7 -dimethyloctyloxy)-
PFO: poly(9,9 - 1,4-phenylenevinylene]), and PFO [poly(9,9 -dioctylfluorene)] when blended with polyethyl-
dioctylfluorene) eneoxide (the latter forming the so-called β-phase). This processing is similar to that performed
by Niles et al. (56) in that it induces microstructures that are distinct from those formed in films
processed from organic solvents. Interestingly, Hellman et al. observed that the transition from
predominantly H-like to J-like features in the absorption spectrum in P3HT occurs for polymers
with weight-average molecular weight above ∼30 kg/mol, which begin to display two-phase
microstructures comprising crystalline moieties embedded in amorphous phases in solution-
processed films (114). Therefore, for sufficiently high molecular weights, microstructures that
predominantly produce H-like photophysical aggregates can be controlled by processing to yield
predominantly J-like aggregates. Elucidating the exact character of such microstructures is a
fundamental challenge for the polymer science community.
One important issue arising from the comparison of P3HT and MEH-PPV presented here
pertains to the apparently weaker intrachain bandwidth (or stronger interchain bandwidth) for
P3HT compared to other conjugated polymers (47, 76). Processing in more polar environments
than in commonly used solvents like chloroform can induce signatures of J-like coupling in P3HT
and, as discussed above, only for materials that have sufficiently high molecular weight; when
processing P3HT in weakly polar media, as in common organic solvents, the spectroscopic signa-
tures are decidedly H-like and are stable to thermal treatment. This contrasts with MEH-PPV,
for example, in which a second-order phase transition can be observed by cooling a pristine film
down to 200 K (67). Similarly, the β-phase is induced in polyfluorenes by a variety of processing
protocols but was first observed upon cycling temperature between room temperature and liquid
nitrogen temperatures (105). Why is P3HT apparently different than other polymeric semicon-
ductors in this respect? Spano et al. (115) recently showed that H- and J-aggregates can be created
by the short-range wave-function overlap coupling induced by Frenkel/charge-transfer exciton
mixing in cases in which the conventional Coulombic coupling is much smaller by comparison.
The sign of the coupling depends on the exact registry between neighboring chains. It may be
that in P3HT, the interchain coupling is comparatively higher than in other materials as a result
of this phenomenon, and therefore, the H-like signatures are generally observed, save exceptional
processing conditions. Generally, understanding the differences in photophysical properties
between the different classes of emissive polymers will require more detailed knowledge of the

494 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

relationship between the various realizations of disorder and the inter- and intrachain exciton
bandwidths.
We end by discussing the role of charge-transfer excitons in neat polymeric semiconductors.
Silva and colleagues (116, 117) suggested that regions of higher structural disorder, characteristic
of gradual transitions between lamellar and chain-entangled, amorphous phases in semicrystalline
microstructures, are responsible for exciton dissociation to tightly bound geminate-polaron pairs.
The role of charge-transfer states in H- and J-aggregate coupling in polymeric semiconductors
is an area of research that promises nonincremental insight into the physics of this exciting class
of materials.

SUMMARY POINTS
1. Single chains of emissive conjugated polymers behave photophysically like J-aggregates.
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

2. The photophysics of conjugated polymer π -stacks is determined by competition between


J-favoring intrachain interactions and H-favoring interchain interactions.
3. The polymer P3HT can behave as both an H-type aggregate and a J-type aggregate, de-
pending on the morphology (preparation method). This is most likely a general property
for all emissive conjugated polymers.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
We are indebted to numerous colleagues and collaborators that have provided the basis for the
intellectual impetus for this article: Jenny Clark, John Gray, Anna Köhler, Heinz Bässler, Natalie
Stingelin, Francis Paquin, Hajime Yamagata, Nick Hestand, and the late Gianluca Latini. F.C.S.
acknowledges support from the National Science Foundation under grant DMR-1203811. C.S.
acknowledges support from the Canada Research Chair in Organic Semiconductor Materials, the
Natural Sciences and Engineering Research Council of Canada, the Université de Montréal, and
the Leverhulme Trust.

LITERATURE CITED
1. Heeger AJ. 2010. Semiconducting polymers: the third generation. Chem. Soc. Rev. 39:2354–71
2. Bredas J-L, Norton JE, Cornil J, Coropceanu V. 2009. Molecular understanding of organic solar cells:
the challenges. Acc. Chem. Res. 42:1691–99
3. Kim Y, Cook S, Tuladhar SM, Choulis SA, Nelson J, et al. 2006. A strong regioregularity effect in
self-organizing conjugated polymer films and high-efficiency polythiophene:fullerene solar cells. Nat.
Mater. 5:197–203
4. Kim JY, Lee K, Coates NE, Moses D, Nguyen TQ, et al. 2007. Efficient tandem polymer solar cells
fabricated by all-solution processing. Science 317:222–25
5. Brabec CJ, Dyakonov V, Parisi J, Sariciftci NS, eds. 2003. Organic Photovoltaics: Concepts and Realization.
New York: Springer

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 495


PC65CH21-Silva ARI 10 March 2014 15:47

6. Brabec CJ. 2004. Organic photovoltaics: technology and market. Sol. Energy Mater. Sol. Cells 83:273–92
7. Erb T, Zhokhavets U, Gobsch G, Raleva S, Stuhn B, et al. 2005. Correlation between structural and opti-
cal properties of composite polymer/fullerene films for organic solar cells. Adv. Funct. Mater. 15:1193–96
8. Schilinsky P, Asawapirom U, Scherf U, Biele M, Brabec CJ. 2005. Influence of the molecular weight of
poly(3-hexylthiophene) on the performance of bulk heterojunction solar cells. Chem. Mater. 17:2175–80
9. Yu G, Gao J, Hummelen JC, Wudl F, Heeger AH. 1995. Polymer photovoltaic cells: enhanced efficiencies
via a network of internal donor-acceptor heterojunctions. Science 270:1789–91
10. Granstrom M, Petritsch K, Arias AC, Lux A, Andersson MR, Friend RH. 1998. Laminated fabrication
of polymeric photovoltaic diodes. Nature 395:257–60
11. Blom PWM, Mihailetchi VD, Koster LJA, Markov DE. 2007. Device physics of polymer:fullerene bulk
heterojunction solar cells. Adv. Mater. 19:1551–66
12. Gunes S, Neugebauer H, Sariciftci NS. 2007. Conjugated polymer-based organic solar cells. Chem. Rev.
107:1324–38
13. Yang X, Loos J. 2007. Toward high-performance polymer solar cells: the importance of morphology
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

control. Macromolecules 40:1353–62


14. Groves C, Reid OG, Ginger DS. 2010. Heterogeneity in polymer solar cells: local morphology and
performance in organic photovoltaics studied with scanning probe microscopy. Acc. Chem. Res. 43:612–
20
15. Liang YY, Yu LP. 2010. A new class of semiconducting polymers for bulk heterojunction solar cells with
exceptionally high performance. Acc. Chem. Res. 43:1227–36
16. van Bavel S, Veenstra S, Loos J. 2010. On the importance of morphology control in polymer solar cells.
Macromol. Rapid Commun. 31:1835–45
17. Clarke TM, Durrant JR. 2010. Charge photogeneration in organic solar cells. Chem. Rev. 110:6736–67
18. Brabec CJ, Gowrisanker S, Halls JJM, Laird D, Jia SJ, Williams SP. 2010. Polymer-fullerene bulk-
heterojunction solar cells. Adv. Mater. 22:3839–56
19. Mullen K, Scherf U, eds. 2006. Organic Light Emitting Devices: Synthesis, Properties and Applications.
New York: Wiley
20. Burroughes JH, Bradley DDC, Brown AR, Marks RN, Mackay K, et al. 1990. Light-emitting diodes
based on conjugated polymers. Nature 347:539–41
21. Friend RH, Gymer RW, Holmes AB, Burroughes JH, Marks RN, et al. 1999. Electroluminescence in
conjugated polymers. Nature 397:121–28
22. Kafafi ZH, ed. 2002. Organic Light-Emitting Materials and Devices V. Washington, DC: SPIE
23. Adachi C, Baldo MA, Thompson ME, Forrest SR. 2001. Nearly 100% internal phosphorescence effi-
ciency in an organic light emitting device. J. Appl. Phys. 90:5048–52
24. Spano FC. 2005. Modeling disorder in polymer aggregates: the optical spectroscopy of regioregular
poly(3-hexylthiophene) thin films. J. Chem. Phys. 122:234701. Erratum. 2007. J. Chem. Phys. 126:159901
25. Clark J, Silva C, Friend RH, Spano FC. 2007. Role of intermolecular coupling in the photophysics of
disordered organic semiconductors: aggregate emission in regioregular polythiophene. Phys. Rev. Lett.
98:206406
26. Clark J, Chang JF, Spano FC, Friend RH, Silva C. 2009. Determining exciton bandwidth and film
microstructure in polythiophene films using linear absorption spectroscopy. Appl. Phys. Lett. 94:163306
27. Spano FC, Clark J, Silva C, Friend RH. 2009. Determining exciton coherence from the photolumines-
cence spectral line shape in poly(3-hexylthiophene) thin films. J. Chem. Phys. 130:074904
28. Yamagata H, Spano FC. 2011. Vibronic coupling in quantum wires: application to polydiacetylene.
J. Chem. Phys. 135:054906
29. Hochstrasser RM, Kasha M. 1964. Application of the exciton model to monomolecular lamellar systems.
Photochem. Photobiol. 3:317–31
30. Kasha M. 1963. Energy transfer mechanisms and the molecular exciton model for molecular aggregates.
Radiat. Res. 20:55–70
31. McRae EG, Kasha M. 1958. Enhancement of phosphorescence ability upon aggregation of dye molecules.
J. Chem. Phys. 28:721–22
32. Yan M, Rothberg LJ, Papadimitrakopoulos F, Galvin ME, Miller TM. 1994. Spatially indirect excitons
as primary photoexcitations in conjugated polymers. Phys. Rev. Lett. 72:1104–7

496 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

33. Yan M, Rothberg LJ, Kwock EW, Miller TM. 1995. Interchain excitations in conjugated polymers. Phys.
Rev. Lett. 75:1992–95
34. Spano FC. 2006. Absorption in regioregular poly(3-hexylthiophene) thin films: Fermi resonances, in-
terband coupling and disorder. Chem. Phys. 325:22–35
35. Spano FC. 2010. The spectral signatures of Frenkel polarons in H- and J-aggregates. Acc. Chem. Res.
43:429–39
36. Spano FC, Yamagata H. 2011. Vibronic coupling in J-aggregates and beyond: a direct means of determin-
ing the exciton coherence length from the photoluminescence spectrum. J. Phys. Chem. B 115:5133–43
37. Brown PJ, Thomas SD, Köhler A, Wilson JS, Kim J-S, et al. 2003. Effect of interchain interactions on
the absorption and emission of poly(3-hexylthiophene). Phys. Rev. B 67:064203
38. Barford W. 2013. Excitons in conjugated polymers: a tale of two particles. J. Phys. Chem. A 117:2665–71
39. Barisien T, Legrand L, Weiser G, Deschamps J, Balog M, et al. 2007. Exciton spectroscopy of red
polydiacetylene chains in single crystals. Chem. Phys. Lett. 444:309–13
40. Dubin F, Berréhar J, Grousson R, Guillet T, Lapersonne-Meyer C, et al. 2002. Optical evidence of
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

a purely one-dimensional exciton density of states in a single conjugated polymer chain. Phys. Rev. B
66:113202
41. Dubin F, Berréhar J, Grousson R, Schott M, Voliotis V. 2006. Evidence of polariton-induced trans-
parency in a single organic quantum wire. Phys. Rev. B 73:121302
42. Dubin F, Melet R, Barisien T, Grousson R, Legrand L, et al. 2006. Macroscopic coherence of a single
exciton state in an organic quantum wire. Nat. Phys. 2:32–35
43. Horvath A, Weiser G, Lapersonne-Meyer C, Schott M, Spagnoli S. 1996. Wannier excitons and Franz-
Keldysh effect of polydiacetylene chains diluted in their single crystal monomer matrix. Phys. Rev. B
53:13507–14
44. Lecuiller R, Berréhar J, Ganière JD, Lapersonne-Meyer C, Lavallard P, Schott M. 2002. Fluorescence
yield and lifetime of isolated polydiacetylene chains: evidence for a one-dimensional exciton band in a
conjugated polymer. Phys. Rev. B 66:125205
45. Lecuiller R, Berréhar J, Lapersonne-Meyer C, Schott M, Ganière JD. 1999. Fluorescence quantum yield
and lifetime of ‘red’ polydiacetylene chains isolated in their crystalline monomer matrix. Chem. Phys. Lett.
314:255–60
46. Schott M. 2006. Optical properties of single conjugated polymer chains (polydiacetylenes). In Photophysics
of Molecular Materials: From Single Molecules to Single Crystals, ed. G Lanzani, pp. 49–145. Weinheim,
Ger.: Wiley-VCH
47. Yamagata H, Spano FC. 2012. Interplay between intrachain and interchain interactions in semiconduct-
ing polymer assemblies: the HJ-aggregate model. J. Chem. Phys. 136:184901. Erratum. 2012. J. Chem.
Phys. 137:249901
48. Soos ZG, Hayden GW, McWilliams PCM, Etemad S. 1990. Excitation shifts of parallel conjugated
polymers due to π -electron dispersion forces. J. Chem. Phys. 93:7439–48
49. Manas ES, Spano FC. 1998. Absorption and spontaneous emission in aggregates of π -conjugated
molecules. J. Chem. Phys. 109:8087–101
50. Cornil J, dos Santos DA, Crispin X, Silbey R, Bredas JL. 1998. Influence of interchain interactions
on the absorption and luminescence of conjugated oligomers and polymers: a quantum mechanical
characterization. J. Am. Chem. Soc. 120:1289–99
51. Gierschner J, Huang YS, Van Averbeke B, Cornil J, Friend RH, Beljonne D. 2009. Excitonic ver-
sus electronic couplings in molecular assemblies: the importance of non-nearest neighbor interactions.
J. Chem. Phys. 130:044105
52. Barford W. 2007. Exciton transfer integrals between polymer chains. J. Chem. Phys. 126:134905
53. Banerji N, Cowan S, Vauthey E, Heeger AJ. 2011. Ultrafast relaxation of the poly(3-hexylthiophene)
emission spectrum. J. Phys. Chem. C 115:9726–39
54. Parkinson P, Mueller C, Stingelin N, Johnston MB, Herz LM. 2010. Role of ultrafast torsional relaxation
in the emission from polythiophene aggregates. J. Phys. Chem. Lett. 1:2788–92
55. Banerji N. 2013. Sub-picosecond delocalization in the excited state of conjugated homopolymers and
donor-acceptor copolymers. J. Mater. Chem. C 1:3052–66

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 497


PC65CH21-Silva ARI 10 March 2014 15:47

56. Niles ET, Roehling JD, Yamagata H, Wise AJ, Spano FC, et al. 2012. J-aggregate behavior in poly-3-
hexylthiophene nanofibers. J. Phys. Chem. Lett. 3:259–63
57. Pichler K, Halliday DA, Bradley DDC, Burns PL, Friend RH, Holmes AB. 1993. Optical spectroscopy
of highly ordered poly( p-phenylene vinylene). J. Phys. Condens. Matter 5:7155–72
58. da Silva MAT, Dias IFL, Duarte JL, Laureto E, Silvestre I, et al. 2008. Identification of the optically
active vibrational modes in the photoluminescence of MEH-PPV films. J. Chem. Phys. 128:094902
59. Chang R, Hsu JH, Fann WS, Yu J, Lin SH, et al. 2000. Aggregated states of luminescent conjugated
polymers in solutions. Chem. Phys. Lett. 317:153–58
60. Zeng QG, Ding ZJ. 2004. Photoluminescence and Raman spectra study of para-phenylenevinylene at
low temperatures. J. Phys. Condens. Matter 16:5171–78
61. Ho PKH, Kim J-S, Tessler N, Friend RH. 2001. Photoluminescence of poly( p-phenylenevinylene)-silica
nanocomposites: evidence of duel emission by Franck-Condon analysis. J. Chem. Phys. 115:2709–20
62. Hagler TW, Pakbaz K, Voss KF, Heeger AJ. 1991. Enhanced order and electronic delocalization in
conjugated polymers oriented by gel processing in polyethylene. Phys. Rev. B 44:8652–66
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

63. Rauscher U, Bässler H, Bradley DDC, Hennecke M. 1990. Exciton versus band description of the
absorption and luminescence spectra in PPV. Phys. Rev. B 42:9830–36
64. Hayes GR, Samuel IDW, Phillips RT. 1997. Polarization dependence of the ultrafast photoluminescence
of oriented PPV. Phys. Rev. B 56:3838–43
65. Soci C, Comoretto D, Marabelli F, Moses D. 2007. Anisotropic photoluminescence properties of ori-
ented poly( p-phenylene-vinylene) films: effects of dispersion of optical constants. Phys. Rev. B 75:075204
66. Ho PKH, Friend RH. 2002. π -Electronic and electrical transport properties of conjugated polymer
nanocomposites: poly( p-phenylenevinylene) with homogeneously dispersed silica nanoparticles. J. Chem.
Phys. 116:6782–94
67. Köhler A, Hoffmann ST, Bässler H. 2012. An order-disorder transition in the conjugated polymer
MEH-PPV. J. Am. Chem. Soc. 134:11594–601
68. Yu ZH, Barbara PF. 2004. Low-temperature single-molecule spectroscopy of MEH-PPV conjugated
polymer molecules. J. Phys. Chem. B 108:11321–26
69. Barbara PF, Gesquiere AJ, Park S-J, Lee YJ. 2005. Single-molecule spectroscopy of conjugated polymers.
Acc. Chem. Res. 38:602–10
70. Peet J, Brocker E, Xu YH, Bazan GC. 2008. Controlled β-phase formation in poly(9,9-di-n-
octylfluorene) by processing with alkyl additives. Adv. Mater. 20:1882–85
71. Khan ALT, Sreearunothai P, Herz LM, Banach MJ, Köhler A. 2004. Morphology-dependent energy
transfer within polyfluorene thin films. Phys. Rev. B 69:085201
72. Hwang I, Scholes GD. 2011. Electronic energy transfer and quantum-coherence in π -conjugated poly-
mers. Chem. Mater. 23:610–20
73. Collini E, Scholes GD. 2009. Coherent intrachain energy migration in a conjugated polymer at room
temperature. Science 323:369–73
74. Engel GS, Calhoun TR, Read EL, Ahn TK, Mancal T, et al. 2007. Evidence for wavelike energy transfer
through quantum coherence in photosynthetic systems. Nature 446:782–86
75. Panitchayangkoon G, Hayes D, Fransted KA, Caram JR, Harel E, et al. 2010. Long-lived quantum coher-
ence in photosynthetic complexes at physiological temperature. Proc. Natl. Acad. Sci. USA 107:12766–70
76. Paquin F, Yamagata H, Hestand NJ, Sakowicz M, Bérubé N, et al. 2013. Two-dimensional spatial
coherence of excitons in semicrystalline polymeric semiconductors: the effect of molecular weight. Phys.
Rev. B 88:155202
77. Kobayashi T, Hamazaki J, Kunugita H, Ema K, Endo T, et al. 2003. Coexistence of photoluminescence
from two intrachain states in polythiophene films. Phys. Rev. B 67:205214
78. Jiang XM, Osterbacka R, Korovyanko O, An CP, Horovitz B, et al. 2002. Spectroscopic studies of
photoexcitations in regioregular and regiorandom polythiophene films. Adv. Funct. Mater. 12:587–97
79. Koren AB, Curtis MD, Francis AH, Kampf JW. 2003. Intermolecular interactions in π -stacked conju-
gated molecules: synthesis, structure, and spectral characterization of alkyl bithiazole oligomers. J. Am.
Chem. Soc. 125:5040–50
80. Holstein T. 1959. Polaron motion. I. Molecular-crystal model. Ann. Phys. 8:325–42

498 Spano · Silva


PC65CH21-Silva ARI 10 March 2014 15:47

81. Beljonne D, Cornil J, Silbey R, Millie P, Bredas JL. 2000. Interchain interactions in conjugated materials:
the exciton model versus the supermolecular approach. J. Chem. Phys. 112:4749–58
82. Westenhoff S, Abrusci A, Feast WJ, Henze O, Kilbinger AFM, et al. 2006. Supramolecular electronic
coupling in chiral oligothiophene nanostructures. Adv. Mater. 18:1281–85
83. Scharsich C, Lohwasser RH, Sommer M, Asawapirom U, Scherf U, et al. 2012. Control of aggregate
formation in poly(3-hexylthiophene) by solvent, molecular weight, and synthetic method. J. Polym. Sci.
B 50:442–53
84. Meskers SCJ, Janssen RAJ, Haverkort JEM, Wolter JH. 2000. Relaxation of photo-excitations in films
of oligo- and poly-( para-phenylene vinylene) derivatives. Chem. Phys. 260:415–39
85. Knapp EW. 1984. Lineshapes of molecular aggregates: exchange narrowing and intersite correlation.
Chem. Phys. 85:73–82
86. Spano FC, Meskers SCJ, Hennebicq E, Beljonne D. 2007. Probing excitation delocalization in
supramolecular chiral stacks by means of circularly polarized light: experiment and modeling. J. Am.
Chem. Soc. 129:7044–54
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

87. Dahlbom M, Pullerits T, Mukamel S, Sundström V. 2001. Exciton delocalization in the B850 light-
harvesting complex: comparison of different measures. J. Phys. Chem. B 105:5515–24
88. Meier T, Chernyak V, Mukamel S. 1997. Multiple exciton coherence sizes in photosynthetic antenna
complexes viewed by pump-probe spectroscopy. J. Phys. Chem. B 101:7332–42
89. Meier T, Zhao Y, Chernyak V, Mukamel S. 1997. Polarons, localization, and excitonic coherence in
superradiance of biological antenna complexes. J. Chem. Phys. 107:3876–93
90. Kuhn O, Sundström V. 1997. Pump-probe spectroscopy of dissipative energy transfer dynamics in
photosynthetic antenna complexes: a density matrix approach. J. Chem. Phys. 107:4154–64
91. Kohler BE, Schilke DE. 1987. Low-lying singlet states of a short polydiacetylene oligomer. J. Chem.
Phys. 86:5214–15
92. Citrin DS. 1992. Long intrinsic radiative lifetimes of excitons in quantum wires. Phys. Rev. Lett. 69:3393–
96
93. Deboer S, Wiersma DA. 1990. Dephasing-induced damping of superradiant emission in J-aggregates.
Chem. Phys. Lett. 165:45–53
94. Fidder H, Knoester J, Wiersma DA. 1990. Superradiant emission and optical dephasing in J-aggregates.
Chem. Phys. Lett. 171:529–36
95. Khachatryan B, Nguyen TD, Vardeny ZV, Ehrenfreund E. 2012. Phosphorescence superradiance in a
Pt-containing π -conjugated polymer. Phys. Rev. B 86:195203
96. Burn PL, Bradley DDC, Friend RH, Halliday DA, Holmes AB, et al. 1992. Precursor route chemistry
and electronic properties of poly( p-phenylene-vinylene), poly(2,5-dimethyl-p-phenylene)vinylene and
poly(2,5-dimethoxy-p-phenylene)vinylene. J. Chem. Soc. Perkin Trans. 1:3225–31
97. Schwartz BJ. 2003. Conjugated polymers as molecular materials: how chain conformation and film
morphology influence energy transfer and interchain interactions. Annu. Rev. Phys. Chem. 54:141–72
98. Nguyen T-Q, Martini IB, Liu J, Schwartz BJ. 2000. Controlling interchain interactions in conjugated
polymers: the effects of chain morphology on exciton-exciton annihilation and aggregation in MEH-PPV
films. J. Phys. Chem. B 104:237–55
99. Collison CJ, Rothberg LJ, Treemaneekarn V, Li Y. 2001. Conformational effects on the photophysics
of conjugated polymers: a two species model for MEH-PPV spectroscopy and dynamics. Macromolecules
34:2346–52
100. Grey JK, Kim DY, Norris BC, Miller WL, Barbara PF. 2006. Size-dependent spectroscopic properties
of conjugated polymer nanoparticles. J. Phys. Chem. B 110:25568–72
101. Lin HZ, Hania RP, Bloem R, Mirzov O, Thomsson D, Scheblykin IG. 2010. Single chain versus single
aggregate spectroscopy of conjugated polymers: Where is the border? Phys. Chem. Chem. Phys. 12:11770–
77
102. Mirzov O, Scheblykin IG. 2006. Photoluminescence spectra of a conjugated polymer: from films and
solutions to single molecules. Phys. Chem. Chem. Phys. 8:5569–76
103. Chang R, Hsu JH, Fann WS, Liang KK, Chiang CH, et al. 2000. Experimental and theoretical investi-
gations of absorption and emission spectra of the light-emitting polymer MEH-PPV in solution. Chem.
Phys. Lett. 317:142–52

www.annualreviews.org • H- and J-Aggregates in Polymeric Semiconductors 499


PC65CH21-Silva ARI 10 March 2014 15:47

104. Yamagata H, Hestand NJ, Spano FC, Köhler A, Scharsich C, et al. 2013. The red phase of
poly[2-methoxy-5-(2-ethylhexyloxy)-1,4-phenylenevinylene] (MEH-PPV): a disordered HJ-aggregate.
J. Chem. Phys. 139:114903
105. Cadby AJ, Lane PA, Mellor H, Martin SJ, Grell M, et al. 2000. Film morphology and photophysics of
polyfluorene. Phys. Rev. B 62:15604–9
106. Kline RJ, McGehee MD, Kadnikova EN, Liu JS, Fréchet JMJ, Toney MF. 2005. Dependence of
regioregular poly(3-hexylthiophene) film morphology and field-effect mobility on molecular weight.
Macromolecules 38:3312–19
107. Kline RJ, McGehee MD, Kadnikova EN, Liu JS, Fréchet JMJ. 2003. Controlling the field-effect mobility
of regioregular polythiophene by changing the molecular weight. Adv. Mater. 15:1519–22
108. Goh C, Kline RJ, McGehee MD, Kadnikova EN, Fréchet JMJ. 2005. Molecular-weight-dependent
mobilities in regioregular poly(3-hexyl-thiophene) diodes. Appl. Phys. Lett. 86:122110
109. Zhang R, Li B, Iovu MC, Jeffries-EL M, Sauve G, et al. 2006. Nanostructure dependence of field-
effect mobility in regioregular poly(3-hexylthiophene) thin film field effect transistors. J. Am. Chem. Soc.
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

128:3480–81
110. Koch FPV, Rivnay J, Foster S, Müller C, Downing J, et al. 2013. Microstructure development with
molecular weight in semicrystalline polymer semiconductors and its influence on charge transport:
poly(3-hexylthiophene), a model study. Prog. Polym. Sci. In press
111. Ballantyne AM, Chen L, Dane J, Hammant T, Braun FM, et al. 2008. The effect of poly(3-
hexylthiophene) molecular weight on charge transport and the performance of polymer:fullerene solar
cells. Adv. Funct. Mater. 18:2373–80
112. Noriega R, Rivnay J, Vandewal K, Koch FPV, Stingelin N, et al. 2013. A general relationship between
disorder, aggregation, and charge transport in conjugated polymers. Nat. Mater. 12:1038–44
113. Hellmann C, Paquin F, Treat ND, Bruno A, Reynolds LX, et al. 2013. Controlling the interaction of
light with polymer semiconductors. Adv. Mater. 25:4906–11
114. Wunderlich B. 1976. Macromolecular Physics, Vol. 2: Crystal Nucleation, Growth, Annealing. San Diego:
Academic
115. Yamagata H, Pochas CM, Spano FC. 2012. Designing J- and H-aggregates through wave function
overlap engineering: applications to poly(3-hexylthiophene). J. Phys. Chem. B 116:14494–503
116. Paquin F, Latini G, Sakowicz M, Karsenti P-L, Wang L, et al. 2011. Charge separation in semicrystalline
polymeric semiconductors by photoexcitation: Is the mechanism intrinsic or extrinsic? Phys. Rev. Lett.
106:197401
117. Paquin F, Rivnay J, Salleo A, Stingelin N, Silva C. 2013. Multi-phase semicrystalline microstructures
drive exciton dissociation in neat plastic semiconductors. Phys. Rev. X. Manuscript submitted
118. Clark J. 2009. Intermolecular interactions in π -conjugated molecules: optical probes of chain conformation. PhD
Diss., Univ. Cambridge, Cambridge, UK
119. Baghgar M, Labastide J, Bokel F, Dujovne I, McKenna A, et al. 2012. Probing inter- and intrachain
exciton coupling in isolated poly(3-hexylthiophene) nanofibers: effect of solvation and regioregularity.
J. Phys. Chem. Lett. 3:1674–79
120. Baghgar M, Pentzer E, Wise AJ, Labastide JA, Emrick T, Barnes MD. 2013. Cross-linked functionalized
poly(3-hexylthiophene) nanofibers with tunable excitonic coupling. ACS Nano 7:8917–23

500 Spano · Silva


PC65-FrontMatter ARI 17 February 2014 14:33

Annual Review of
Physical Chemistry
Contents Volume 65, 2014

A Journey Through Chemical Dynamics


William H. Miller p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Chemistry of Atmospheric Nucleation: On the Recent Advances on
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

Precursor Characterization and Atmospheric Cluster Composition


in Connection with Atmospheric New Particle Formation
M. Kulmala, T. Petäjä, M. Ehn, J. Thornton, M. Sipilä, D.R. Worsnop,
and V.-M. Kerminen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p21
Multidimensional Time-Resolved Spectroscopy of Vibrational
Coherence in Biopolyenes
Tiago Buckup and Marcus Motzkus p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p39
Phase Separation in Bulk Heterojunctions of Semiconducting
Polymers and Fullerenes for Photovoltaics
Neil D. Treat and Michael L. Chabinyc p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p59
Nitrogen-Vacancy Centers in Diamond: Nanoscale Sensors for
Physics and Biology
Romana Schirhagl, Kevin Chang, Michael Loretz, and Christian L. Degen p p p p p p p p p p p p p83
Superresolution Localization Methods
Alexander R. Small and Raghuveer Parthasarathy p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 107
The Structure and Dynamics of Molecular Excitons
Christopher J. Bardeen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 127
Advanced Potential Energy Surfaces for Condensed Phase Simulation
Omar Demerdash, Eng-Hui Yap, and Teresa Head-Gordon p p p p p p p p p p p p p p p p p p p p p p p p p p p p 149
Ion Mobility Analysis of Molecular Dynamics
Thomas Wyttenbach, Nicholas A. Pierson, David E. Clemmer,
and Michael T. Bowers p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 175
State-to-State Spectroscopy and Dynamics of Ions and Neutrals by
Photoionization and Photoelectron Methods
Cheuk-Yiu Ng p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 197
Imaging Fluorescence Fluctuation Spectroscopy: New Tools for
Quantitative Bioimaging
Nirmalya Bag and Thorsten Wohland p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 225

v
PC65-FrontMatter ARI 17 February 2014 14:33

Elucidation of Intermediates and Mechanisms in Heterogeneous


Catalysis Using Infrared Spectroscopy
Aditya Savara and Eric Weitz p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 249
Physicochemical Mechanism of Light-Driven DNA Repair
by (6-4) Photolyases
Shirin Faraji and Andreas Dreuw p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 275
Advances in the Determination of Nucleic Acid
Conformational Ensembles
Loı̈c Salmon, Shan Yang, and Hashim M. Al-Hashimi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 293
The Role of Ligands in Determining the Exciton Relaxation Dynamics
Access provided by National Cheng Kung University on 12/18/23. For personal use only.

in Semiconductor Quantum Dots


Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

Mark D. Peterson, Laura C. Cass, Rachel D. Harris, Kedy Edme, Kimberly Sung,
and Emily A. Weiss p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 317
Laboratory-Frame Photoelectron Angular Distributions in Anion
Photodetachment: Insight into Electronic Structure and
Intermolecular Interactions
Andrei Sanov p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 341
Quantum Heat Engines and Refrigerators: Continuous Devices
Ronnie Kosloff and Amikam Levy p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 365
Approaches to Single-Nanoparticle Catalysis
Justin B. Sambur and Peng Chen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 395
Ultrafast Carrier Dynamics in Nanostructures for Solar Fuels
Jason B. Baxter, Christiaan Richter, and Charles A. Schmuttenmaer p p p p p p p p p p p p p p p p p p 423
Nucleation in Polymers and Soft Matter
Xiaofei Xu, Christina L. Ting, Isamu Kusaka, and Zhen-Gang Wang p p p p p p p p p p p p p p p p 449
H- and J-Aggregate Behavior in Polymeric Semiconductors
Frank C. Spano and Carlos Silva p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 477
Cold State-Selected Molecular Collisions and Reactions
Benjamin K. Stuhl, Matthew T. Hummon, and Jun Ye p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 501
Band Excitation in Scanning Probe Microscopy: Recognition and
Functional Imaging
S. Jesse, R.K. Vasudevan, L. Collins, E. Strelcov, M.B. Okatan, A. Belianinov,
A.P. Baddorf, R. Proksch, and S.V. Kalinin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 519
Dynamical Outcomes of Quenching: Reflections on a
Conical Intersection
Julia H. Lehman and Marsha I. Lester p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 537
Bimolecular Recombination in Organic Photovoltaics
Girish Lakhwani, Akshay Rao, and Richard H. Friend p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 557

vi Contents
PC65-FrontMatter ARI 17 February 2014 14:33

Mapping Atomic Motions with Ultrabright Electrons: The Chemists’


Gedanken Experiment Enters the Lab Frame
R.J. Dwayne Miller p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 583
Optical Spectroscopy Using Gas-Phase Femtosecond
Laser Filamentation
Johanan Odhner and Robert Levis p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 605

Indexes

Cumulative Index of Contributing Authors, Volumes 61–65 p p p p p p p p p p p p p p p p p p p p p p p p p p p 629


Cumulative Index of Article Titles, Volumes 61–65 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 632
Access provided by National Cheng Kung University on 12/18/23. For personal use only.
Annu. Rev. Phys. Chem. 2014.65:477-500. Downloaded from www.annualreviews.org

Errata

An online log of corrections to Annual Review of Physical Chemistry articles may be


found at http://www.annualreviews.org/errata/physchem

Contents vii

You might also like