You are on page 1of 13

Journal of Physics B: Atomic, Molecular and Optical Physics

PAPER You may also like


- Self-assembled semiconducting
Impact of charge-transfer excitons in regioregular monolayers in organic electronics
Alexey S. Sizov, Elena V. Agina and
polythiophene on the charge separation at Sergey A. Ponomarenko

- Fine-tuning of electronic properties in


polythiophene-fullerene heterojunctions donor–acceptor conjugated polymers
based on oligothiophenes
Ichiro Imae, Hitoshi Sagawa and Yutaka
To cite this article: M Polkehn et al 2018 J. Phys. B: At. Mol. Opt. Phys. 51 014003 Harima

- The effect of crystallinity in donor groups


on the performance of photovoltaic
devices based on an
oligothiophene–fullerene dyad
View the article online for updates and enhancements. Takeshi Nishizawa, Keisuke Tajima and
Kazuhito Hashimoto

This content was downloaded from IP address 203.110.242.12 on 15/06/2022 at 07:14


Journal of Physics B: Atomic, Molecular and Optical Physics

J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 (12pp) https://doi.org/10.1088/1361-6455/aa93d0

Impact of charge-transfer excitons in


regioregular polythiophene on the charge
separation at polythiophene-fullerene
heterojunctions
M Polkehn1, H Tamura2 and I Burghardt1
1
Institute for Physical and Theoretical Chemistry, Goethe University Frankfurt, Max-von-Laue-Str. 7,
D-60438 Frankfurt, Germany
2
Department of Chemical System Engineering, School of Engineering, The University of Tokyo, Tokyo
113-8656, Japan

E-mail: burghardt@chemie.uni-frankfurt.de

Received 6 July 2017, revised 1 October 2017


Accepted for publication 16 October 2017
Published 30 November 2017

Abstract
This study addresses the mechanism of ultrafast charge separation in regioregular
oligothiophene-fullerene assemblies representative of poly-3-hexylthiophene
(P3HT)-[6,6]-phenyl-C61 butyric acid methyl ester (PCBM) heterojunctions, with special
emphasis on the inclusion of charge transfer excitons in the oligothiophene phase. The formation
of polaronic inter-chain charge separated species in highly ordered oligothiophene has been
demonstrated in recent experiments and could have a significant impact on the net charge
transfer to the fullerene acceptor. The present approach combines a first-principles parametrized
multi-site Hamiltonian, based on time-dependent density functional theory calculations, with
accurate quantum dynamics simulations using the multi-layer multi-configuration time-
dependent Hartree method. Quantum dynamical studies are carried out for up to 182 electronic
states and 112 phonon modes. The present analysis follows up on our previous study of (Huix-
Rotllant et al 2015 J. Phys. Chem. Lett. 6 1702) and significantly expands the scope of this
analysis by including the dynamical role of charge transfer excitons. Our investigation highlights
the pronounced mixing of photogenerated Frenkel excitons with charge transfer excitons in the
oligothiophene domain, and the opening of new transfer channels due the creation of such
charge-separated species. As a result, it turns out that the interfacial donor/acceptor charge
transfer state can be largely circumvented due to the presence of charge transfer excitons.
However, the latter states in turn act as a trap, such that the free carrier yield observed on
ultrafast time scales is tangibly reduced. The present analysis underscores the complexity of the
transfer pathways at P3HT-PCBM type junctions.
Keywords: quantum dynamics, organic photovoltaics, excitons

(Some figures may appear in colour only in the online journal)

1. Introduction investigated over recent years [1–3]. A particular focus has


been directed at investigating the conditions facilitating ultra-
Exciton dissociation in prototypical blends of poly-3- fast long-range electron–hole separation, as experimentally
hexylthiophene (P3HT) and the fullerene derivative observed especially in regioregular materials [4, 5]. A priori, a
[6,6]-phenyl-C61 butyric acid methyl ester (PCBM), and related much slower, thermally activated dynamics is expected in the
polymer/fullerene combinations have been extensively presence of an effective Coulomb barrier, according to the

0953-4075/18/014003+12$33.00 1 © 2017 IOP Publishing Ltd Printed in the UK


J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Onsager–Braun model [6]. To explain these observations, band [32]. As mentioned above, these Frenkel-type species
transient effects due to hot states [5, 7, 8], charge delocalization may be intimately coupled, though, to charge transfer exci-
[3, 7, 9–11], the role of the driving energy [7, 12], destabili- tons involving inter-chain charge separation [20–22]. Both
zation of the interfacial charge transfer state [13], the details of types of species are coupled, in turn, to donor/acceptor
the microelectrostatics at the interface [14, 15], and entropic charge separated states. The coupling and interconversion
effects [16] have been invoked. In our recent work [17, 18], we between these three types of electronic configurations will be
have pointed out that the effective lowering of the Coulomb at the center of our discussion.
barrier due to charge delocalization, and the vibronically hot In line with the notation introduced in [17, 18, 23], an
nature of the primary excitonic and charge transfer states can e–h lattice Hamiltonian is formulated and parametrized based
play a central role. upon electronic structure (here, TDDFT) calculations for
In the present study, we build upon our recent investi- relevant fragments. To simplify the model, the fullerene
gations [17–19] of regioregular P3HT-PCBM type systems moiety is represented by an effective, coarse-grained super-
and extend the latter studies so as to include charge separated particle site (ν = 0) [17, 18], while N oligothiophene frag-
species within the P3HT type donor domain. Indeed, as has ments are treated explicitly as individual oligomer sites
been shown in recent experimental investigations [20–22], (ν = 1, K, N). Localized e–h pairs ∣nnñ (for ν = 1, K, N)
charge transfer excitons, or polarons, are observed in oli- correspond to Frenkel excitonic (XT) configurations on the
gothiophene phases, presumably by inter-chain polaron pair OT moieties, ∣XTnñ = ∣n = n , m = nñ. Conversely, non-local
formation in highly ordered, lamellar domains. This raises the e–h states ∣nmñ, m ¹ n (ν = 0, μ = 1, K, N), represent charge
question whether the presence of charge transfer excitons separated (CS) configurations, ∣CSn ñ = ∣n , m = n + nñ,
affects the yield of charge separated species between donor where ν = 0. The CS1 state corresponds to the interfacial
and acceptor domains and, hence, the power conversion charge-transfer state (denoted CT in [17]). Finally, charge-
efficiency. Indeed, a non-negligible amount of polaronic separated inter-chain species in the OT domain are included,
species in the donor phase could open up additional transfer i.e., charge transfer exciton configurations of type ∣CTXn, n ¢ñ =
pathways and enhance the effective charge transfer to the ∣n = n , m = n¢ñ where n , m = 1,¼, N .
acceptor phase. Conversely, polaronic species could turn out In the dynamical treatment detailed below, delocalized
to be stabilized in the donor phase and effectively reduce the Frenkel excitonic states, and delocalized mixtures of Frenkel
reservoir of active electron–hole pairs available for interfacial and charge transfer excitonic states arise as coherent super-
charge separation. positions of e–h configurations ∣nmñ; for the full wavefunction
To investigate the role of charge transfer excitons in the including phonon modes, see equation (7).
oligothiophene domain, we use electronic structure studies at The Hamiltonian is defined in the above e–h basis, in line
a time-dependent density functional theory (TDDFT) level of with [17, 18], i.e.,
relevant molecular fragments to parametrize a multi-state
lattice Hamiltonian [17–19, 23], which is subsequently el
Hˆ = Hˆ + Hˆ
e - ph
combined with quantum dynamical simulations using the on ‐ site coup e - ph
multi-layer multi-configuration time-dependent Hartree (ML- = (Hˆ + Hˆ ) + Hˆ (1 )
MCTDH) method [24–30]. In the present study, high-
el
dimensional lattice models with up to 182 electronic states comprising an electronic part (Ĥ ) and an electron–phonon
and 112 vibrational (phonon) modes are investigated. This coupling part (Ĥ
e - ph
). The electronic part, in turn, is split into
model contains the essential ingredients permitting the on-site contributions (Hˆ
on ‐ site
) and electronic couplings
investigation of the above questions. coup
(Ĥ ). In the following, the individual parts of the Hamil-
The remainder of this article is organized as follows. We
tonian are detailed.
first describe the electron–hole lattice model that is employed
in the present study. Next, we illustrate the modified ener-
2.1. On-site Hamiltonian
getics of the system in the presence of charge transfer exci-
tons. We then turn to a high-dimensional quantum dynamical The on-site part of the Hamiltonian is given as
investigation using the ML-MCTDH method. Finally, we
offer a discussion and conclusion. N N
=  XT å ∣XTnñáXTn∣ + å  nCS∣CSn ñáCSn ∣
on ‐ site

n=1 n=1
N N
2. Electron–hole lattice Hamiltonian
+ åå n ¢= 1
n, n ¢ ∣CTXn, n ¢ñáCTXn, n ¢∣ ,
 CTX (2 )
n=1
In the model system depicted in figure 1, representing a n ¢¹ n

regioregular oligothiophene (OT)-fullerene domain repre-


sentative of P3HT-PCBM, the primary OT photoexcitations where the excitonic states ∣XTnñ are taken to have identical on-
in the self-assembled lamellar structures are inter-chain, site energies ( XT ) while the energies of the charge-separated
H-aggregate type excitons [31, 32]. These delocalized exci- states ∣CSn ñ and the OT charge transfer excitonic states
tations are typical of face-to-face aligned aggregates whose ∣CTXn, n ¢ñ are determined by an effective Coulomb barrier as
bright state is located at the high-energy edge of the excitonic further discussed next.

2
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

The effective barrier potential that defines the on-site couplings within the XT and CS manifolds are nearest-
energies of the ∣CSn ñ states is depicted in figure 1 (lower neighbor interactions. As a new feature of the extended model
panel); this effective potential describes the Coulombic including OT charge transfer excitons, a coupling arises
interaction between delocalized electron and hole distribu- between the ∣CTX1,2ñ state and the ∣CS2 ñ state (κ = 0.16 eV),
tions, and includes the effect of the internal electric field. As by electron transfer to the fullerene moiety. Next, the Frenkel-
detailed in [17, 19, 33], the barrier was determined from DFT type excitonic ∣XTnñ states couple to the charge transfer
calculations in conjunction with a tight-binding model, where excitons ∣CTXn, n ¢ñ either by electron transfer to one of the
a representative field value of E = 10 V μm−1 was used. The neighboring sites (k1 = 0.357 eV) or by hole transfer to the
barrier height is a sensitive function of electron delocalization neighboring sites (k 2 = 0.139 eV).
within the fullerene moiety and, hence, depends on the The following terms in equation (3) describe the coupling
effective size of the fullerene super-particle in our model. In between nearest-neighbor excitons ( j = 0.1 eV), the hole
the present study, two characteristic barrier heights are con- transfer between neighboring ∣CSn ñ states (th = -0.12 eV),
sidered: (i) around 0.25 eV, for small fullerene clusters (here, along with electron transfer (le = 0.08 eV) and hole transfer
a 7-C60 cluster) representative of disordered domains, (ii) (lh = 0.08 eV) between neighboring ∣CTXn, n ¢ñ states. Note
around 0.05 eV, as a result of delocalization across several that le and lh are taken to be equal, which is a reasonable
fullerene layers (here, five 61-C60 layers) [17]. These barrier approximation for the present system.
types will be referred to as barrier I and barrier II in the Table 1 lists all electronic couplings, and figure 2 illus-
following (see figure 1). trates the relevant electronic configurations and couplings for
Likewise, the on-site energies of the OT charge transfer a minimal model comprising two donor sites. All electronic
excitonic states ∣CTXn, n ¢ñ are determined by a Coulomb bar- couplings were computed by a diabatization procedure as
rier, which was taken to be of the standard form detailed further in [33–35]. Clearly, a complex scenario of
CTX
Vbarrier = -1 (4p 0 r rnn ¢ ). Here the relative permittivity coupled pathways emerges due to the presence of the CTX
r = 3.5 was used and rnn ¢ = (∣n - n¢∣rru + r0 ) is the e–h states.
distance as a function of the sites (n , n¢), with rru the repeat Given that the coupling constants λ and κ—which gen-
unit length and r0 the intrinsic e–h distance for a Frenkel erate charge-separated ∣CSn ñ states from either the interfacial
exciton. ∣XT1ñ state or the charge-transfer excitonic ∣CTX1,2ñ state—are
of comparable magnitude, competing transfer channels are
2.2. Electronic couplings expected to arise for the photogenerated ∣XTñ states: either a
direct transfer to the interfacial ∣CS1ñ state takes place, or else
Next, the electronic coupling part of the Hamiltonian com- OT charge transfer excitons ∣CTX1,2ñ are generated first, fol-
prises multiple transfer channels and reads as follows, lowed by a transfer to the ∣CS2 ñ state. The generation of
coupl charge-transfer excitons should be efficient due to the large
Hˆ = l (∣XT1ñáCS1∣ + h.c.) + k (∣CTX1,2ñáCS2 ∣ + h.c.)
couplings (k1, k 2 ) and favorable energetics. The pathway
⎛N -1 N ⎞
+ k1 ⎜ å ∣XTn ñáCTX n + 1, n∣ + å ∣XTn ñáCTX n - 1, n∣ + h.c.⎟ ∣XTñ ⟶ ∣CTX1,2ñ ⟶ ∣CS2 ñ bypasses the trapped inter-
⎝ n=1 n=2 ⎠ facial ∣CS1ñ state and could modify the overall efficiency of
⎛N -1 N ⎞ exciton dissociation at the interface.
+ k 2 ⎜ å ∣XTn ñáCTX n, n + 1∣ + å ∣XTn ñáCTX n, n - 1∣ + h.c.⎟ The electronic couplings k1 and k 2 that mediate transfer
⎝ n=1 n=2 ⎠
between the ∣XTñ and ∣CTXñ states are a sensitive function of
N N -1
+ j å (∣XTn ñáXTn + 1∣ + h.c.) + th the inter-chain distance, and take large values at short inter-
å (∣CSn ñáCSn +1∣ + h.c.)
n=1 n =1 chain distances (here, 3.5 Å) characteristic of regioregular
⎛N -1 N domains. By contrast, these couplings are expected to be

+ le ⎜ å å ∣CTX n, n ¢ñáCTX n + 1, n ¢∣ + h.c. strongly reduced in regiorandom phases where inter-chain
⎜ n = 1 n ¢= 1 distances tend to be larger. For comparison, we have calcu-
⎝ n ¢¹ n, n + 1

lated (k1, k 2 ) for several representative distances and found a
N N
⎟ very rapid decrease in magnitude. As compared with the
+ å å ∣CTX n, n ¢ñáCTX n - 1, n ¢∣ + h.c.⎟
n =2 n ¢= 1 ⎟ present values (k1= 0.36 eV, k 2 = 0.14 eV) for 3.5 Å (see
n ¢¹ n, n - 1 ⎠ figure 2), we obtained (k1= 0.17 eV, k 2 = 0.06 eV) for 4.0 Å,
⎛ N -1 N and (k1= 0.08 eV, k 2 = 0.02 eV) for 4.5 Å. While changes in

+ lh ⎜ å å ∣CTX n, n ¢ñáCTX n, n ¢+ 1∣ + h.c. the energetics and oligomer conformations also need to be
n=1
⎝n ¢= 1n ¹ n ¢ , n ¢+ 1 taken into account in the description of regiorandom phases, it
N -1 N
⎞ is clear from the reduced couplings at larger distances that

+ å å ∣CTX n, n ¢ñáCTX n - 1, n ¢∣ + h.c.⎟. charge-transfer excitons will play a much less prominent role.
n=2
n ¢= 1 ⎠
n ¹ n ¢ , n ¢+ 1 (3 )
2.3. Electron–phonon coupling
The first term relates to the coupling between the excitonic
and donor/acceptor charge transfer manifolds (l = 0.2 eV), The electron–phonon coupling part of the Hamiltonian is
which is restricted to the interfacial ∣XT1ñ and ∣CS1ñ states. All included within a linear vibronic coupling model, using

3
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Figure 1. Illustration of the oligothiophene-fullerene (P3HT-PCBM type) aggregate model of equations (1)–(3), where the fullerene acceptor
is considered as an effective super-particle. The Coulomb barrier that defines the on-site energies of the successive charge-separated states is
illustrated in the lower panel. Two types of barriers are considered, as explained in the text.

Table 1. Parameters for the coupling Hamiltonian equation (3), and for N ´ NOT oligothiophene modes (where N corre-
specified in eV. sponds to the number of OT units),
λ κ κ1 κ2 j th le lh
N NOT
OT 1
0.2 0.165 0.357 0.139 0.1 −0.12 0.08 0.08 Hˆ e- ph = å å 2 wlOT (( pˆnOT
, l ) + (qn, l ) )
2 ˆ OT 2
n =1 l=1
N NOT
+ å å cCS,
OT
l qˆn, l ∣CSn ñáCSn ∣
OT

fragment-based spectral densities [36] that were obtained n =1 l=1


N NOT
from state-specific Franck–Condon gradients, followed by a
truncated effective-mode expansion [18, 23, 35]. The elec- + å å cXT,
OT
l qˆn, l ∣XTnñáXTn∣
OT

n =1 l=1
tron–phonon coupling Hamiltonian is split into fullerene N N NOT
(C60) and oligothiophene (OT) contributions, + å å å (cXT,
OT
l qˆn, l + cXT, l qˆn ¢ , l )
OT OT OT

C60 OT n = 1 n ¢= 1 l = 1
Hˆ e- ph = Hˆ e- ph + Hˆ e- ph, (4 ) ´ ∣CTXn, n ¢ñáCTXn, n ¢∣. (6 )
where the Hamiltonian reads as follows in mass and fre-
Here, it is assumed that the OT electron–phonon couplings
quency weighted (dimensionless) coordinates, for NC60 full-
are identical for all fragments, such that the vibronic cou-
erene super-particle modes,
plings do not carry site indices n. Furthermore, it is assumed
C60
NC60
1 C60 C60 2 that the vibronic couplings for the ∣CTXn, n ¢ñ states are given,
Hˆ e- ph = å 2
wl (( pˆl ) + (qˆlC60)2 )
to a good approximation, by the sum of XT vibronic cou-
l=1
N NC60 plings of individual OT fragments (see [37]).
+ å å clC60 qˆlC60 ∣CSn ñáCSn ∣ (5 ) As in [18], we used NC60 = NOT = 8 effective modes per
n =1 l=1 fragment (hence, 112 modes overall). The relevant vibronic

4
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Figure 2. For illustration, the full set of relevant electronic configurations is shown for a simplified system comprising two donor sites. Upper
panel: excitonic (∣XTnñ), charge transfer exciton (∣CTXn, n ¢ñ) and charge separated (∣CSn ñ) configurations. Lower panel: on-site energies and
electronic couplings in the basis of these configurations.

couplings are tabulated in table 2. Low-frequency inter-frag- Table 2. Frequencies and vibronic couplings of effective modes
ment modes do not modify the dynamics significantly [17] relating to the electron–phonon coupling part of the Hamiltonian,
and were omitted from the present study. Ĥe - ph , see also supplementary material provided in [18]. All
quantities are specified in eV.
Overall, the Hamiltonian of equations (1)–(6) generalizes
the model of [18] such as to include OT charge transfer Mode
OT OT
excitons. The dynamics resulting from the above Hamiltonian index l wlC60 wlOT clC60 cCS, l cXT, l
reduces to the dynamics described in [18] when the couplings
1 0.2000 0.4012 0.0639 0.0099 0.0057
(k, k1, k 2 ) between XT and CTX manifolds, as well between 2 0.1842 0.3977 0.0929 −0.0001 0.0041
the CS and CTX manifolds are set to zero. 3 0.1778 0.1827 −0.0569 −0.0959 −0.1834
In the following, three different variants of the above 4 0.1411 0.1785 −0.0247 0.0815 0.0663
Hamiltonian will be considered which comprise, besides 13 5 0.0939 0.1345 0.0396 −0.0567 −0.0465
∣XTnñ states and 13 ∣CSn ñ states, (i) a single ∣CTXn, n ¢ñ state at 6 0.0799 0.1118 −0.0192 0.0165 0.0517
the interface, with n = 1, n¢ = 2 (i.e., 27 states overall), (ii) a 7 0.0558 0.0426 −0.0335 −0.0152 −0.0285
subset of ∣CTXn, n ¢ñ states, with ∣n - n¢∣ = 1, yielding a total 8 0.0332 0.0183 0.0139 −0.0174 −0.0109
of 50 electronic states, and (iii) the complete ∣CTXn, n ¢ñ
manifold, yielding a total of 182 electronic states. All systems
involve the full set of 112 vibrational modes.
27-state system involving the XT and CS manifolds along
with the ∣CTX1,2ñ state. The figure shows the eigenstates
3. Coupled XT, CTX, and CS manifolds (panels (a) and (c)) and the energetics (panels (b) and (d)), for
both barrier I and barrier II, evaluated at the Franck–Condon
Complementary to the site-local representation of the Hamil- geometry. The energetics of the manifolds is illustrated in a
tonian, it is instructive to analyze the properties of the elec- partially adiabatic representation (labeled I) where the elec-
el
tronic part, Hˆ = Hˆ
on ‐ site
+ Hˆ
coup
, in terms of its localized tronic coupling between the different manifolds has been set
and delocalized eigenstates. In particular, the mixing of the XT, to zero, and a full adiabatic representation (labeled II) where
el
CTX, and CS manifolds, and the role of trapped interfacial the complete electronic Hamiltonian Ĥ is diagonalized.
states, can be best appreciated from this perspective. From inspection of the energetics, it is clear that an extensive
To this end, figure 3 shows the energetic distribution and mixing of the XT, CTX and CS manifolds is taking place for
eigenstates of the simplest system under study, i.e., the both barriers.

5
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Figure 3. Electronic eigenstates and energetics for barriers I and II. Panels (a) and (c): for barrier I (panel (a)) and barrier II (panel (c)), the
electronic eigenstates of the 27-state model system are shown in ascending energetic order. The abscissa defines a series of electron–hole
basis states pertaining to the OT excitonic manifold (∣XTnñ, n = 1, ¼, 13), the OT charge transfer exciton (∣CTX1,2ñ), and the OT/C60 charge
separated states (∣CSn ñ, n = 1, ¼, 13). Panels (b) and (d): the energetics of the relevant states are shown, for a local adiabatic representation
(labeled I) where electronic couplings between the XT, CTX, and CS manifolds have been turned off and sub-blocks pertaining to the
different manifolds are diagonalized, and for the full adiabatic representation (labeled II) where all electronic couplings are included. The
bright exciton is highlighted by a red ellipse, and the CTX state is highlighted by a blue ellipse.

As can be immediately seen from the site-resolved discussion of [18], though, we expect several competing
eigenstate representation (panels (a) and (c)), the lowest- pathways, all of which involve participation of vibronic
energy states are localized at the OT/C60 interface and couplings. In particular, ultrafast internal conversion
involve contributions from the lowest-lying XT, CTX, and [18, 38, 39] within the XT manifold, channeling the popu-
CS states. At higher energies, a distribution of strongly lation to localized interfacial states, might happen faster than
delocalized states is present, which involves pronounced delocalization towards long-range charge-separated states.
mixing of contributions of the XT, CTX, and CS manifolds. In a complementary fashion, figure 4 illustrates the
For barrier I, a separate high-energy manifold is visible which energetics for the 50-state system which comprises the subset
essentially consists of CS states around the top of the Cou- of ∣CTXn, n ¢ñ states, with ∣n - n¢∣ = 1, as explained above.
lomb barrier. Overall, the picture is similar to the situation The participation of a large number of CTX states modifies
described in [18], except for the noticeable contribution of the the distribution of states substantially. Importantly, a set of
∣CTX1,2ñ state, especially to the low-lying states. The latter low-energy states appears that mainly exhibit a mixture of XT
states can play a crucial role since they give rise to interfacial and CTX character, with smaller participation of CS char-
trapping effects. acter. These states could act as long-lived traps within the OT
In figure 3 (panels (b) and (d)), the energetic positions of domain.
the bright exciton and the CTX state are highlighted. This In the following, a detailed quantum dynamical analysis
picture underscores that a substantial amount of excess energy will be reported to analyze which pathways will turn out to be
is present, such that immediate delocalization might take dominant, and to what extent the picture of [18] is modified in
place, involving participation of the CTX state. Following the the presence of the CTX manifold.

6
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Figure 4. Analogously to panels (b) and (d) of the preceding figure 3, the energetics is illustrated for the 50-state model, for barrier I (left) and
barrier II (right). As before, a local adiabatic representation (labeled I) is shown, where electronic couplings between the XT, CTX, and CS
manifolds have been turned off and sub-blocks pertaining to the different manifolds are diagonalized, along with the full adiabatic representation
(labeled II) where all electronic couplings are included. Conspicuously, a set of low-energy states appears which exhibit a mixture of XT, CTX,
and CS character. Indeed, it turns out in the case of barrier I that the lowest energetic state exhibits around 50% XT and CTX character,
respectively, and very little admixture of CS character. In the case of barrier II, the lowest energetic states exhibit a strong state mixing involving
XT, CTX, and CS character. The lowest state carrying predominant CS character is located at −0.05 eV (barrier I) and at −0.2 eV (barrier II).

4. Quantum dynamics of charge separation in the with expansion coefficients A j1 , ¼ , jf , mn (t ) º AJ , nm (t ) and time-
presence of CTX states dependent SPF configurations kf = 1 j(jk) (qk , t ) º FJ ({qk}, t ).
k
In the ML-MCTDH approach [28, 30], the time-dependent
Quantum dynamical calculations have been carried out using
the ML-MCTDH method [28–30], as briefly summarized in the SPFs are in turn expanded in sum-over-products form.
following. We then proceed to the discussion of various results, The wavefunction equation (7) describes correlated vibronic
for the minimal 27-state system including the ∣CTX1,2ñ state, as states involving coherent superpositions of multiple e–h
well as the 50-state and 182-state models described above. configurations.
For the three model systems addressed above, all of
4.1. ML-MCTDH simulations which include 112 vibrational (phonon) modes and a variable
number of electronic states (27, 50, and 182), a 6-layer
The ML-MCTDH method [28, 30] is a generalization of the representation of the wavefunction was used. Table 3 lists the
MCTDH technique [24–27] (Heidelberg Package [40]), i.e., a numbers of SPFs per layer. A primitive harmonic oscillator
tensor approximation scheme permitting an efficient propa- DVR representation was chosen for all vibrational modes
gation of high-dimensional molecular systems. The method comprising 32 DVR points per mode. Typically propagation
employs a hierarchical representation of the wavefunction in
times between 300 fs and 1 ps were used.
conjunction with variational equations of motions for time-
Initial conditions correspond either to individual elec-
dependent coefficients and so-called single particle functions
tron–hole states, or else a superposition of such states. In
(SPFs). The latter are in turn represented by a static basis of
particular, preparation of an initial localized ∣XT1ñ state is
discrete variable representation (DVR) type.
considered, as compared with a delocalized bright exciton
In the MCTDH approach [24–27], the time-dependent
wavefunction representing a set of vibrational modes initial condition constructed by imaginary time propagation
{q1, ¼, qf } and a set of e–h states ∣nmñ is expanded in terms of (‘relaxation’ procedure in MCTDH; here, the electronic sub-
Hartree products, i.e., products of time-dependent SPFs, system is decoupled from the vibrational modes3).

∣y (q1, ¼, qf , t )ñ
N n1 nf f
3
=å å ¼ å A j1 ,¼,jf ,nm (t )  j(jk) (qk , t )∣nmñ The initial bright delocalized exciton state of figures 5(b) and (d) is generated
k by changing the sign of the excitonic coupling, such that the excitonic ground
n,m j1 jf k= 1
state wavefunction for the corresponding J-aggregate species (with negative
N excitonic coupling) is produced. The latter bright-state wavefunction is
=å å AJ , nm (t ) FJ ({qk}, t )∣nmñ (7 ) subsequently used as an initial state for the real-time propagation for the
n,m J H-aggregate system.

7
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Figure 5. Time-evolving electron–hole state populations resulting from ML-MCTDH calculations, for barrier II. Exciton (XT) state
populations are shown in red, OT/C60 charge separated (CS) populations are shown in blue, and charge transfer exciton (CTX) populations
are shown in light blue. Panels (a) and (b): reference system in the absence of the CTX manifold, for a localized initial condition
corresponding to the interfacial ∣XT1ñ state (panel (a)) and for a delocalized, bright exciton initial condition (panel (b)). Panels (c) and
(d): corresponding time-evolving populations for 27-state simulations including the ∣CTX1,2ñ state. The presence of the CTX state is seen to
significantly modify the nature of the low-energetic trapped states which accumulate a significant part of the overall population. Interestingly,
the population of the ∣CS1ñ state, which is predominantly populated in the 26-state simulations shown in panels (a) and (b), is strongly reduced
in the 27-state simulations of panels (c) and (d). Instead, the ∣CTX1,2ñ state and the ∣CS2 ñ state acquire a non-negligible weight.

Table 3. Wavefunction partitioning for the ML-MCTDH calculations reported in this work. All simulations were carried out for 112
vibrational (phonon) modes.
Layer 1 2 3 4 5 6
27 states SPF [39, 39, 27] [35, 35] [29, 29] [19, 19] [12, 12, 7] [4, 4, 7]
50 states SPF [45, 45, 50] [39, 35] [31, 31] [32, 32] [7, 12, 10] [4, 7, 7]
182 states SPF [32, 32, 182] [29, 29] [24, 24] [15, 15] [7, 7, 7] [4, 4, 4]

4.2. Quantum dynamics of minimal model including ∣CTX1;2 〉 delocalized (bright exciton) initial condition (see panels (a)
state and (c), and panels (b) and (d), respectively).
In line with our previous studies [17, 18, 36], we expect a
Figure 5 compares the result of ML-MCTDH calculations for markedly nonexponential short-time dynamics, in agreement
the 27-state model comprising the ∣CTX1,2ñ state (panels (c) with experimentally observed ultrafast initial charge separation
and (d)) with the corresponding dynamics in the absence of on a time scale of t ~ 50–200 fs [41, 42]. The initial dynamics
the CTX state (panels (a) and (b)). All results illustrated in the involves pronounced vibronic effects that mediate the transfer
figure refer to barrier II. The figure shows site-resolved to the CS manifold. Vibronic effects also give rise to trapping
populations of the XT (red), CTX (light blue), and CS (blue) effects, due to the strong coupling of both inter-chain excitons
manifolds as a function of time, for both a localized and and charge-separated states to vibrations—in particular

8
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Figure 6. As an extension to the representation of figure 5, the time-dependent population of the ∣CS1ñ state is shown for the 26-state model
(in the absence of the CTX manifold) as compared with the 27-state model, the 50-state model, and the 182-state model, for barriers I and II
and for both localized and delocalized initial conditions. While the ∣CS1ñ state population is already reduced in the 27-state model, this effect
is found to be much more pronounced in the 50-state model and the 182-state model. Here, the ∣CS1ñ state population is essentially quenched.
The 50-state and 182-state models yield similar results such that the 50-state model—which includes charge-transfer excitons with an
electron–hole separation ∣n - n¢∣ = 1—can be taken to be a good approximation to the full dynamics.

high-frequency CC modes—whose ultrafast reorganization leads Participation of the ∣CTX1,2ñ state, as shown in panels
to energetically stabilized electronic distributions. These effects (c) and (d), is seen to lead to a tangible reduction of the ∣CS1ñ
occur on the same time scale (t ~ 20–100 fs) and cannot be population, while the ∣CS2 ñ population rises more rapidly.
separated from the charge separation dynamics in the present This is due to the fact that the ∣CTX1,2ñ state directly couples
system. Also, intra-chain trapping as revealed by transient to the ∣CS2 ñ state, as explained above. The amount of long-
anisotropy measurements [43] is not fully resolved in the present range CS states is comparable to the reference calculations of
study since we define e–h states in terms of ‘coarse-grained’ panels (a) and (b). These observations are in line with the
oligomer units (for comparison, e–h states can be alternatively electronic eigenstate picture of the preceding section where a
strong mixing of XT, CTX and CS character in the low-
defined in a monomer basis [44], permitting a full analysis of
energy interfacial states was observed.
intra-chain trapping).
Referring to the reference calculations of panels (a) and
(b), we note that (i) the localized initial condition (panel (a)) 4.3. Quantum dynamics of 50-state and 182-state models
immediately generates a significant population of the inter-
We now turn to the larger model systems, including 50 or 182
facial ∣CS1ñ state, on a time scale of t ~ 30 fs, but (ii) a non-
electronic states. Since the site-resolved representation of the
negligible participation of delocalized CS states is observed
propagation results tends to be cumbersome, we focus on
early on as well. A somewhat lower participation of deloca- several integrated properties, notably (i) the population of the
lized CS states is seen for the delocalized, bright-state initial interfacial ∣CS1ñ state and (ii) the free carrier population,
condition, where the rise of the interfacial ∣CS1ñ state is found which is calculated as the integrated population over CS states
to follow a slower build-up kinetics, with a t ~ 500 fs time with an electron–hole separation ∣n - n¢∣ > 8, similarly to the
scale. This build-up can be interpreted in terms of an internal procedure of [17, 18].
conversion process [18, 38, 39] which leads to the gradual Figure 6 shows the time-dependent population of the
population of low-energy states within the excitonic manifold. interfacial charge separated ∣CS1ñ state, for the three types of
At the same time, long-range charge-separated states emerge simulations described above (27, 50, and 182 states), as
and build up steadily. These observations are consistent with compared to reference simulations in the absence of the CTX
the result that a lower barrier favors ultrafast long-range manifold. Both localized initial conditions (∣XT1ñ) and delo-
charge separation as discussed in [17, 18]. calized initial conditions (bright exciton) are considered, for

9
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

Figure 7. In a complementary fashion to figure 6, the time-evolving free carrier populations are shown, for the 26-state model (in the absence
of the CTX manifold) as compared with the 27-state model, the 50-state model, and the 182-state model, for barriers I and II and for both
localized and delocalized initial conditions. Free carrier populations are obtained as integrated populations for the charge separated states
∣CSn ñ, n = 8, ¼, 13. While free carrier populations are not strongly affected by addition of the ∣CTX1,2ñ state in the 27-state model, they are
much reduced in the 50-state and 182-state models.

both types of barriers. Barrier I generates the highest popu- yield, both in the case of localized and delocalized initial
lation of the ∣CS1ñ state, with an immediate onset in the case of conditions. Again, results for the 50-state and 182-state
the localized (∣XT1ñ) initial condition and a gradual rise for the models are very similar.
bright-state initial condition, in line with the dynamics
depicted in figure 5. For barrier II, the ∣CS1ñ state population
follows the same trends, but is significantly reduced. 5. Discussion and conclusions
In the case of a single ∣CTX1,2ñ state, the ∣CS1ñ state
population is reduced to a significant extent, due to the The above observations lead us to conclude that the presence
alternative ∣XT1ñ  ∣CTX1,2ñ  ∣CS2 ñ pathway that was dis- of the CTX manifold, representing inter-chain charge-sepa-
cussed above (see section 2). Adding more CTX states, in the rated polaronic species in the regioregular oligothiophene
50-state calculations (comprising all CTX states with an domain, reduces the role of the interfacial charge transfer state
electron–hole separation ∣n - n¢∣ = 1) and in the 182-state (∣CS1ñ); in fact, a dynamical scenario results that entirely
calculations (comprising all CTX states), leads to an almost bypasses the ∣CS1ñ state. However, at the same time the CTX
complete quenching of the ∣CS1ñ state population. This is the manifold leads to efficient trapping of electron–hole pairs of
case for both types of barriers. Clearly the pathway via the mixed XT/CTX character in the oligothiophene donor
CTX manifold becomes dominant. Also, the simulations for domain. Hence, despite the availability of additional CTX-CS
50 states versus 182 states are found to give nearly equal coupling pathways, the CTX mediated channel entails a net
results, such that one may conclude that CTX states with reduction of the free carrier population.
∣n - n¢∣ = 1 dominate the picture. One should note, though, that these conclusions depend
Figure 7 shows the time-dependent free-carrier popula- in a sensitive fashion on the energetics and electronic cou-
tion, i.e., the integrated ∣CSn ñ state populations for n > 8, for plings between the different manifolds. For example, for
the same simulations as in figure 6. In line with our previous electronic couplings between the XT and CTX manifolds
findings [17, 18], the free carrier yield is tangibly larger for reduced by a factor of two (k1 = 0.18 eV, k 2 = 0.07 eV), the
barrier II as compared with barrier I. While the 27-state free carrier yield is found to be close to the reference calcu-
simulations with a single ∣CTX1,2ñ state lead to a comparable lations in the absence of CTX states. It is conceivable that
free carrier yield as the reference calculations, the simulations further modification of the energetics, e.g., involving the XT-
for 50 and 182 states show a strongly reduced free carrier CS energetic offset, could even lead to an enhancement of

10
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

free carrier yields due to the CTX pathway. In the actual [4] Gélinas S, Rao A, Kumar A, Smith S L, Chin A W, Clark J,
material, static and dynamic disorder are expected to con- van der Poll T S, Bazan G C and Friend R H 2014 Science
tribute to fluctuations in both on-site energies and electronic 343 512
[5] Grancini G, Maiuri M, Fazzil D, Petrozzal A, Egelhaaf H,
couplings. Brida D, Cerullo G and Lanzani G 2013 Nat. Mater. 4 1602
From the perspective of quantum dynamical simula- [6] Braun C L 1984 J. Chem. Phys. 80 4157
tions, the use of the ML-MCTDH technique provides a [7] Bakulin A A, Rao A, Pavelyev V G, van Loosdrecht P,
unique tool to accurately describe quantum dynamical Pshenichnikov M S, Niedzialek D, Cornil J, Beljonne D and
evolution of hundreds of electronic states and vibrational Friend R H 2012 Science 335 1340–4
[8] Jailaubekov A E et al 2013 Nat. Mater. 12 66–73
modes. Importantly, the full quantum dynamical description [9] Caruso D and Troisi A 2012 Proc. Natl Acad. Sci. 109
permits the inclusion of vibronic mechanisms like internal 13498–502
conversion [18, 38, 39] which play a critical role in med- [10] Vázquez H and Troisi A 2013 Phys. Rev. B 88 205304
iating electronic deexcitation on ultrafast time scales. Fur- [11] Zheng Z, Tummala N R, Fu Y, Coropceanu V and Brédas J
ther, a quantum dynamical description is needed to account 2017 ACS Appl. Mater. Interfaces 9 18095
[12] Bakulin A A, Dimitrov S D, Rao A, Chow P C Y, Nielsen C B,
for the complex patterns of delocalization and trapping that Schroeder B C, McCulloch I, Bakker H J, Durrant J R and
evolve on ultrafast time scales, which cannot be captured by Friend R H 2012 J. Phys. Chem. Lett. 4 209
kinetic models. While the present calculations were carried [13] Vandewal K et al 2014 Nat. Mater. 13 63
out at zero temperature—which is appropriate to describe [14] D’Avino G, Mothy S, Muccioli L, Zannoni C, Wang L,
the ultrafast part of the photoinduced dynamics—temper- Cornil J, Beljonne D and Castet F 2013 J. Phys. Chem. C
117 12981
ature effects will be included in future studies, along with [15] D’Avino G, Muccioli L, Olivier Y and Beljonne D 2016
energetic and structural disorder. J. Phys. Chem. Lett. 7 536
To summarize, the inclusion of the CTX manifold con- [16] Hood S N and Kassal I 2016 J. Phys. Chem. Lett. 7 4495
siderably enhances the complexity of transfer pathways at [17] Tamura H and Burghardt I 2013 J. Am. Chem. Soc. 135
P3HT-PCBM type junctions, as a result of the coupling 16364–7
[18] Huix-Rotllant M, Tamura H and Burghardt I 2015 J. Phys.
between Frenkel-type excitonic manifolds, charge transfer Chem. Lett. 6 1702
excitons, and donor/acceptor CS species. Given that recent [19] Polkehn M, Eisenbrandt P, Tamura H and Burghardt I 2017
experimental observations [20–22], in agreement with our Int. J. Quantum Chem. (https://doi.org/10.1002/
calculations and related electronic structure work [45], are qua.25502)
strongly indicative of pronounced XT/CTX mixing effects in [20] Reid O G, Pensack R D, Song Y, Scholes G D and Rumbles G
2013 Chem. Mater. 26 561
ordered polythiophene materials, it is important to include [21] Magnanelli T J and Bragg A E 2015 J. Phys. Chem. Lett. 6 438
these effects in the modeling of donor–acceptor junctions [22] De Sio A et al 2016 Nat. Commun. 7 13742
involving regioregular thiophene-based domains. [23] Tamura H, Hughes K H, Martinazzo R, Wahl J, Binder R and
Burghardt I 2017 Ultrafast energy and charge transfer in
functional molecular nanoscale aggregates Ultrafast
Dynamics at the Nanoscale: Biomolecules and
Acknowledgments Supramolecular Assemblies ed I Burghardt and S Haacke
(Singapore: Pan Stanford Publishers) ch 7, p 257
Funding by the Deutsche Forschungsgemeinschaft (DFG) and [24] Beck M H, Jäckle A, Worth G A and Meyer H D 2000 Phys.
Rep. 324 1
the Agence Nationale de la Recherche (ANR) in the frame- [25] Worth G A, Meyer H, Köppel H, Cederbaum L S and
work of the project MolNanoMat (BU-1032-2), as well as Burghardt I 2008 Int. Rev. Phys. Chem. 27 569
support by the NAKAMA funds are gratefully acknowledged. [26] Meyer H D, Manthe U and Cederbaum L S 1990 Chem. Phys.
Further, we are grateful to Konstantin Falahati for his assis- Lett. 165 73
tance with the preparation of one of the figures. [27] Manthe U, Meyer H D and Cederbaum L S 1992 J. Chem.
Phys. 97 3199
[28] Wang H and Thoss M 2003 J. Chem. Phys. 119 1289
[29] Manthe U 2008 J. Chem. Phys. 128 164116
[30] Vendrell O and Meyer H 2011 J. Chem. Phys. 134 044135
ORCID iDs [31] Jiang X M, Österbacka R, Korovyanko O, An C P, Horovitz B,
Janssen R and Vardeny Z V 2002 Adv. Funct. Mater. 12
I Burghardt https://orcid.org/0000-0002-9727-9049 587–97
[32] Spano F C and Silva C 2014 Annu. Rev. Phys. Chem. 65
477–500
[33] Tamura H and Burghardt I 2013 J. Phys. Chem. C 117
References 15020–5
[34] Tamura H 2016 J. Phys. Chem. A 120 9341
[35] Tamura H, Burghardt I and Tsukada M 2011 J. Phys. Chem. C
[1] Sariciftci N S, Smilowitz L, Heeger A J and Wudl F 1992 115 10205–10
Science 258 1474 [36] Tamura H, Martinazzo R, Ruckenbauer M and Burghardt I
[2] Deibel C, Strobel T and Dyakonov V 2010 Adv. Mater. 22 2012 J. Chem. Phys. 137 22A540
4097–111 [37] Polkehn M, Popp W, Tamura H and Burghardt I 2017 Ultrafast
[3] Ryno S M, Ravva M K, Chen X, Li H and Brédas J 2017 Adv. dynamics of polaron formation in neat oligothiophene
Energy Mater. 7 1601370 aggregates in preparation

11
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 014003 M Polkehn et al

[38] Nitzan A 2006 Chemical Dynamics in Condensed Phases: [41] Brabec C J, Zerza G, Cerullo G, De Silvestri S, Luzzati S,
Relaxation, Transfer, and Reactions in Condensed Hummelen J C and Sariciftci N S 2001 Chem. Phys. Lett.
Molecular Systems (Oxford: Oxford University Press) 340 232
[39] Jumper C C, Anna J M, Stradomska A, Schins J, [42] Pensack R D, Banyas K M and Asbury J B 2010 J. Phys.
Myahkostupov M, Prusakova V, Oblinsky D G, Chem. B 114 12242
Castellano F N, Knöster J and Scholes G D 2014 Chem. [43] Grage M, Zaushitsyn Y, Yartsev A, Chachisvilis M,
Phys. Lett. 599 23 Sundström V and Pullerits T 2003 Phys. Rev. B 67 205207
[40] Worth G A, Beck M H, Jäckle A and Meyer H 2014 The [44] Binder R, Polkehn M, Ma T and Burghardt I 2017 Chem. Phys.
MCTDH Package, Version 9 (2007) 482 16
H D Meyer Version 8.5 (2014), see http://pci.uni-heidelberg. [45] Li H, Nieman R, Aquino A, Lischka H and Tretiak S 2014
de/tc/usr/mctdh/ J. Chem. Theory Comput. 10 3280

12

You might also like