Kinetic Monte Carlo study of sensitiviy of OLED efficiency and
lifetime to materials parameters
Citation for published version (APA):
Coehoorn, R., Eersel, van, H., Bobbert, P. A., & Janssen, R. A. J. (2015). Kinetic Monte Carlo study of sensitiviy
of OLED efficiency and lifetime to materials parameters.
Advanced Functional Materials
,
25
(13), 2024-2037.
https://doi.org/10.1002/adfm.201402532
Document license:
TAVERNE
DOI:
10.1002/adfm.201402532
Document status and date:
Published: 01/01/2015
Document Version:
Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)
Please check the document version of this publication:
• A submitted manuscript is the version of the article upon submission and before peer-review. There can be
important differences between the submitted version and the official published version of record. People
interested in the research are advised to contact the author for the final version of the publication, or visit the
DOI to the publisher's website.
• The final author version and the galley proof are versions of the publication after peer review.
• The final published version features the final layout of the paper including the volume, issue and page
numbers.
Link to publication
General rights
Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners
and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.
• Users may download and print one copy of any publication from the public portal for the purpose of private study or research.
• You may not further distribute the material or use it for any profit-making activity or commercial gain
• You may freely distribute the URL identifying the publication in the public portal.
If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please
follow below link for the End User Agreement:
www.tue.nl/taverne
Take down policy
If you believe that this document breaches copyright please contact us at:
providing details and we will investigate your claim.
Download date: 28. Aug. 2024
FULL PAPER
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
2024
wileyonlinelibrary.com
and excitonic processes and determining
the values of the relevant material-specifi c
parameters, ii) supporting the layer stack
design process by providing an effi cient
route for developing understanding of the
benefi ts and disadvantages of the use of
novel materials and layer stack concepts.
The third (vertical) axis in the fi gure indi-
cates examples of the performance charac-
teristics which may be studied using the
simulations. We envisage that simulations
can be used to clarify the sensitivity of the
OLED device performance to materials
and layer stack parameters, and can pro-
vide insights in the functioning of OLEDs
beyond the capability of experiments. For
example by a much more precise con-
trol of the device structure, by providing
insight in all processes at an ultimate
(molecular-scale) spatial resolution and at
the shortest relevant time scales, and by
providing the opportunity to explore the
performance under conditions which are
experimentally not realizable. A long-term goal of our research
is therefore to develop and apply a simulation method which
can address any combination of the three application types,
and which provides a platform within which novel insights
as obtained in each of the fi elds indicated in the fi gure can be
readily adopted.
So far, commonly used OLED simulations methods are based
on one-dimensional drift-diffusion approaches (see Coehoorn
and Bobbert and references therein)
[ 2 ]
in which the charge
transport and excitonic processes are included in a semi-empir-
ical manner. However, it is well-known that as a result of the
energetic disorder, resulting from the structural disorder, the
current density in OLEDs is fi lamentary, rather than uniform.
[ 3 ]
We have shown from molecular-scale three-dimensional kinetic
Monte Carlo (3D-KMC) modelling that, as a consequence, the
emission takes place on certain preferred molecular sites, rather
than being laterally uniform.
[ 4 ]
It is thus necessary to develop a
full 3D-KMC model for accurately simulating the effect of the
loss processes, such as triplet-polaron quenching and triplet-tri-
plet annihilation in phosphorescent OLEDs. These losses lead
to an effi ciency decrease with increasing current density and
luminance (roll-off of the internal quantum effi ciency,
η
IQE
).
We have recently extended the 3D-KMC method to include
the charge transport and all excitonic processes in an inte-
gral manner.
[ 5 ]
The simulations were found to provide a good
description of the roll-off of archetypical green and red emitting
OLEDs studied intensively in the literature,
[ 6 ]
and were used to
Kinetic Monte Carlo Study of the Sensitivity of OLED
Effi ciency and Lifetime to Materials Parameters
Reinder Coehoorn , * Harm van Eersel , Peter A. Bobbert , and René A. J. Janssen
The performance of organic light-emitting diodes (OLEDs) is determined by
a complex interplay of the optoelectronic processes in the active layer stack.
In order to enable simulation-assisted layer stack development, a three-
dimensional kinetic Monte Carlo OLED simulation method which includes
the charge transport and all excitonic processes is developed. In this paper,
the results are presented of simulations including degradation processes in
idealized but realizable phosphorescent OLEDs. Degradation is treated as a
result of the conversion of emitter molecules to non-emissive sites upon a
triplet-polaron quenching (TPQ) process. Under the assumptions made, TPQ
provides the dominant contribution to the roll-off. There is therefore a strong
relationship between the roll-off and the lifetime. This is quantifi ed using a
“uniform density model”, within which the charge carrier and exciton densi-
ties are assumed to be uniform across the emissive layer. The simulations
give rise to design rules regarding the energy levels, and are used to study the
sensitivity of the roll-off and lifetime to various other materials parameters,
including the mobility, the phosphorescent dye concentration, the triplet
exciton emissive lifetime and binding energy, and the type of TPQ process.
DOI: 10.1002/adfm.201402532
Prof. R. Coehoorn, H. van Eersel
Philips Research Laboratories
High Tech Campus 4, 5656AE , Eindhoven
The Netherlands
E-mail: reinder[email protected]
Prof. R. Coehoorn, H. van Eersel,
Dr. P. A. Bobbert, Prof. R. A. J. Janssen
Department of Applied Physics
Eindhoven University of Technology, P.O. Box 513
5600MB , Eindhoven , The Netherlands
1. Introduction
Organic light-emitting diode (OLED) technology for large area
lighting is presently developing from design-focused special-
lighting applications towards general-lighting applications for
homes and offi ces. Future products are expected to show added
functionality such as fl exibility, transparency in the off-state,
and color-tunability. Continuous improvements of the system
power effi ciency and long-term stability arise from the use of
novel organic semiconductor materials, novel stack architectures
and methods for improved light outcoupling.
[ 1 ]
The development of OLED active layer stacks is supported
by the use of simulation methods. As indicated schematically
in Figure 1 , two types of applications may be distinguished:
i) revealing the mechanism of the relevant charge transport
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER
2025
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
develop a design rule for an optimal power effi ciency, resulting
from a small
η
IQE
-roll-off and a low voltage.
In this paper, we demonstrate how 3D-KMC simulations
can also be used to predict the dependence of the lifetime
of phosphorescent OLEDs on the materials and layer stack
parameters. For that purpose, we have included in the simula-
tions irreversible changes of the molecular sites, according to a
chosen scenario, and simulate the resulting time-dependence
of the emission. Whereas a wide variety of degradation mecha-
nisms can be operative,
[ 7,8 ]
recent work has indicated that in
phosphorescent OLEDs triplet-polaron quenching followed by
defect formation can play a major role.
[ 9,10 ]
We investigate two
degradation scenarios in which upon triplet-polaron quenching
the dye molecule which is involved is with a certain probability
converted in a non-emissive molecule. These explicit 3D-KMC
lifetime simulations can be computationally intensive. We
have therefore also studied to what extent it is already possible
to predict lifetime trends using the loss due to exciton-polaron
quenching as obtained from results of the steady-state simula-
tions. As a fi rst step towards that goal, we focus on the case of
monochrome OLEDs with a single emissive layer (EML), and
show how under the simplifying assumption that the charge
carrier density and exciton generation are uniform across the
thickness of the EML the roll-off can be expressed in terms of
the materials and layer stack parameters. The approach is a
refi nement of earlier work,
[ 11,12 ]
by including the charge car-
rier density dependence of the mobility and by including a
mechanistic expression for the triplet-polaron quenching rate.
As a second step, it is shown how using this uniform den-
sity model (UDM) variations of the lifetime with layer stack
parameters, such as the EML thickness or dye concentration,
can then be predicted from the roll-off. Within the degrada-
tion scenarios considered, the absolute lifetime is proportional
to the inverse of the probability that upon a quenching event
degradation takes place. That is the only additional parameter
which is required. Within the framework of our approach, it
can be determined by one or a few “calibration” experiments.
We stress that the UDM only provides a starting point for
analyses of roll-off and lifetime. The simulations show that
changes of the shape of the charge carrier density and emis-
sion profi les with the voltage and current density also play an
important role.
The paper is organized as follows. Section 2 contains an
overview of the simulation method and a simulation example.
In Section 3, the uniform density model is developed, and
analytical expressions are given for the dependence of the
roll-off and lifetime on the material parameters, neglecting
exciton dissociation and exciton diffusion. In Sections 4 and
5, the sensitivity of the roll-off and lifetime, respectively, to
various materials parameters is studied from 3D-KMC simu-
lations. Two design rules for the most optimal energy level
structures are established, and the sensitivity of the roll-off to
the emissive lifetime of the dye molecules and to the triplet
binding energy is examined. The UDM is shown to provide a
fair prediction of the OLED lifetime for small dye concentra-
tions, for which exciton diffusion to already degraded sites
may be neglected. Section 6 contains a summary, conclusions
and outlook.
2. Simulation Method
In this section, fi rstly an overview is given of the 3D-KMC sim-
ulation method and of the simulation parameters used in this
paper (subsection 2.1). Subsequently, more detailed descrip-
tions are given on the treatment of charge transport, exci-
tonic processes and degradation (subsections 2.2, 2.3 and 2.4,
respectively). Finally, as an example the simulations are applied
in subsection 2.5 to a symmetric three-layer OLED, for which
the roll-off curve, emission profi le, carrier density profi le and
time-dependent decrease of the emission due to degradation
is shown. In Section 4, the energy level structure assumed for
this example will be shown to quite optimally fulfi ll two design
rules leading to a small roll-off, and in Section 5, the sensitivity
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Figure 1. Application areas of OLED device simulations (horizontal and diagonal axes) and examples of the performance characteristics which may be
studied (vertical axis). TADF denotes Thermally Assisted Delayed Fluorescence.
[ 7,8 ]
Simulation studies for a variety of performance characteristics, layer
stack designs and materials can help to develop an improved material-specifi c understanding of the relevant physical processes. Realizing progress in
that understanding is needed to make simulation-assisted OLED stack development more powerful.
FULL PAPER
2026
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
of the OLED lifetime when varying the Ir-dye concentration in
this OLED will be discussed.
2.1. Simulation Parameters—Overview
Within the 3D-KMC simulation method developed, the
molecules are treated as point sites, without any internal struc-
ture. The sites are placed on a simple cubic grid with a lattice
parameter a = N
t
1/3
, with N
t
the average molecular site density.
Any layer sequence and mixing of different types of molecules
in each layer can be defi ned, such as random mixing, aggre-
gation of similar molecules,
[ 13 ]
mixed interface layers, or, for
example, a concentration gradient in the emissive layer.
[ 14,15 ]
At
any moment in time, a site in the box is either empty, or occu-
pied by one electron, one hole, or by one electron and one hole.
When at a site a compound species (exciton, one electron and
one hole) is present, the exciton is further specifi ed as either
a singlet or a triplet exciton. Typically, for each set of external
conditions (voltage, temperature) fi ve parallel simulations are
carried out for nominally equal boxes, but with different dis-
order realizations, with a lateral area of 50 × 50 sites. From the
time-resolved and time-averaged results as obtained for these
simulation boxes, the statistical uncertainty can be judged.
The simulations are specifi ed by four sets of parameters:
i) parameters defi ning the system structure (thicknesses
and composition of all sub-layers), ii) electrical parameters,
iii) excitonic parameters, iv) the parameters specifying the
external conditions (voltage, V , and temperature, T ). The
charge transport and excitonic processes are included in a
manner as described by Mesta et al.
[ 4 ]
and by van Eersel et al.
[ 5 ]
A brief description is given in the next subsections. Table 1
gives an overview of the basic electrical and excitonic param-
eters, and their values used as a default in the simulations dis-
cussed in this paper. We note that the code allows for various
refi nements, not mentioned further here, such as non-isotropic
mobilities, non-Gaussian shapes of the density of states, and
long-distance (e.g., Förster-type) exciton-polaron quenching
and exciton-exciton annihilation. Furthermore, it allows for
carrying out simulations under special conditions, such as
steady-state or time-resolved photoluminescence
[ 11,16,17 ]
or
electroluminescence conditions.
[ 18 ]
All simulations were done
for T = 298 K.
The accessible material and device parameter range is
determined by practical (real time, CPU time, memory) limi-
tations. The total required real simulation time depends on
the level of time-averaged, spatially resolved or time-resolved
detail needed and the number of nominally equal OLEDs with
different disorder realizations run in parallel. When consid-
ering fi ve 50 nm × 50 nm of such OLEDs in parallel, the IQE
roll-off curve simulations discussed in this paper are obtained
typically in 310 days on a 3.0 GHz core [Intel Xeon X5675],
depending on the current density. Simulations in which the
fraction of quenched excitons or the fraction of radiatively
decaying excitons is small (at a very small current or very large
current density, respectively) are most demanding, since a
large number of events needs to be recorded to gather enough
statistics on the most rare event. The lifetime results discussed
in this paper are based on simulations over periods ranging
from a few days up to two months. Achieving suffi cient sta-
tistics is slowed down when considering larger device dimen-
sions, larger energy barriers at interfaces, enlarged charge
carrier or exciton hop distances (due to enlarged wavefunction
decay lengths or Förster radii, at large fi elds, or in very dilute
systems), enlarged density of states widths or trap depths, and
enlarged exciton lifetimes (so that at a given time more exci-
tons are present in the device).
2.2. Charge Transport
The hop rate
ν
of electrons and holes over a distance R and
from a molecule with energy E
i
(initial state energy) to a mole-
cule with E
f
= E
i
+ Δ E (fi nal state energy) is assumed to be given
by the Miller-Abrahams formalism. When the hopping occurs
between two molecules of types A and B the rate is assumed to
be given by:
exp 2 exp
2
1,A 1,B
B
Ra
EE
kT
ννν α
[]
()
=−
Δ−Δ
(1)
where
ν
1, X
is the hopping attempt rate between two molecules
of type X at a distance a (fi rst neighbor), for the case Δ E = 0,
and
α
1/
λ
is the inverse of the wavefunction decay length (
λ
).
The square-root form of the prefactor used arises when consid-
ering the transfer integral due to the overlap of the two expo-
nentially decaying wavefunctions. The expression applies to
any pair of molecules, also to molecules in different layers. So
no special parameters are introduced to describe the hop rates
across internal interfaces. In this paper, materials with spatially
uncorrelated Gaussian disorder are considered. The hole and
electron site energies are taken from a random Gaussian
distribution with a material-specifi c mean value and with a
material-specifi c standard deviation
σ
h
or
σ
e
, respectively,
modifi ed by i) the Coulomb interaction energy with the other
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Table 1. Overview of the electrical and excitonic parameters used in the
3D-KMC simulations, and the default values used in this paper.
Parameter Description Value
ν
1
Hopping attempt rate to the fi rst neighbor
3.3 × 10
10
s
1
σ
Width (standard deviation) of the Gaussian
density of states
0.10 eV
N
t
Site density
10
27
m
3
λ
Wavefunction decay length 0.3 nm
ε
r
Relative dielectric constant 3.0
E
T, b
Triplet exciton binding energy 1.0 eV
k
D,0
Dexter transfer prefactor
1.6 × 10
10
s
1
R
F,diff
Förster radius for exciton diffusion between
dye molecules
1.5 nm
Γ
rad
Radiative decay rate dye molecules
a)
0.544 µs
1
Γ
nr
Non-radiative decay rate dye molecules
a)
0.181 µs
1
a)
Values as obtained experimentally for the orange-red emitter Ir(MDQ)
2
(acac) in
an
α
-NPD matrix, assuming a radiative decay effi ciency
η
rad
= 0.75, intermediate
between the values obtained from analyses assuming randomly oriented or mainly
parallel emission dipole moments.
[ 19 ]
FULL PAPER
2027
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
charge-carriers and with its image charges in the electrodes
and by ii) the electrostatic potential difference due to the fi eld
arising from the applied voltage. As in the transport processes
studied holes or electrons are removed or added, the material-
specifi c mean energy level values for holes and electrons
should be termed properly as a mean ionization potential or
a mean electron affi nity, respectively. However, we indicate
these energies instead as a highest occupied molecular orbital
(HOMO) energy, E
HOMO
, or a lowest unoccupied molecular
orbital (LUMO) energy, E
LUMO
, respectively, as is common prac-
tice in the literature.
By using this mechanistic approach, we circumvent the
need to describe the charge transport by means of complicated
expressions for the mobility and diffusion coeffi cient, which
are used in one-dimensional drift-diffusion simulations. The
mobility, e.g., depends on the charge carrier density, electric
eld and temperature, and is affected by the shape of the den-
sity of states, the type of disorder, Coulomb interaction effects
in accumulation regions near interfaces, and the presence of
guest or trap molecules.
[ 2 ]
Furthermore, the mobility may effec-
tively be time (or frequency) dependent due to charge carrier
relaxation
[ 20,21 ]
and it is effectively layer thickness dependent
below a certain critical length scale.
[ 22 ]
In a uniform layer, the mobility at any fi eld, temperature and
carrier density is determined by four material parameters: a ,
σ
,
ν
1
and
λ
. These parameters can be determined from dedicated
transport studies of single-carrier single-layer devices, as shown
for several polymeric
[ 23–25 ]
and small-molecule materials.
[ 26,27 ]
The lattice parameter a which is obtained from such studies, is
quite similar to the value expected from the experimental molec-
ular site density. In this paper, we take a = 1 nm,
σ
= 0.1 eV,
and
ν
1
= 3.3 × 10
10
s
1
, which are typical values for small-mol-
ecule OLED materials (see van Eersel et al.
[ 5 ]
for a motivation
of the choice for
ν
1
). The wavefunction decay length
λ
is only
relevant in dilute host-guest systems in which guest-guest
hopping over distances larger than a is signifi cant. From such
studies, we deduce that
λ
is larger than the value of 0.1 × a (i.e.,
0.1 nm for a = 1 nm) which has been used in earlier modeling
studies of transport in one-component materials.
[ 4,28,29 ]
In this
paper, we take
λ
= 0.3 nm, the value used by van Eersel et al.
to describe the roll-off in Ir(ppy)
3
and PtOEP based OLEDs.
[ 5 ]
Within a more refi ned approach, the hop rates should be
described using Marcus theory.
[ 30 ]
However, as an additional
parameter would be involved (the reorganization energy), to
which the shape of the current voltage-curves will be quite
insensitive,
[ 31 ]
we have not adopted that approach. Based on
similar arguments (see also Cottaar et al.),
[ 31 ]
a simple cubic
grid is used in all cases instead of treating the grid structure as
a degree of freedom.
2.3. Excitonic Processes
Exciton generation (dissociation) is treated as a result of a hop of
a charge carrier to (away from) a site at which already a carrier of
the opposite polarity resides, using Equation (1) , albeit that Δ E
includes now a gain (loss) equal to the on-site singlet or triplet
exciton binding energy, E
S(T),b
. This takes into account that the
actual on-site exciton binding energy is fi nite, whereas within
the point-site model used the on-site electron-hole Coulomb
interaction would be infi nite. In this paper, we focus on phos-
phorescent OLEDs and take the triplet binding energy equal for
all sites. Exciton generation and dissociation are thus a natural
consequence of the electron-hole attraction; no additional rate
coeffi cients are introduced. The Coulomb interaction energy
between an electron and hole on a nearest neighbor site
(a charge-transfer exciton) is within our formalism equal to
E
CT,b
= e /(4
π
ε
0
ε
r
a ), with e the electron charge,
ε
0
the vacuum per-
mittivity and
ε
r
the relative dielectric constant (taken equal to 3),
that is, 0.48 eV. In a homogeneous material, a signifi cant dis-
sociation rate will thus be obtained unless the exciton binding
energy is well above E
CT,b
. In Section 4, we study the sensitivity
of the roll-off to E
T, b
. In this paper, we assume that all excitons
which are generated are triplets. As we consider only devices in
which almost all the excitons are formed on the phosphorescent
dye molecules, so that singlet-to-triplet conversion due to inter-
system crossing is fast, that is an excellent approximation.
Radiative and non-radiative decay are assumed to be mono-
exponential processes with rates Γ
rad
and Γ
nr
, respectively,
which may be deduced from the measured total decay rate
Γ = Γ
rad
+ Γ
nr
τ
1
(with
τ
the effective decay time) and the
radiative decay effi ciency
η
rad
(which may be taken equal
to the photoluminescence effi ciency) using the expressions
Γ
rad
=
η
rad
Γ and Γ
nr
= (1 –
η
rad
) Γ.
The exciton transfer rate from a site of type A to a site of type
B, at a distance R , is described as a sum of contributions due to
Förster transfer, with a rate
F,AB
F,diff ,AB
6
k
R
R
(2)
with
R
F,diff,AB
the Förster radius for exciton diffusion between
the two types (A, B) of molecules involved, and a Dexter contri-
bution given by
exp 2 exp
2
D D,0A D,0B
B
kkk R
EE
kT
α
[]
=−
Δ+Δ
(3)
where
k
D,0,A(B)
are material-type specifi c prefactors,
α
is the
inverse of the wavefunction decay length, and Δ E is the dif-
ference between the fi nal and initial state exciton energies.
Such an energy difference can occur when the two mol-
ecule types are different, or as a result of excitonic disorder.
In this paper, excitonic disorder is neglected. Only few
studies on triplet exciton diffusion in phosphorescent host-
guest systems have been carried out. Based on the work of
Kawamura et al.,
[ 32 ]
we take R
F,diff
= 1.5 nm. It has also been
argued that exciton diffusion between Ir-dye molecules is
predominantly due to Dexter transfer.
[ 33,34 ]
In view of this
uncertainty, we take the Dexter prefactors such that at a dis-
tance equal to the Förster radius both rates are approxi-
mately equal. At much larger distances the diffusion is
then predominantly due to the Förster process, whereas at
distances of 1–2 nm also Dexter transfer contributes.
The detailed mechanisms of triplet-polaron quenching
(TPQ) and triplet-triplet annihilation (TTA) are not well known.
In this paper, we describe both processes as immediate when
the two interacting species are at a distance a . A distinction
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER
2028
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
is made between TPQ processes in which the triplet is trans-
ferred to the site at which the polaron resides (“TPQ-p pro-
cess”) or one in which the polaron hops to the site at which
the triplet resides (“TPQ-t process”) (see Figure 2 ). Making this
distinction does not affect the roll-off. However, it can affect
the degradation rate (see Section 5). After a TTA process one
triplet exciton is assumed to be left, randomly chosen to be on
one of the dye sites. Under these assumptions, no free para-
meters describing TPQ and TTA are introduced. The TPQ rate
is then determined predominantly by the polaron density and
the polaron diffusion coeffi cient (see the next section), as for
the dilute systems considered the triplet diffusion coeffi cient
is relatively small. The TTA rate is at the low voltages consid-
ered for lighting applications very small, in view of the small
triplet volume density obtained for the commonly used short
radiative lifetime phosphorescent Ir-based emitters, and the
small triplet diffusion coeffi cient.
[ 5 ]
In a forthcoming paper, we
investigate to what extent TPQ still provides the dominant con-
tribution to the roll-off when describing TPQ and TTA as long-
range Förster-type processes with Förster radii up to 5 nm,
[ 35 ]
as
suggested, for instance, for the case of TTA from the work of
Reineke et al.
[ 16 ]
2.4. Degradation
In this paper, we describe OLED degradation as a result of the
conversion of emitter sites to non-emissive sites with an infi -
nite value of Γ
nr
when on that site a TPQ process occurs, with
a probability p
degr
. All other parameters are kept constant. The
effect of degradation is only included when mentioned explic-
itly. In such a case, we fi rst carry out a steady-state 3D-KMC
simulation without degradation ( p
degr
= 0), in order to obtain
equilibrated charge and exciton distributions. Subsequently,
we continue the simulation under the condition p
degr
= 1. Our
approach thus employs the largest possible acceleration, in
order to make the simulations feasible. We fi nd that such an
approach is permitted, as simulations using a smaller value of
p
degr
lead to a decrease of the decay rate which is proportional
to p
degr
. We remark that p
degr
is in reality expected to be a very
small number. The decay of the emission intensity is found to
show in general a stretched-exponential-like shape, as is also
observed experimentally. The lifetime is shorter for the case of
TPQ at the site where the triplet resides (TPQ-t) than for the
case of TPQ at the site where the polaron resides (TPQ-p), as
only in the former case all TPQ processes occur on an emitter
molecule (see Section 5).
2.5. Example: Mixed Matrix Symmetric OLED
As a fi rst example, we discuss the results of 3D-KMC simu-
lations for a symmetric OLED with a layer structure (anode |
HTL (40 nm) | EML (20 nm) | ETL (40 nm) | cathode), with
a mixed-matrix emissive layer containing equal concentra-
tions of the hole transport layer (HTL) and electron transport
layer (ETL) materials (host molecules) and 4 mol% of emitter
molecules (guest). The energy level structure is shown in
Figure 3 a, and the simulation parameters as given in Table 1
are used. The shallow LUMO and deep HOMO energies of the
HTL and ETL, respectively, ensure perfect carrier blocking, so
the recombination effi ciency is 100%. We already showed for
similar single-matrix OLEDs that the carefully chosen guest
trap depth (0.2 eV for electrons and holes) leads to a quite
optimal balance between a small roll-off and a low operating
voltage.
[ 5 ]
First, the OLED performance under steady-state
conditions (no degradation) is discussed (Figures 3 b–e). Sub-
sequently, the decrease of the emission in the presence of deg-
radation is shown (Figure 3 f).
Figure 3 b shows the roll-off of the IQE as a function of the
current density and (inset) the current–voltage ( J ( V )) curve in
the range 3–6 V. The full curve is a fi t to the roll-off using the
often-used empirical expression
1
IQE
rad
50
J
J
m
η
η
=
+
(4)
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Figure 2. The two types of triplet-polaron quenching processes considered: a) TPQ-p process, and b) TPQ-t process. Empty circles: neutral molecules
in the ground states. Filled circles: molecules on which a polaron resides, in the polaronic ground state. Stars (initial confi guration): neutral molecules
in the lowest triplet exciton state. Stars (intermediate confi guration): polaron in an excited state.
FULL PAPER
2029
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
with J
50
the current density at which the IQE has been
reduced by a factor 2, and m an exponent which increases with
increasing steepness of the roll-off around J
50
. It may be seen
that this expression provides an excellent fi t. Alternatively, the
IQE may be written as
η
IQE
= [Γ/(Γ + Γ
q
)]
η
rad
, with Γ
q
the effective
rate of all loss processes due to quenching (and annihilation).
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Figure 3. 3D-KMC simulation example, as discussed in Section 2.5. a) Energy level structure of the OLED considered. b) Current density dependence
of internal quantum effi ciency. The full curve is a fi t using Equation ( 4) . c) Current density dependence of the relative quench rate. d) Hole and electron
concentration profi les at 3 V. The full curves are guides-to-the-eye. The molecular layers in the EML are at a distance of 41 to 60 nm from the anode,
and the cathode is at a distance of 101 nm from the anode. e) Emission profi le at 3 V. The full curve is a guide-to-the-eye. f) Time-dependence of the
normalized emission at 6 V (degradation study), assuming only TPQ-t processes and assuming p
degr
= 1, with and without exciton diffusion (closed
and open symbols, respectively). The dashed curve is a stretched-exponential fi t to the simulation results without diffusion, using the parameter values
discussed in the text. In (b,c) the statistical uncertainty of the data points is smaller than the size of the data points. In (d–f) the statistical uncertainty
may be judged from the variability with the position and time, respectively.
FULL PAPER
2030
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
It follows from Equation ( 4) that the relative quench rate, Γ
q
/ Γ,
may be expressed as
1
rad
IQE 50
J
J
q
m
η
η
Γ
Γ
=−=
(5)
A double logarithmic plot of the quantity [(
η
rad
/
η
IQE
) 1]
should then yield a straight line with a slope m and an x -axis
zero-crossing at J = J
50
. We call this a “quench rate plot”. Pos-
sible deviations from Equation ( 4) for small current densities
can be seen more clearly from such a plot than from the roll-off
curve. Figure 3 c shows that Equation ( 4) is indeed well obeyed
within the current density range studied. The fi t parameters
are J
50
= 10.0 kA/m
2
and m = 0.80. Such a fi t makes it possible
to quite accurately derive the roll-off at small current densities,
where the simulation results are somewhat less accurate due
to relatively larger statistical fl uctuations. The current density
J
90
= (1/9)
1/ m
J
50
, which is a more commonly used measure for
the roll-off,
[ 12 ]
is equal to 642 A/m
2
. This value is higher than
the highest values of J
90
for phosphorescent OLEDs reported to
date, which range up to 300 A/m
2
.
[ 12 ]
We attribute the improved
roll-off in the simulated device to the effect of carrier trapping
on the mobility in the EML. Figure 3 d,e shows the charge car-
rier concentration and emission profi les, respectively, at 3 V.
As may be seen from the fi gures, strong accumulation of elec-
trons at the anode-side of the EML and of holes at the cathode-
side of the EML is prevented, and the emission profi le is quite
uniform across the thickness of the EML. A study in Section
4 of the effect of the guest trap depth supports this picture.
We nd that in the current density range studied the roll-off is
almost completely due to TPQ. We found a similar conclusion
for Ir(ppy)
3
and PtOEP based devices, where the small contri-
bution of TTA was attributed to the relatively small triplet con-
centration in the devices at the current densities studied.
[ 5 ]
Figure 3 f shows for the same device the decrease of the
emission at 6 V as a function of the time during a degrada-
tion simulation, assuming only TPQ-t processes and for p
degr
=
1. The excited polaron is then in all cases located on the dye
molecule, so that each TPQ process leads to degradation of the
dye site involved. The simulations were carried out with and
without triplet exciton diffusion. Switching off diffusion makes
the simulations computationally less intensive, so that for the
same total simulation (CPU) time the simulated time is much
larger. Under these strongly accelerated conditions, the time at
which the emission is reduced to 90% of the t = 0 value (LT90)
is (5.5 ± 1.0) and (7 ± 2) µs, respectively. For the small emitter
concentration used, the effect of diffusion of excitons to already
degraded sites is thus quite small. The fi gure shows that a
stretched exponential function (dashed curve, I exp[( t /
τ
life
)
β
)
with a (1/e) lifetime
τ
life
80 µs and
β
0.91) provides a rea-
sonable, although not perfect, fi t. In Section 4 we show that dif-
fusion becomes more important at larger dye concentrations.
Simulations of this type are expected to provide a prediction of
the real lifetime for any other dye concentration and EML thick-
ness after extrapolation of the simulation results to smaller cur-
rent densities, and after a “calibration” experiment to deduce
the actual value of p
degr
.
3. The Uniform Density Model—The Relationship
Between Roll-Off and Lifetime
In this section, it is fi rst shown how within the framework of
the uniform density model (UDM) simplifi ed expressions may
be obtained for the dependence of the IQE roll-off on the mate-
rial and device parameters (subsection 3.1). The analysis leads
to a shape of the roll-off curves which is consistent with the
empirical roll-off given in Equation ( 4) , and obtained from the
3D-KMC simulations shown in Figure 3 . In subsection 3.2, it is
shown how, under the assumption that upon a triplet-polaron
quenching process subsequently (with a small probability) a
degradation process takes place, the lifetime is related to mate-
rial and device parameters. The analysis suggests that the expo-
nent which gives the superlinear decrease of the lifetime with
increasing current density in an accelerated lifetime experi-
ment may be obtained from the shape of the roll-off curve. In
Sections 4 and 5, the predictions from the model will be com-
pared to simulation results for various OLED layer structures.
3.1. Roll-Off
Within the TPQ-related degradation mechanisms considered in
this paper, the roll-off and lifetime are related. Even if for cer-
tain operational conditions the IQE loss due to TPQ is small,
it will thus be of interest to understand how it depends on the
material parameters. In general, the electron, hole and exciton
densities show a dependence on the position in the EML,
precluding the development of an analytical model for the effi -
ciency roll-off. However, we fi nd that it is nevertheless useful
to further develop a simplifi ed model within which the carrier
density non-uniformity across the EML or across the part of the
EML within which the emission takes place (“emission zone”)
is neglected.
[ 11,12 ]
Within this uniform density model, an OLED
is regarded as a “chemical reaction vessel”, with “reactants
(electrons and holes) and “products” (excitons). The interactions
between all these species can be complex, but the neglect of a
spatial dependence strongly simplifi es the analysis. Figure 3 e
shows that it is indeed possible to obtain under some condi-
tions a quite uniform emission profi le in the EML. In this
section, we extend the UDM by including the effect of a charge-
carrier density of the mobility and diffusion coeffi cient, and by
including a mechanistic description of the TPQ rate, leading to
a refi ned expression for the roll-off.
We consider a phosphorescent OLED with in the emission
zone uniform and equal electron and hole volume densities,
n
e
= n
h
= n , and a uniform triplet density, n
T
. Under steady-state
conditions, the exciton generation rate is equal to the exciton
loss rates due to radiative and non-radiative decay, TPQ, TTA
and dissociation. Neglecting TTA (as the simulations shown in
Section 2.5 and shown by van Eersel et al.
[ 5 ]
suggest that TPQ is
more relevant) and dissociation, the triplet density balance may
under steady-state conditions be written as
d
d
2
2
0
TT
TPQ T
2
n
t
n
knn
e
nn
τε
μ
()
=− + =
(6)
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER
2031
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
with k
TPQ
the TPQ rate coeffi cient, µ ( n ) the charge density
dependent mobility, which is assumed to be equal for holes and
electrons, and with the last (generation) term as expected from
the Langevin formula, neglecting the electric fi eld dependence
of the exciton generation rate and neglecting a correction to
the mobility in bipolar devices.
[ 36 ]
The radiative decay rate will
be proportional to the triplet density, so that from Equation ( 6)
the IQE roll-off is given by
012
IQE
T
T TPQ
rad
rad
TPQ
J
n
nk k n
ηη
η
τ
()
()
=
=
=
+
(7)
Simple expressions for the current density dependent functions
n ( J ) and k
TPQ
( J ) may be obtained as follows. The condition of
charge conservation (the number of injected carriers is equal to
the number of excitons created, both per unit of time and area)
implies that
2
2
J
e
e
nnd
ε
μ
()
(8)
with
d the thickness of the recombination zone (or, for a well-
designed OLED, the thickness of the EML). We assume that the
mobility may be expressed as
1
t
n
n
N
b
μμ
()
(9)
where the prefactor µ
1
is the mobility as obtained by extrapola-
tion from the carrier density range of interest to the total site
density, N
t
(i.e., at a carrier concentration c n / N
t
equal to 1),
and with b > 0. The electric fi eld dependence of the mobility has
been neglected. Equation ( 9) is exact for the case of an expo-
nential density of states (DOS),
[ 37 ]
and provides for a Gaussian
DOS a fair approximation in charge carrier density ranges with
a width of at least one order of magnitude.
[ 29 ]
In both cases,
b increases with increasing energetic disorder. An effective
carrier density dependence of the mobility arises also in host-
guest systems with a bimodal Gaussian density of states.
[ 38 ]
The
exponent b depends then on the energy distance between the
host and guest states. Combining Equations ( 8) and ( 9) leads to
2
2
1
1
2
n
N
ed
J
t
b
b
ε
μ
=
+
(10)
Note that n does not depend on the triplet-polaron quench rate.
Triplet-polaron quenching is the result of encounters of dif-
fusing triplets and polarons. In phosphorescent OLEDs, with
typical dye concentrations of 510 mol%, the diffusion length
of triplets is very small (of the order of the Förster radius, which
we take 1.5 nm). TPQ is therefore predominantly due to the
diffusion of the charge carriers, so that k
TPQ
is proportional
to the charge carrier diffusion coeffi cient, D . From diffusion
theory,
[ 39 ]
it is known that for the case of a capture radius R
c
,
the rate coeffi cient is then given by
44
TPQ c
B
c
kn DnR
kT
e
nR
ππμ
() () ()
==
(11)
where the second step has been made using the Einstein equa-
tion which relates the diffusion coeffi cient to the mobility. The
TPQ rate coeffi cient is thus carrier density dependent. The
electric fi eld dependence of the diffusion coeffi cient has been
neglected. For simplicity, we have used the standard form of the
Einstein equation, for non-interacting particles, instead of the
generalized Einstein equation which would lead to a correction
at high carrier densities.
[ 40 ]
From Equation ( 7–11) it follows that
18
2
IQE
rad
B
c
1
t
1
2
2
1
2
J
kT
e
R
Ned
J
b
b
b
η
η
πτ
με
()
=
+
+
+
+
(12)
This simplifi ed analysis leads thus to expression for the roll-off
which is consistent with the empirical expression (Equation 4 )
which is often found experimentally in OLEDs. Also the roll-
off as obtained from simulations for the devices studied in this
paper is quite well described using Equation 4 (see Section 2.5
and Section 4). The parameter values m and J
50
are given in
Table 2 for the general case and for the specifi c case of a constant
mobility ( b = 0). If all other factors remain the same, the roll-off
can thus be reduced by i) decreasing the mobility in the EML (as
then the TPQ rate decreases), ii) increasing the emission zone
thickness (as then the polaron density decreases), iii) decreasing
the radiative decay time (as then radiative decay better outcom-
petes TPQ), and iv) reducing the TPQ capture radius.
The model suggests that the slope of the quench rate curve
will be in the range 0.5 to 1.0, depending on the (effective) dis-
order of the mobility in the EML. The value m = 0.80 found
for the example discussed in Section 2.5 is well in that range.
The effective mobility exponent b is then approximately equal
to 3. Interestingly, only a relatively weak sensitivity of the rel-
ative quench rate and J
50
to the mobility prefactor
μ
1
is then
expected, viz. Γ
q
/ Γ µ
1
0.2
and J
50
µ
1
0.25
.
3.2. Lifetime
Within the UDM, the degradation rate of an emitter molecule
in the EML is given by
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Table 2. Summary of roll-off parameters for the case of uniform and
equal carrier densities, equal electron and hole mobility functions, equal
exciton-polaron quench rates, and no dissociation. For the case
b = 0, we have µ
1
= µ .
General case
Constant mobility ( b = 0)
m
b
b
1
2
=
+
+
m
1
2
=
J
ed
N
kT
e
R
b
b
b
b
2
8
50
2
1
t
1
1
B
c
2
1
ε
μ
πτ
=
+
+
+
J
ed
kTR32
50
4
Bc
2
εμ π τ
()
=
×
FULL PAPER
2032
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
d
d
1
degr
q qd degr degr
f
t
gp p p f
()
× × ×
(13)
with
f
degr
( t ) the fraction of the emitter molecules which on time
t has already degraded, p
q
the probability that an exciton is lost
due to TPQ, p
qd
the probability that the TPQ process occurs on a
dye site (which depends on the type of TPQ-process considered
and on the emitter density, see Section 5), and p
degr
the proba-
bility that an excited polaronic dye molecule degrades. The value
of p
degr
depends on the chemical stability of the molecule, and
must be determined experimentally. Within the framework of
the UDM, the exciton generation rate per dye site is equal to
dye
g
J
edn
=
××
(14)
with
n
dye
the volume density of emitter molecules in the EML.
If the light emission intensity I ( t ) is proportional to the fraction
of non-degraded sites, as expected in the absence of exciton dif-
fusion, it is (from Equation ( 13) ) given by
10exp 0
degr
life,nd
It f I
t
I
τ
()
() () ()
=− =
(15)
with a (1/e) lifetime when no exciton diffusion is included,
τ
life,nd
, given by
1
life,nd
qqddegr
dye
pp p
edn
J
τ
××
(16)
Using
that p
q
= (1
η
IQE
/
η
rad
), when writing
η
IQE
=
η
rad
/
[1+( J / J
50
)
m
], where J
50
and m can be obtained from Table 2 , it
follows that p
q
( J / J
50
)
m
for small current densities ( J << J
50
),
so that then
1
life,nd
qd degr
dye 50
1
pp
edn J
J
m
m
τ
≅×
×× ×
+
(17)
The model thus establishes a relationship between the roll-
off and the lifetime in the absence of exciton diffusion. If
valid, Equation ( 17) can be used to predict
τ
life,nd
already from
a steady-state simulation, instead of from a usually computa-
tionally more demanding explicit 3D-KMC lifetime simulation.
For the case of the systems discussed in Section 2.5, the dye
concentration dependence of the lifetime predicted by Equation
( 17) will be investigated in Section 5.
Equation ( 17) furthermore leads to an algebraic dependence
of the lifetime on the current density,
τ
life
J
n
, with a current
density acceleration exponent n = m + 1. As m was predicted to
fall in the range 0.5 to 1, n will be in the range 1.5 to 2. Such a
current density dependence is indeed often observed in acceler-
ated lifetime experiments.
[ 41 ]
4. Sensitivity of the Roll-Off to Material
Parameters
The benefi ts of using the symmetric mixed-matrix OLEDs of
the type shown in Figure 3 a are found to be quite similar to
those of the otherwise identical single-matrix OLEDs studied by
van Eersel et al.
[ 5 ]
Due to the use of dyes which act as hole and
electron traps with a trap depth, Δ
1
, around 0.2 eV, the carriers
are slowed down in the EML, so that no charge accumulation
occurs in the EML near the interfaces. At voltages around 3 V
and below, the emission profi les are quite uniform, leading to a
small IQE-loss. In this section we further analyze the sensitivity
of the roll-off to Δ
1
, to a hole (electron) injection barrier Δ
2
from
the HTL to the host in the EML (from the ETL to the host in
the EML), and to various other material parameters including
the triplet emissive lifetime, triplet binding energy and the
Förster radius for triplet diffusion. Figure 4 a shows the energy
level structure of the devices studied. The analysis leads to two
energy level design rules from which the roll-off can be mini-
mized without introducing a large voltage increase.
Figure 4 b shows the quench rate curves for devices of the
type shown in Figure 4 a, with variable guest trap depths Δ
1
in the range 0–0.5 eV and with Δ
2
= 0 eV. The simulations
were carried out for a series of voltages in the range 3–6 V, for
the parameter values given in Table 1 and for a typical Ir-dye
concentration of 4 mol%. Results for the case Δ
1
= 0.2 eV were
already given in Figure 3 . The fi gure confi rms in a quantita-
tive manner the more qualitative picture mentioned above. It
may be seen that for all values of Δ
1
, the empirical expression
(Equation ( 4) ) for the roll-off is well obeyed, with a value of J
0
which increases until it saturates for Δ
1
0.2 eV. The use of
deeper traps leads to a signifi cant reduction of the current den-
sity at a given voltage, and is therefore unfavorable from a point
of view of optimizing the power effi ciency. That may be seen
from the current density at a fi xed voltage, that is, 3 V and 4 V
(indicated in the fi gures by large and fi lled symbols). We fi nd
that variations of i) the disorder energy
σ
, from 0.075 to 0.125 eV,
ii) the EML layer thickness d , from 10 to 30 nm (at a fi xed 100 nm
total thickness), or iii) the wavefunction decay length
λ
(from
0.3 to 0.1 nm) have no signifi cant or only little effect on this
picture. For single-matrix OLEDs we fi nd that for Δ
1
< 0.2 eV
the quench rate is a factor 1.5 larger than for the mixed-matrix
devices, and that saturation occurs at Δ
1
0.25–0.3 eV; the
quench rate curves coincide then to those shown in Figure 4 b.
We attribute this to the somewhat larger mobility in single-
matrix OLEDs, as all host molecules contribute to the transport,
so that deeper traps are needed to obtain a uniform emission
profi le. A fi rst design rule for symmetric OLEDs is thus that the
dye trap depth Δ
1
should ideally be 0.2 eV or 0.25–0.3 eV, for
mixed-matrix or single-matrix OLEDs, respectively.
As a second design rule, we fi nd that the HTL-EML(host)
and ETL-EML(host) injection barriers Δ
2
should ideally be
(0.0 ± 0.1) eV. This may be concluded from Figure 4 c, which
shows the effect of variations of Δ
2
from 0.3 to +0.3 eV, for a fi xed
value Δ
1
= 0.2 eV. For positive injection barriers, accumulation of
charges occurs in the transport layers, which enhances the TPQ
rate near the EML interfaces. When decreasing the barrier to
Δ
2
0, the decrease of the relative quench rates saturates.
Although a further reduction of Δ
2
does not further diminish the
roll-off, it gives rise to an increase of the voltage required to obtain
a certain current density, as dissipation occurs due to the energy
loss when the carriers enter the EML.
Figure 4 c shows that the quench rate at small current den-
sities and the slope parameter m characterizing the relative
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER
2033
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
quench rate curves do not only depend on the transport and
excitonic parameters in the EML, but also on the injection
boundary conditions. We fi nd that their effect on the shape of
the emission profi les, and thereby on the roll-off, cannot be
neglected. Giving a detailed quantitative analysis is beyond the
scope of this paper; we limit ourselves here to a brief qualita-
tive discussion. In the absence of injection barriers (Δ
2
0 eV),
the emission profi les for systems with Δ
1
0.2 eV are found
to be quite uniform at low voltages. However, above 3 V, only
part of the exciton generation occurs in the bulk of the EML. A
fraction of the charges passes the EML and is fi nally blocked.
Charge accumulation zones develop near the EML interfaces,
and the emission originates to an increasing extent from these
interfacial zones. The gradual splitting of the emission profi le
into two narrow peaks near the interfaces is expected to con-
tribute to the large slope of the relative quench rate curves.
Using Equation ( 10) , we deduce for such devices a value of
b 1.33 from an analysis of the carrier density dependence of
the current density in the center of the EML (where due to the
symmetry the electron and hole densities are equal). Within the
framework of the UDM, this value would lead to m 0.7, which
is indeed somewhat smaller than the value m 0.80 obtained
from the quench rate curves. In the presence of large barriers
for carrier injection into the EML (Δ
2
0.2 eV), charge accumu-
lation occurs already at small voltages near the EML interfaces.
The emission is then almost completely confi ned to the two
thin interfacial zones. As a result of the large local carrier den-
sity, TPQ leads already at low current densities to a relatively
large quench rate. With increasing voltage (and current den-
sity), the effect of the injection barriers at the EML interfaces
is reduced. There is then less charge accumulation near the
interfaces, and a larger fraction of the exciton formation occurs
in the bulk of the EML, where the charge density is relatively
small. These two effects explain the small slope of the relative
quench rate curve for such cases, for example, m 0.20 for
devices with Δ
2
= 0.3 eV (as may be deduced from Figure 4 c).
As a next step, we have investigated the sensitivity of the
roll-off to a variation of the hopping attempt rate to the fi rst
neighbor,
ν
1
, throughout the entire device. This corresponds to
a variation of the mobility prefactor µ
1
introduced in Section 3.
The simulations confi rm the expectation that for the symmetric
devices studied the relative quench rate curves are only weakly
sensitive to µ
1
. For simulations with
ν
1
enhanced or decreased
by a factor of 100 as compared to the value given in Table 1 ,
the relative quench rates as extrapolated to J = 10 A/m
2
are
found to differ by only a factor 5.
[ 42 ]
From Equation ( 12) ,
a factor 6.3 would be expected when assuming the value
b = 3.0 which would follow from the observed slope parameter
m = 0.80. At larger current densities an even smaller effect of
changing
ν
1
was found; the J
0
-values were only different by a
factor 2. We emphasize that this result does not imply that
the hopping attempt rates are not important to the roll-off of
OLEDs. Layer-specifi c or charge carrier-specifi c differences, for
example, may result in a non-uniform emission profi le and
hence in an increased effi ciency loss.
Figure 5 a,b show the sensitivity of the quench rate curves to
the effective radiative decay time
τ
, at a fi xed value of the radia-
tive decay effi ciency (
η
rad
= 0.75), for a triplet exciton binding
energy, E
T, b
, equal to 1.0 and 0.5 eV, respectively. Simulation
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Figure 4. a) Energy level structure of the symmetric OLEDs studied. In
the EML, a 50:50 mixture of two materials with a 4.0 eV single-particle
energy gap is used. b) Quench rate curves for OLEDs as show in (a),
with a varying energy distance Δ
1
between the host and guest levels in
the EML and with Δ
2
= 0 eV. The results shown in the fi gure provide a
motivation for the fi rst OLEDs design rule, discussed in Section 4. c)
Quench rate curves for OLEDs as shown in (a), with fi xed energy levels
in the EML (Δ
1
= 0.2 eV), but with varying barrier heights Δ
2
from the
HTL and the ETL to the host levels in the EML. The Δ
2
= 0.0 eV case is
as shown in Figure 3 . The results shown in the fi gure provide a motiva-
tion for the second OLEDs design rule, discussed in Section 4. In both
gures, large and fi lled symbols indicate the data points obtained at 3
and 4 V. The statistical uncertainty of the data points is for J > 10
2
A/m
2
of the order of the symbol sizes, as may be judged from a comparison
with the linear dashed guides-to-the-eye, but becomes somewhat larger
for smaller current densities.
FULL PAPER
2034
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
results for 0.75 eV (not shown) are essentially the same as for
1.0 eV. As mentioned in Section 2.4, a signifi cant dissociation
rate is expected unless E
T, b
is well above E
CT,b
= 0.48 eV. This
is confi rmed by simulation results. If E
T, b
is 0.75 eV or higher,
the quenching loss is essentially independent of E
T, b
. Further-
more, it is simply inversely proportional to
τ
(dashed lines) as
expected from Equation ( 12) . The roll-off is sensitive to E
T, b
below a value of 0.75 eV, as may be seen from the J
50
( E
T, b
)
dependence shown in Figure 6 (closed symbols). For E
T, b
=
0.5 eV, the quench rate is signifi cantly enhanced, in particular
at small current densities, so that the roll-off curves do not
obey Equation ( 4) anymore. The observed reduction of the
slope of the quench rate curves, to values which can be smaller
then 0.5, may be understood from an extension of the UDM
which includes dissociation.
[ 42 ]
The observation that at very
large current densities the effect of dissociation becomes less
may be explained by considering that then such a high volume
density of charge carriers is available for exciton formation
that very soon after dissociation already new pairs are formed.
At the 4 mol% dye concentration assumed in the pre-
ceding simulations, and for the small Förster radius for triplet
diffusion assumed ( R
F,diff
= 1.5 nm), triplet diffusion does
not contribute signifi cantly to the TPQ rate. In Figure 7 , the
results of a study of the role of triplet diffusion are shown.
For clarity, Dexter-type diffusion has been switched off. Var-
ying R
F,diff
in steps of 1.5 nm reveals only a signifi cant con-
tribution to the roll-off for R
F,diff
equal to 3 nm and larger,
for the default 4 mol% dye concentration as well as for
8 mol% devices. The fi gure also shows that increasing the
dye concentration increases the sensitivity to triplet diffusion,
as expected. The effect of triplet diffusion is largest at small
current densities, so that the quench rate curves become non-
linear, as due to the carrier density dependence of the polaron
diffusion coeffi cient the relative role of polaron diffusion is
larger at large current densities. We conclude that for dye con-
centrations of 8 mol% or less, the uncertainty concerning the
most appropriate description of triplet diffusion is expected to
contribute only a small uncertainty to the simulation results
presented above. We note that triplet diffusion can also con-
tribute to less roll-off, viz. when in the case of a strongly non-
uniform emission profi le the triplets migrate to regions with a
smaller polaron density.
[ 43 ]
Realizing OLEDs which satisfy the fi rst design rule (equal
0.2 eV hole and electron trap depths due to the guest levels
in the EML) is in practice not always well possible, due to a
lack of suitable host materials with accurately tuned HOMO
and LUMO levels. It is therefore of interest to investigate the
sensitivity to deviations of the precise relative positions of the
host and guest energy levels. The open circles in Figure 5 show
the results of such a study, carried out for asymmetric mixed-
matrix OLEDs with Δ
1,h
= 0.3 eV hole trap depth and a Δ
1,e
=
0 eV electron trap depth (as shown in the inset). By choosing
unequal concentrations of the hole and electron transporting
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Figure 5. Current density dependence of the relative quench rate for
OLEDs as shown in Figure 3 a for various effective radiative decay times
relative to the default value (
τ
0
= 1.38 µs, see Table 1 ), for a) E
T, b
=
1.0 eV, and b) E
T, b
= 0.5 eV. Thick curves indicate the results for
τ
=
τ
0
, and
dashed lines indicate the linear
τ
/
τ
0
dependence which is expected in the
absence of dissociation, and using the E
T, b
= 1.0 eV result as a reference.
The statistical uncertainty of the data points is of the order of the symbol
sizes, or smaller.
Figure 6. Closed circles: dependence of the 50% roll-off current density,
J
50
, on the triplet exciton binding energy, for symmetric mixed-matrix
OLEDs with an energy level structure as given in Figure 3 a (closed sym-
bols). Open circles: J
50
for asymmetric mixed OLEDs with the same HTL,
EML (host), and ETL energy level structure, but with a guest HOMO
and LUMO level structure as shown in the inset and with in the EML a
composition as indicated. The statistical uncertainty of the data points is
of the order of the symbol size.
FULL PAPER
2035
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
host materials in the EML (76 and 20 mol%, respectively), the
reduction of the effective hole mobility due to the trapping
of holes at dye sites is partially compensated. The simulation
results show a much stronger sensitivity of the roll-off to E
T, b
than for the case of a symmetric OLED with Δ
1,h
= Δ
1,e
= 0.2
eV (closed spheres). This may be attributed to the strongly
enhanced triplet exciton dissociation probability, as also CT
states consisting of a hole on a dye site and an electron on a
host site can be formed.
In view of the sensitivity of the roll-off to E
T, b
, it is impor-
tant to be able to obtain its value accurately from experiment.
By defi nition, E
T, b
= E
LUMO
E
HOMO
E
T
. It can therefore be
obtained, at least in principle, from a determination of the
single particle energy gap, E
LUMO
E
HOMO
, and a measurement
of the triplet energy. However, the accuracy of commonly used
methods for obtaining E
HOMO
and E
LUMO
from e.g., cyclic vol-
tammetry, from a combination of photoelectron spectroscopy
and inverse photoelectron spectroscopy, or from density func-
tional theory calculations, is a subject of current debate.
[ 5,44,45 ]
Partial information follows from the difference between the
singlet and triplet exciton energies, which for the commonly
used green emitter material Ir(ppy)
3
is approximately 0.4 eV.
[ 46 ]
Assuming a singlet exciton binding energy of a few tenths of an
eV, E
T, b
would then be around 0.6–0.9 eV.
5. Sensitivity of the OLED Lifetime to the Ir-Dye
Concentration
In this section, a comparison is given between the results of
explicit 3D-KMC simulations of the lifetime of OLEDs of
the type shown in Figure 3 a, for a range of Ir-dye concentra-
tions, and the prediction obtained from the UDM discussed
in Section 4. When exciton diffusion may be neglected, a fair
agreement is obtained for the TPQ-p and TPQ-t degradation
scenarios. From a comparison of the simulated lifetimes with
the experimental lifetime in state-of-the-art devices, an estimate
is made of the degradation probability per TPQ process, p
degr
.
Figure 8 shows the results of 3D-KMC simulations of the
lifetime of OLEDs of the type shown in Figure 3 a as a function
of the Ir-dye concentration in the EML, at a fi xed voltage (6 V),
for the case of TPQ-p and TPQ-t processes (see Figure 2 ). The
simulations were carried out with the purpose to investigate the
usefulness of the UDM as a means to predict the OLED life-
time from the results of steady-state simulations, and to inves-
tigate to what extent exciton diffusion (which was neglected in
the UDM) affects the results. Typical examples of the raw simu-
lation data were already shown in Figure 2 f.
In the case of a TPQ-t process, the excited-state polaron is
created on the molecule where the exciton was located. With a
very high probability, around 0.97 and 0.98 for Ir-dye concentra-
tions x
Ir
= 2 and 4 mol%, respectively, and for higher concen-
trations essentially 1, the excitons are located on the dye sites.
Each TPQ process then leads to the conversion of an emissive
dye molecule into a degraded (non-emissive) dye molecule, so
that p
qd
1 in Equation ( 17) . It follows from Equation ( 17) that
the lifetime is then expected to be proportional to the volume
density of Ir-dye molecules, if all other factors remain constant.
However, the actual concentration dependence is affected by
the concentration dependence of the current density, which
is found to be minimal around x
Ir
9 mol%. At smaller con-
centrations, the dyes act predominantly as trap sites, so that
the mobility then decreases with increasing x
Ir
, whereas at
larger concentrations direct dye-dye transport is the predomi-
nant transport mechanism. The mobility then increases with
increasing x
Ir
. Figure 8 (closed spheres) shows that the pre-
dicted LT90 lifetime indeed increases for small x
Ir
, but that it
shows a maximum around 15 mol%. The lifetime as obtained
from explicit 3D-KMC simulations is for small dye concentra-
tions within a factor 2 of the predicted value, but is smaller by
a factor of approximately 4 for large concentrations. The differ-
ence at large dye concentrations is mainly due to the neglect of
exciton diffusion in the UDM, as may be concluded from the
observation that simulations within which diffusion is switched
off (small circles) agree then excellently with the prediction
from the UDM. The small discrepancy at small concentrations
might be due to deviations from a uniform emission profi le,
either laterally or in the direction perpendicular to the EML.
Upon a TPQ-p process, the excited-state polaron is cre-
ated on the molecule at which the polaron was located. From
the steady-state simulations, we fi nd that the probability that
TPQ occurs on a dye site ( p
qd
) is to an excellent approximation
proportional to x
Ir
, and approximately 0.42 for x
Ir
= 0.1. As a
result, the lifetime as predicted from Equation ( 17) is approxi-
mately independent of x
Ir
(closed squares in Figure 8 ), and
larger than as expected for the TPQ-t scenario. The explicit
3D-KMC simulation results (open squares) confi rm the
latter expectation. However, the x
Ir
dependence is larger than
expected. A preliminary analysis suggests that the statistics of
the occurrence of emitter pairs strongly affects the lifetime for
this case. As degradation only occurs when the exciton and
polaron reside both on a dye molecule, it is for very small
x
Ir
rarer than would follow from continuum statistics. That
explains the higher than expected lifetime for the 2 mol%
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Figure 7. Dependence of the quenching curves on the Förster radius for
triplet exciton diffusion, R
F,diff
, for two values of the Ir-dye concentration,
obtained from simulations for the OLED shown in Figure 3 a. The full and
dashed curves are guides-to-the-eye, consistent with the numerical accu-
racy of the simulation data. The statistical uncertainty of the data points
is equal to symbol sizes, or smaller, except for the data points close to
J = 10
2
A/m
2
, where it is slightly larger.
FULL PAPER
2036
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
case, and the observation of saturation of the degradation, viz.
when each cluster of Ir-dye molecules contains a degraded
molecule. The emission occurs then from the isolated dye
molecules. Giving more in-depth analyses for both scenarios
is beyond the scope of this paper.
The LT90 lifetime obtained from the simulations is in the
range 5–50 µs when taking p
degr
= 1, and for x
Ir
in the 5–10
mol% range. The current density at 6 V as obtained from the
simulations is approximately 4000 A/m
2
. In effi cient stacked
state-of-the-art white OLEDs operated at a luminance equal
to 1000 cd/m
2
, the current density is much smaller, around
10 A/m
2
. Assuming a current density acceleration exponent
equal to 1.8 (as expected from the analysis given in Section 3.2),
an actual LT90 lifetime of 10,000 hours would then imply an
average value p
degr
of the order 10
8
. A method for deducing
that probability from the observed luminance degradation
versus time and from transient photoluminescence studies
of degraded devices has been demonstrated by Giebink et
al.
[ 9 ]
Interestingly, for the systems studied, with in the EML a
typical blue-emitting Ir-dye molecule embedded in a 4,4-bis(3-
methylcarbazole-9-yl)-2,2-biphenyl (mCBP) host, a similar deg-
radation probability per triplet-polaron encounter was found,
viz. roughly 2 × 10
9
.
We note that the validity of the assumption that the degrada-
tion probability p
degr
is independent of the material composition
and simulation parameters has yet to be proven. It is conceiv-
able that protective mechanisms leading to fast de-excitation of
the excited polarons depend on, for example, the dye concentra-
tion or the applied electric fi eld. If needed, such mechanisms
can be included in the 3D-KMC approach developed.
6. Summary, Conclusions, and Outlook
Results have been presented of three-dimensional kinetic
Monte Carlo (3D-KMC) simulations of the degradation in phos-
phorescent OLEDs, based on scenarios in which the emissive
dye molecules are converted to non-emissive molecules upon
a triplet-polaron quenching (TPQ) process. For the simulation
parameters assumed, TPQ processes provide the predominant
contribution to the IQE roll-off with increasing current density.
Therefore, the roll-off and lifetime are related. From a sim-
plifi ed model, assuming uniform charge carrier and exciton
densities across the emissive layer (EML), an analytical expres-
sion for the roll-off has been obtained in terms of the mate-
rials parameters. Within the same approach, an expression
has been obtained (Equation 17 ) from which the lifetime due
to TPQ-induced degradation can be obtained from steady-state
simulations. The model is found to provide a fair prediction of
the Ir-dye concentration dependence of the lifetime in a model
system, provided that exciton diffusion may be neglected.
The 3D-KMC model provides a means to investigate the
sensitivity of the roll-off and lifetime to the physical processes
assumed, to the materials-specifi c parameters and to the layer
stack architecture. In this paper, we have explored some of
these relationships, including the sensitivity to energy level
values, the mobility, the triplet binding energy, the emissive
lifetime, energetic disorder, the mobility and exciton diffu-
sion. We have established how in idealized symmetric OLEDs
the roll-off can be minimized by following design rules for the
trap depths of the dyes in the EML and for the injection bar-
riers from the transport layers to the EML. Furthermore, it was
found that the roll-off is quite sensitive to the triplet exciton
binding energy, E
T, b
, if that would be smaller than 0.75 eV (i.e.,
within the framework of our formalism less than 0.25 eV above
the binding energy of CT-excitons with the electron and hole on
nearest neighbor sites).
It follows from our results that the development of more
accurate methods for determining the HOMO and LUMO
energy levels and E
T, b
, with an accuracy better than 0.1–0.2 eV,
would make the simulations more versatile. Furthermore,
simulation-assisted OLED lifetime studies should be carried
out in order to establish the degradation mechanism. The sim-
ulation results suggest that it would be of interest to vary in
such a study the emitter concentration as a means to make a
distinction between different degradation scenarios.
Acknowledgements
The authors would like to thank the Philips Research e-Science
department for technical support. This research was supported by the
Dutch nanotechnology program NanoNextNL.
Received: July 28, 2014
Revised: September 25, 2014
Published online: November 4, 2014
[1] S. Reineke , M. Thomschke , B. Lüssem , K. Leo , Rev. Mod. Phys.
2013 , 85 , 1245 .
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com
Figure 8. Results of explicit 3D-KMC lifetime for the OLEDs shown in
Figure 3 a (open symbols) as a function of the Ir-dye concentration, at 6 V,
assuming TPQ-p processes (the triplet exciton hops to the site on which
the polaron resides, open squares) or TPQ-t processes (the polaron hops
to the site on which the triplet exciton resides, open circles). For the
latter case, also simulation results without diffusion are included (small
circles). The closed symbols give the predictions from the uniform den-
sity model (Equation ( 17) ). The vertical scale gives the actual LT90 device
lifetime, multiplied by the (very small) probability p
degr
that upon a TPQ
process degradation takes place. The results were obtained with p
degr
=
1. The curves are guides-to-the-eye. The statistical uncertainty of the data
points is equal to symbol sizes, or smaller.
FULL PAPER
2037
wileyonlinelibrary.com
©
2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
[2] R. Coehoorn , P. A. Bobbert , Phys. Status Solidi A 2012 , 209 ,
2354 .
[3] J. J. M. van der Holst , M. A. Uijttewaal , B. Ramachandhran ,
R. Coehoorn , P. A. Bobbert , G. A. de Wijs , R. A. de Groot , Phys. Rev.
B 2009 , 79 , 085203 .
[4] M. Mesta , M. Carvelli , R. J. de Vries , H. van Eersel , J. J. M. van
der Holst , M. Schober , M. Furno , B. Lüssem , K. Leo , P. Loebl ,
R. Coehoorn , P. A. Bobbert , Nat. Mater. 2013 , 12 , 652 .
[5] H. van Eersel , P. A. Bobbert , R. A. J. Janssen , R. Coehoorn , Appl.
Phys. Lett. 2014, 105, 143303.
[6] N. C. Giebink , S. R. Forrest , Phys. Rev. B 2008 , 77 , 235215 .
[7] D. Y. Kondakov , In Organic Electronics ; CRC Press, Taylor & Francis
Group, Boca Raton 2009 , p 211 .
[8] S. Schmidbauer , A. Hohenleutner , B. König , Adv. Mater. 2013 , 25 ,
2114 .
[9] N. C. Giebink , B. W. D’Andrade , M. S. Weaver , P. B. Mackenzie , J.
J. Brown , M. E. Thompson , S. R. Forrest , J. Appl. Phys. 2008 , 103 ,
044509 .
[10] N. C. Giebink , B. W. D’Andrade , M. S. Weaver , J. J. Brown ,
S. R. Forrest , J. Appl. Phys. 2009 , 105 , 124514 .
[11] M. A. Baldo , C. Adachi , S. R. Forrest , Phys. Rev. B 2000 , 62 , 10967 .
[12] C. Murawski , K. Leo , M. C. Gather , Adv. Mater. 2013 , 25 , 6801 .
[13] S. Reineke , G. Schwartz , K. Walzer , M. Falke , K. Leo , Appl. Phys. Lett.
2009 , 94 , 163305 .
[14] F. Lindla , M. Boesing , C. Zimmermann , F. Jessen , P. van
Gemmern , D. Bertram , D. Keiper , N. Meyer , M. Heuken , H. Kalisch ,
R. H. Jansen , Appl. Phys. Lett. 2009 , 95 , 213305 .
[15] N. C. Erickson , R. J. Holmes , Appl. Phys. Lett. 2010 ,
97 , 083308 .
[16] S. Reineke , K. Walzer , K. Leo , Phys. Rev. B 2007 , 75 , 125328 .
[17] H. van Eersel , P. A. Bobbert , R. Coehoorn , unpublished .
[18] C. Weichsel , L. Burtone , S. Reineke , S. I. Hintschich , M. C. Gather ,
K. Leo , B. Lüssem , Phys. Rev. B 2012 , 86 , 075204 .
[19] T. D. Schmidt , D. S. Setz , M. Flämmich , J. Frischeisen , D. Michaelis ,
B. C. Krummacher , N. Danz , W. Brütting , Appl. Phys. Lett. 2011 , 99 ,
163302 .
[20] W. C. Germs , J. J. M. van der Holst , S. L. M. van Mensfoort ,
P. A. Bobbert , R. Coehoorn , Phys. Rev. B 2011 , 84 , 165210 .
[21] M. Mesta , J. Cottaar , R. Coehoorn , P. A. Bobbert , Appl. Phys. Lett.
2014 , 104 , 213301 .
[22] A. Massé , R. Coehoorn , P. A. Bobbert , Phys. Rev. Lett. 2014, 113 ,
116604.
[23] S. L. M. van Mensfoort , S. I. E. Vulto , R. A. J. Janssen , R. Coehoorn ,
Phys. Rev. B 2008 , 78 , 085208 .
[24] O. Rubel , S. D. Baranovskii , P. Thomas , S. Yamasaki , Phys. Rev. B
2004 , 69 , 014206 .
[25] H. T. Nicolai , A. J. Hof , M. Lu , P. W. M. Blom , R. J. de Vries ,
R. Coehoorn , Appl. Phys. Lett. 2011 , 99 , 203303 .
[26] S. L. M. van Mensfoort , V. Shabro , R. J. de Vries , R. A. J. Janssen ,
R. Coehoorn , J. Appl. Phys. 2010 , 107 , 113710 .
[27] M. Schober , M. Anderson , M. Thomschke , J. Widmer , M. Furno ,
R. Scholz , B. Lüssem , K. Leo , Phys. Rev. B 2011 , 84 , 165326 .
[28] H. Bässler , Phys. Status Solidi B 1993 , 175 , 15 .
[29] W. F. Pasveer , J. Cottaar , C. Tanase , R. Coehoorn , P. A. Bobbert ,
P. W. M. Blom , D. M. de Leeuw , M. A. J. Michels , Phys. Rev. Lett.
2005 , 94 ,
206601 .
[30] V. Coropceanu , J. Cornil , D. A. da Silva Filho , Y. Olivier , R. Silbey ,
J.-L. Brédas , Chem. Rev. 2007 , 107 , 926 .
[31] J. Cottaar , L. J. A. Koster , R. Coehoorn , P. A. Bobbert , Phys. Rev. Lett.
2011 , 107 , 136601 .
[32] Y. Kawamura , J. Brooks , J. J. Brown , H. Sasabe , C. Adachi , Phys. Rev.
Lett. 2006 , 96 , 017404 .
[33] E. B. Namdas , A. Ruseckas , I. D. W. Samuel , S.-C. Lo , P. L. Burn ,
Appl. Phys. Lett. 2005 , 86 , 091104 .
[34] J. C. Ribierre , A. Ruseckas , K. Knights , S. V. Staton , N. Cumpstey ,
P. L. Burn , I. D. W. Samuel , Phys. Rev. Lett. 2008 , 100 , 017402 .
[35] H. van Eersel , R. A. J. Janssen , P. A. Bobbert , R. Coehoorn ,
unpublished .
[36] J. J. M. van der Holst , F. W. A. van Oost , R. Coehoorn , P. A. Bobbert ,
Phys. Rev. B 2009 , 80 , 235202 .
[37] M. C. J. M. Vissenberg , M. Matters , Phys. Rev. B 1998 , 57 , 12964 .
[38] R. Coehoorn , Phys. Rev. B 2007 , 75 , 155203 .
[39] S. Chandrasekhar , Rev. Mod. Phys. 1943 , 15 , 1 .
[40] Y. Roichman , N. Tessler , Appl. Phys. Lett. 2002 , 80 , 1948 .
[41] C. Kristukat , T. Gerloff , M. Hoffmann , K. Diekmann , In Organic
Light-Emitting Diodes (OLEDs): materials, devices and applications, Ch.
20 , (Ed: A. Buckley ), Woodhead Publishing, Cambridge, UK 2013 .
[42] These simulations were carried out under the condition of zero
exciton diffusion, in order to allow a proper comparison with the
roll-off as obtained from the UDM.
[43] G. He , M. Pfeiffer , K. Leo , M. Hofmann , J. Birnstock , R. Pudzich ,
J. Salbeck , Appl. Phys. Lett. 2004 , 85 , 3911 .
[44] B. W. D’Andrade , S. Datta , S. R. Forrest , P. Djurovich , E. Polikarpov ,
M. E. Thompson , Org. Electron. 2005 , 6 ,
11 .
[45] P. I. Djurovich , E. I. Mayo , S. R. Forrest , M. E. Thompson , Org.
Electron. 2009 , 10 , 515 .
[46] T. Tsuboi , J. Lumin. 2006 , 119–120 , 288 .
Adv. Funct. Mater. 2015, 25, 2024–2037
www.afm-journal.de
www.MaterialsViews.com