Вы находитесь на странице: 1из 104

Contents:

0 Introduction

I Concepts and Materials

I.1 Chemistry of Carbon


I.1.a Hybrid orbitals
I.1.b The benzene ring
I.1.c Conjugated molecules
I.2 Examples of organic semiconductors
I.3 Excitations in organic semiconductors
I.3.a Polarons and excitons
I.3.b Light emission from organic molecules
I.3.c Controlling the bandgap
I.4 Charge carrier injection

II Organic semiconductor applications and devices

II.1 Synthetic Metals


III.1.a Water- based synthetic metals
III.1.b Applications of synthetic metals
II.2 Organic field effect transistors
III.2.a Description of organic FET operation
III.2.b Requirements on OFET materials
II.3 Organic light emitting devices

1
II.3.a Overview over basic phenomena
II.3.b Bipolar carrier injection
III.4.c Exciton formation

2
0 Introduction

Knowledge of the following concepts and terms


from solid state physics is essential and will be
presumed:

Metal, semiconductor, Fermi level, work function,


conduction band, valence band, band gap, doping,
hole, chemical / electrochemical potential, wave
vector, quasiparticle, polaron, (Wannier)
exciton, conductivity, (positive / negative)
temperature coefficient (PTC / NTC).

Please read up in a solid state physics textbook


about any of these concepts you feel you are not
familiar with.

3
I.1 Chemistry of Carbon (C)
For our context, the key difference between
organic and inorganic solid matter is that
excitations in inorganic matter are delocalised
and best described by a wave vector k, while in
organic matter, excitations usually are localised
and k is not a good quantum number. To understand
organic semiconductors (and, maybe synthetic
metals?), we have to understand how something
like a ‘bandgap’ can arise within a single
molecule. The key to this understanding lies in
the chemistry of carbon.

12
The most common carbon isotope is C (nucleus has
6 neutrons, 6 protons), but there is a natural
13
abundance of 1.2% the C isotope with 7 neutrons,
6 protons. This has a nuclear magnetic momentum,
which is used in NMR. In atomic carbon, the 6
electrons occupy the following orbitals, table
I.1:

Table I.1

Orbital: 1s 2s 2px 2py 2pz


No. of electrons: 2 2 1 1 0

Table I.1: Electronic configuration of carbon. 1s


means principle quantum number (QN) n=1, Orbital
QN l = 0, consequently magnetic QN 0; 2 electrons
go into 1s due to 2 spins. Briefly, this is
2 2 2
written as 1s 2s 2p .

4
I.1.a Hybrid orbitals
Carbon, like most chemical elements, forms
covalent bonds. The driving force for chemical
reaction is the desire to share electrons between
different atoms to complete electronic shells.
Thus, usually,

Atomic Orbitals Æ Molecular Orbitals

Carbon should form 2 bonds to add 2 electrons to


complete the vacancies in the two incomplete p
orbitals (px and py): Carbon should be divalent
(form 2 single bonds).

In reality, carbon forms 4 bonds. In C (and some


other atoms), chemical bonding proceeds via
intermediate steps: ‘Promotion’ and
‘Hybridisation’:

Atomic orbitals Æ Hybrid Orbitals Æ Molecular


Orbitals

For hybridisation, carbon ‘promotes’ one 2s


electron into the empty pz orbital, we arrive at
2 1 3
1s 2s 2p . Then, C combines (‘hybridizes’) the
remaining 2s electron and either:

three 2p orbitals Æ sp3 hybrids


or two 2p orbitals Æ sp2 hybrids
or one 2p orbital Æ sp hybrid

5
sp3 hybrid orbitals have 4 ‘fingers’ pointing
into space ‘symmetrically’, i.e. into the corners
of a tetraeder. The angle between the ‘fingers’
o
is 109.5 . In this form, C can form 4 bonds, e.g.
by sharing electrons with Hydrogen’s 1s shell:
CH4 (methane), or with other sp3 carbon (e.g. H3C-
CH3, ethane). The C-C bond in ethane is called a
σ bond. σ bonds are very strong: Diamond consists
of carbon held together by σ bonds entirely.

sp2 hybrid orbitals have 3 ‘fingers’ in a plane,


o
with 120 to each other, plus one remaining p
orbital perpendicular to the plane. In this form,
C needs another sp2 hybrid C to form a molecule,
e.g. H2C=CH2: 2 of the 3 ‘fingers’ of each C bond
to H, as before, the third overlaps with another
C sp2 orbital to form a bond (σ bond). The
remaining p orbitals of either C overlap, as
well, to form another carbon/carbon bond, the so-
called π bond. (1σ+1π bond: Carbon double bond)
This is a weaker bond, and the respective orbital
is more ‘delocalized’, i.e. occupies relatively
large space rather far away from its original
carbon.

sp hybrid orbitals have 2 ‘fingers’ along one


o
axis (say, x) 180 to each other, plus 2
remaining p orbitals (along y and z axis). In
this form, C can bond e.g. with 2 H and another
sp hybrid. It forms one σ bond between the sp
orbitals, plus the remaining 2 p orbitals of each

6
molecule overlap to form two π bonds (carbon
triple bond). This is Ethene (acetylene), HC≡CH.

The bond lengths are: C-C ≅ 1.45 Å, C=C ≅ 1.33 Å,


C≡C z Å
-10
(1 Å = 10 m)

I.1.b The Benzene ring


o
sp2 hybrid orbitals have an angle of 120 with
respect to each other. Hence, by σ-bonding 6 sp2
carbons we can form a regular hexagon. Each C
will form 2 σ bonds, one with each of its
neighbours. There remains one sp2 orbital per C
to be ‘capped’, e.g. by a H. The remaining p
orbitals will again overlap to form π bonds. The
resulting benzene molecule may look like shown in
Fig. I.1:

Or

Fig. I.1: The two possible ‘borderline’


structures of benzene.

It is not quite clear where the π bonds should


be. In reality, a quantum mechanical
superposition of the two borderline states is
adopted, wherein it is impossible to assign
double bonds: The π electrons are completely

7
delocalized to form a ‘cloud’ that spans the
entire molecule.

= 1 / 2

Fig. I.2: The structure of benzene. The side of a


ring is 1.39 Å in length, intermediate between C-
C and C=C bond lengths.

The benzene ring is one of the most important and


versatile building blocks of organic chemistry.
Its delocalized π electrons have remarkable
properties with respect to their interaction with
light, and many molecules containing benzene
rings can donate or accept electrical charges
with relative ease. Much of molecular physics,
including organic semiconductor physics, is
concerned with molecules containing benzene
rings.

I.1.c Conjugated Molecules


The benzene ring is the prototype of conjugated
molecules, that is molecules with alternating
single/double or single/triple carbon bonds. In
conjugated molecules, π electrons delocalize

8
throughout the entire molecule and are relatively
loosely bound.

We can think of conjugated molecules being built


step-by-step by binding hybridised carbons
together (linear combination of atomic orbitals,
LCAO). In the molecular orbitals (MO)
description, we instead imagine a given, rigid
set of points at which atomic nuclei are fixed,
and ‘fill’ that skeleton with electrons to arrive
at the molecule. The LCAO and MO approach
correspond to two schools of quantum
chemistry/computer simulation. Both approaches,
of course, should lead to the same molecules, but
in the MO picture, the correspondence to
semiconductors is easier to see. We have to put N
electrons into the molecule to balance N positive
charges. The first electrons will cluster closely
to the atomic nuclei, resulting in almost
undisturbed atomic orbitals – that is equivalent
to saying that e.g. the carbon 1s electrons do
not participate in chemical bonds. But the last
few electrons will go into what we have called
delocalized π orbitals. Although we can trace π
orbitals to the hybridised atomic orbitals of
carbon, we have seen that in a conjugated
molecule, they may delocalize far from their
‘original’ carbon – hence, for the π cloud, the
MO picture is more appealing.

9
The last pair of electrons (one for each spin
state) to be ‘filled’ into the molecule occupies
a molecular orbital that is called ‘highest
occupied molecular orbital’ (HOMO). Note that a
half-filled HOMO would imply an unpaired spin,
i.e. the molecule would be a radical. Since the
completion of electronic shells is the driving
force behind chemical reactions, radicals usually
appear at intermediate stages of chemical
reactions, but not as end product. The next
molecular orbital beyond the HOMO is called
‘lowest unoccupied molecular orbital’ (LUMO).
HOMO and LUMO together are also referred to as
frontier orbitals.

We recognize a close similarity between the HOMO


(LUMO) of a molecule with the valence
(conduction) band in a crystalline semiconductor.
Due to this analogy, the energy difference
between LUMO and HOMO in a conjugated molecule is
often called the ‘bandgap’ of a conjugated
molecule. However, we will soon address the
limits of the analogy between inorganic and
organic semiconductors. There will be
considerable ambiguity how precisely to give a
quantitative figure for an organic bandgap from
experimental data. Note also that the distinction
between direct and indirect bandgap makes no
sense when k is not a good quantum number.

I.2 Examples of organic semiconductors

10
Here follows a reference list of conjugated
molecules and polymers, which are of importance
in the field of organic semiconductors, together
with acronyms by which they will be referred to
throughout this text. The sheer length of this
(hopelessly incomplete!) list underscores one of
the key assets of organic semiconductors: The
practically unlimited diversity of synthetic
organic chemistry allows tailoring materials with
a large property portfolio.

Historically, the organic semiconductor


discipline distinguishes between ‘polymeric’ and
‘low molecular weight’ organic semiconductors.
This distinction is nowadays blurred due to the
advent of a number of ‘hybrid’ materials, that
combine properties and attributes of low-
molecular weight and polymeric materials. A few
examples of these are included in the following
list, table I.2. We will refer to these materials
throughout the text, often with their acronyms
only as given in table I.2.

Table I.2

a.) Low molecular weight organic


semiconductors

S S S
S S S
6T

11
Pentacene

Perylene

N N

H3C CH3
TPD

N
N
O

PBD

C60

12
N O
N
O Al

O
N

Alq3

N
N Pt N
N

PtOEP

N O
Ir
O
S

2
btpacac

13
N O
Eu
N O 3

ADS053RE
N
C7H15O
S SC12H25
7O-PBT-S12

C6H13 C6H13
S S

C6H13 S S C6H13

S S
C6H13 C6H13
HHTT

ABTo O

S ABTo
O
O
HO N O
N N
HO
N
Ru N
N Ru N S
N
N N
HO N S S N

O S
ABTo
O
ABTo O
N3 Black dye

14
O

O2N NO2

O2N
TNF

b.) polymeric organic semiconductors

*
n *
PPV

MEH-PPV

N
C

n CN-PPV

* n *
PPE

15
* n *
PPP

C10H21

H3C

* *

H3C

C10H 21

MeLPPP
S
* n
*

R PAT

S n *
S

PTV

PTAA

16
* n *

R R PF

* n *

N N
C8H17 C8H17 S
F8BT
S
* n *
S

C8H17 C8H17
F8T2

c.) ‘hybrid’ materials

*
n *
N

PVK

17
N

N N

ST 638

sQP
O
O

O O

(H2C)nO O(CH2)n

N N

oxTPD

18
N

NDSB Dendron
(G2)

d.) synthetic metals

*
* n
PA

n *
*
PDA

19
NH N

N
NH

PAni

* S
n *

O O
PEDOT

Table I.2: A selection of organic semiconductors


and synthetic metals.

Let us briefly discuss the attributes of the


listed materials immediately. This discussion
will preview many of the topics that are
introduced more systematically later in the
description of organic semiconductor devices;
nevertheless the reader is invited to immerse
her/himself into the fascinating world of
molecular diversity at this stage.

a.) low molecular weight materials:


Hexithiophene (6T), Pentacene, Perylene, and TPD
(N,N'-bis-(m-tolyl)-N,N'diphenyl-1,1-biphenyl-
4,4'-diamine) are hole transporting and more or
less strongly fluorescent organic semiconductors
(Perylene more, 6T and TPD less). 6T is one
representative of the thiophene family of organic

20
semiconductors, which are known for their fast
hole mobilities and are often used in organic
transistors. PBD (2-(biphenyl-4-yl)-5-(4-tert-
butylphenyl)-1,3,4-oxadiazole) is an electron
conductor. Both TPD and PBD have been used as
carrier injection layers in multilayer device
architectures. C60 is a material with very high
electron affinity, and C60 derivatives have been
used as electron acceptors in organic
photovoltaic devices. However, electron transport
in C60 is very sensitive to even traces of oxygen,
which limits its practical potential. Alq3
(tris(8-quinolinolato)aluminum(III)) is an
organometallic Al chelate complex with efficient
green electroluminescence and remarkable
stability. Alq3 was used as the emissive material
in the first double layer organic light- emitting
device. PtOEP is a red phosphorescent porphyrine
derivative. The central Pt atom facilitates spin-
orbit coupling that allows light emission from
triplet excitons. btp2Ir(acac) (bis(2-(2'-
benzothienyl)-pyridinato-N,C-3')
iridiumacetylacetonate) is a representative of a
family of highly efficient phosphors that have
been used successfully as triplet- harvesting
emitters in efficient electrophosphorescent
devices. ADS053 RE is a trade name for the red-
emitting organolanthanide
Tris(dinapthoylmethane)mono (phenanthroline)-
europium(III). Organolanthanides transfer both
singlet- and triplet excitons to an excited
atomic state of the central lanthanide, resulting
in very narrow emission lines, i.e. spectrally

21
pure colours (here, the 612 nm red line of
Europium). 7O-PBT-S12 and HHTT are hole-
transporting calamitic and discotic liquid
crystals, respectively. Due to the stacking of
conjugated cores in smectic (7O-PBT-S12) and some
discotic (HHTT) liquid crystalline phases, both
can display rather high charge carrier
mobilities. ‘N3’ and ‘black dye’ are
organometallic dyes with broad absorption spectra
spanning the red and near infrared, which are
often used for the harvesting of solar photons in
dye- sensitized photovoltaic cells (‘Grätzel
cells’). Trinitrofluorenone (TNF) is often used
as electron acceptor for the formation of charge
transfer complexes with conjugated molecules.

b.) polymeric organic semiconductors


Poly(para-phenylene vinylene) PPV played an
outstanding role in the development of organic
electroluminescence. MEH- PPV and Cyano- PPV (CN-
PPV) are sidechain- substituted PPVs. Sidechains
promote solubility and also can change the
bandgap, and the type of transported charge
carrier. Poly(phenylene ethynylene) PPE and
poly(para phenylene) PPP are variations on a
similar theme. Methylated ladder- type PPP
(MeLPPP) is similar to PPP, but with all backbone
rings forced to be coplanar. PPVs, PPP, PPE, and
MeLPPP have been explored extensively in organic
light emitting devics. Poly(alkylated thiophene)
(PAT) and poly(thienylene vinylene) (PTV) are
less emissive, but have higher hole mobilities

22
and low ionisation potentials, which qualifies
them for use in organic field effect transistors.
However, thiophene- containing materials are
generally sensitive to oxygen, which for
practical organic transistor applications can not
be excluded. Instead, variations of
poly(triarylamine) (PTAA) are being developed for
hole- transporting transistors. Polyfluorene (PF)
is a blue emitter that has recently competed
successfully with PPV as organic light emitting
material. Typically, PF is copolymerized rather
than sidechain- substituted to modify its
properties. F8BT is an electron transporter and
efficient green emitter. F8T2 is a hole
transporter that works well in transistors. PF,
F8BT, and F8T2 also display interesting liquid
crystalline phases.

c.) hybrid materials:


Poly(vinyl carbazole) PVK is historically one of
the first (the first?) polymeric organic
semiconductors. PVK clearly is a polymeric
material, with the film forming and morphological
properties typical of polymers. However, the
semiconducting carbazole units dangle laterally
from a non- conjugated backbone, and are isolated
from each other. The electronic properties of PVK
are therefore very similar to those of low
molecular weight carbazole. PVK is used in
photocopiers, and was the first polymer for which
electroluminescence (EL) was reported. ST 638 is
the tradename for 4,4’,4”-Tris(N-(1-naphthyl)-N-

23
phenyl-amino)-triphenylamine. This is a low
molecular weight material, but due to its
sterically hindred ‘starburst’ architecture it
has a very high glass transition temperature and
does typically not crystallise when spincast from
solution, but forms a glassy film, like many
polymers. The glassy morphology has considerable
advantages for device applications; a tendency to
crystallise is a major problem with hole
transporting small molecules such as TPD. The
same structural theme was employed for the design
of electron transporting starburst- type
phenylquinoxalines (not shown here). Another
structural theme that can be used to suppress
crystallisation in non- polymeric materials is
the use of spiro- links between two (or more)
para-phenylene units, here exemplified by a
spiro- linked pair of quaterphenyls (sQP). Note
the cross- shaped 3- dimensional architecture of
spiro compounds that is difficult to sketch on
paper. oxTPD is clearly a low- molecular weight
compound, but via the oxetane functions that are
attached with flexible spacers it can be
crosslinked in- situ with the help of a suitable
(photo)initiator. The result is a highly
crosslinked, inert hole transporting film with no
crystallisation tendency that has been used
successfully in multilayer devices. NDSB Dendron
(G2) is a second generation, nitrogen- cored
distyryl benzene dendrimer. The core displays
visible absorption and emission, the meta- linked
dendronic sidegroups have a bandgap in the UV and

24
for the purposes of charge injection, transport,
and light emission can be considered as inert.

d.) synthetic metals:


The distinction between ‘organic semiconductors’
and ‘synthetic metals’ is somewhat arbitrary, as
the ‘synthetic metals’ shown here are in the
undoped state, when they display semiconducting
rather than metallic or quasimetallic properties.
Metallic properties are observed only after
chemical ‘doping’. PA (poly(acetylene)) is the
classic example, the (chance) discovery that
iodine- doped PA displays metallic conductivity
was a milestone discovery that earned the 2000
chemistry Nobel prize. Poly(diacetylene) (PDA)
has a widely tunable bandgap if substituted with
suitable sidegroups. Both PA and PDA are of
historic, but no longer of practical interest.
Poly(aniline) (PAni, here shown as ‘emeraldine
base’) and poly(3,4-ethylenedioxythiophene) PEDOT
are more modern developments. These are made
metallic by acid- rather than redox doping.
Water- based PAni and PEDOT preparations are now
commercially available. PEDOT that is acid- doped
with poly(styrene sulfonic acid) (PSS)
(PEDOT/PSS) is now very popular in the OLED
community to modify (or replace) the commonly
used transparent ITO anodes.

As a general remark on the above list of


materials, it is very important to keep in mind
that materials that nominally are the same often

25
show very different performance in devices.
Device performance can be very sensitive to low
levels of impurities and/or chemical defects, and
often, different chemical routes that lead to the
same material introduce different levels and
types of defects and/or impurities. Experience
shows that the Gilch route is superior to Wittig,
Horner, and Stille coupling for high quality
poly(arylenevinylene)s, while Suzuki coupling is
to be preferred over Yamamoto or Kovacic coupling
for poly(arylene)s. The highest regioregularity
in poly(alkylthiophene)s are achieved with the
help of Rieke Zinc. Companies or research groups
that are able to provide conjugated materials in
the quality required for practical devices are
few and far between. Even if chemistry is
‘ideal’, the same material can still display very
different properties when prepared in different
ways. The solvent and casting method used for
solution processing, or deposition rate, type and
temperature of substrate for evaporated films,
thermal treatment cycles, and the presence or
absence of even trace amounts of oxygen and/or
water can have a decisive impact on the resulting
device.

I.3 Excitations in organic semiconductors


In any semiconductor application, the material
will not be in its ground state. To transport
charge, and/or emit light, the semiconductor
needs to sustain excitations, and in the case of
charge transport, these excitations also need to

26
be mobile. In I.2, we will discuss the
fundamental properties of these excitations. In
I.3, we will see how excitations are generated
practically, and how they migrate.

I.3.a Polarons and excitons


When an electron is taken away from the HOMO or
added to the LUMO of a molecule, the resulting
molecule is termed a radical ion, namely a
radical cation for positive charge, and radical
anion for negative charge. The ‘radical’ refers
to the net spin the molecule will have due to the
unpaired remaining (or added) electron. After
removal or addition of the electron, molecular
orbitals and the positions of nuclei will respond
by a relaxation to a new position of minimum
energy. Radical ions are often called polarons,
(electron / hole polaron, respectively), in
analogy to the terminology in inorganic
semiconductors. However, again, the inorganic
polaron will be delocalised and have an
associated wavevector k describing its coherent
movement; the radical ion has not.

Due to the strong coupling between the charge


carrier and the local lattice relaxation,
removing an electron costs somewhat less in
energy than the HOMO suggests, and an electron
joining the molecule gains somewhat more energy
than the LUMO suggests. Instead, the required
energies are called the ionisation potential Ip,
and electron affinity Ea, respectively. Eq. I.1

27
illustrates that in the format of a chemical
reaction:

→ D+ + e −
I
D 
p
Eq. I.1:
A + e−  → A−
a E

Ip and Ea are conceptually closely related to


electrochemical Redox potentials. The main
difference is that Ip and Ea are defined with
respect to electrons in vacuum, while Redox
potentials are normalised with respect to
electrons in a reference electrode.
Experimentally it is much more common to measure
a molecules’ Redox potentials with a technique
called cyclic voltammetry (CV).

Note that in a metal, there is only one frontier


orbital, namely the Fermi level EF. The energy
required to remove an electron from a metal is
called work function Φ. As the frontier orbital
in a metal is the Fermi level, the naïve
expectation is Φ = - EF. This expectation,
however, is not met. The work function is a
surface property, while the Fermi level is a bulk
property. For a more detailed discussion, see
[T4, Ch. 18]. As a practical consequence, when
two metals are in contact, their Fermi levels
will equilibrate, but the work functions will
not. Instead, the work function will be that of
the metal that constitutes the particular surface
through which the electron leaves the metal.

28
As a practically important example for an organic
radical ion, a neutral and a positively charged
polythiophene segment are sketched in Fig. I.3.
Note how the missing electron (hole) leads to a
redistribution of the π-bonds and hence, to
different bond lengths, bond angles, and nuclear
positions.

S S S S

S S S S S

Hole
injection

(oxidation)

S S S S

S S S S S

Fig. I.3: A polythiophene segment and the derived


radical cation. See [T1, Ch. 14] for more
information.

Apart from polarons, the most important


excitation in an organic semiconductor is known
as exciton. This can be visualized as an electron
that is removed from the HOMO, but is placed into
the LUMO instead of being removed entirely. A
typical way of ‘lifting’ an electron from the
HOMO into the LUMO is via the absorption of a
photon. Note that the exciton is electrically
neutral. Alternatively, an exciton can result

29
from the combination of a hole- and an electron
polaron.

After exciton formation, the π electrons


redistribute into the excited π* orbitals, which
are also known as antibonding orbitals, as they
destabilise the molecule. The strong σ bonds are
crucial in keeping the molecule intact
nevertheless: The presence of a σ- bonded
backbone makes the difference between
photophysics and photochemistry. Again, the
excitation leads to a related structural
relaxation of the surrounding molecular geometry.
Fig. I.4 shows the geometric relaxation and
redistribution of electron density in an excited
phenylene- vinylene segment:

Excitation

Fig.I.4: The transition from an aromatic,


‘bonding’ π phenylene- vinylene system to a
quinoidal, ‘antibonding’ π* system on optical or
electrical excitation.

30
The ‘size’ of the exciton is about 3 repeat
units, or 10 nm, and the exciton has clearly
intramolecular, one- dimensional character. This
makes organic excitons Frenkel excitons, while in
inorganic semiconductors, excitons typically are
more delocalised Wannier excitons. Due to the
mutual attraction of electron and hole in the
exciton, and structural relaxation of the
molecule, the energy difference between the
excitonic state and the ground state is lower
than the difference between Ip and Ea suggests
(which in turn is lower than the difference
between HOMO and LUMO). This energy difference is
known as exciton binding energy Eb. Eb is larger
in Frenkel than in Wannier excitons; typical
organic (Frenkel) exciton binding energies are in
the range 0.2 to 0.5 eV. Note how considerable
ambiguity arises in the term ‘bandgap’ when
applied to organic semiconductors, which can mean
either the energy difference between LUMO and
HOMO, or Ip - Ea, or (Ip – Ea) – Eb.

Fig. I.5 gives an alternative, more schematic


representation of the electronic ground state,
radical ions (here called polarons), and excitons
in organic semiconductors.

31
LUMO

Ground state
HOMO

Hole Electron
polaron polaron

Singlet
Triplet exciton
exciton

Fig. I.5: Level diagrams for excitations in


organic semiconductors

Fig. I.5 shows two different types of exciton,


singlet and triplet excitons. The different types
of excitons result from the fact that electrons
as well as holes posses a spin. The quasiparticle
‘exciton’ has an overall wavefunction that
contains a spatial and a spin part, and the spin
part of the wavefunction results from a
combination of the respective electron- and hole
spins in a way that is consistent with the basic

32
rules of quantum mechanics, in particular the
Pauli principle. There is three ways in which
hole and electron spin can combine so that the
resulting overall spin part of the wavefunction
is symmetric under particle exchange, and has
total spin S = 1 - namely ↑↑>,↓↓>, and 1/√2(↑↓>
+ ↓↑>). Excitons with that property are called
triplet excitons. One combination of spins,
namely 1/√2(↑↓> - ↓↑>), results in a spin part
of the wavefunction that is antisymmetric under
particle exchange, and total spin S = 0. This
combination is called a singlet exciton. The
properties of excitons are summarised in table
I.3:

Table I.3

Spin state ket S Sz symmetric(+)/


antisymmetric(
-)

↑↑> 1 1 +

↓↓> 1 -1 +

1/√2(↑↓> +↓↑>) 1 0 +

1/√2(↑↓> -↓↑>) 0 0 -

Table I.3: The possible spin combinations (spin


state kets) of electron / hole pairs (excitons),
the resulting total spin, spin z- component, and

33
symmetry property under particle exchange.
Excitons with a spin ket symmetric under particle
exchange are triplet excitons, excitons with spin
ket antisymmetric under particle exchange are
singlet excitons.

I.3.b Light emission from organic molecules


In most simplistic terms, the absorption of a
photon that has generated an exciton on an
organic molecule can in some cases be reversed
(ignoring some intermediate steps to be discussed
below): The excited electron drops back from the
LUMO into the HOMO, emitting a photon in the
process. This phenomenon is known as
fluorescence. Since excitons can also be
generated electrically by the combination of
polarons; this paves the way to organic
electroluminescence (EL). First, we will discuss
fluorescence in some detail, using a framework
based on ground- state and excited state
molecular orbitals. Then, we will return to
discuss a striking difference between
fluorescence and EL, which is best understood in
the more schematic picture used in Fig. I.5.

Fluorescence in the molecular orbitals picture


The following diagram Fig. I.6 is used to discuss
fluorescence in the MO picture. The two curves in
the potential energy / distance diagram describe
the ground state and first excited state of a
chemical bond, with the equilibrium distance

34
being the bond length. Bond length is longer for
the excited state due to the presence of
π
*
antibonding orbitals. The horizontal lines
represent the vibrational states of the bonds.
Nuclear distances oscillate around the
equilibrium bond length within the limits of the
intersection of the horizontal line with the
potential curve. Vibrational levels are quantised
with a vibronic spacing in the order 0.1 eV ≅
1100 K; consequently, at room temperature, almost
all bonds are in the lowest vibronic level of the
ground state. From there, they may be lifted into
the first excited state by absorption of a
photon.

E
S1

S0

r
R0 R1

Fig. I.6: Energy diagram for molecules in ground-


and excited states, showing potential energy E
vs. nuclear distance. See e.g. [T6, Ch. 17]

35
The absorption process is governed by the Franck-
Condon principle: Electronic transitions are much
faster than nuclear rearrangement. Hence,
transitions always occur ‘vertically’ in the
diagram, from ground state equilibrium position
to the turning point of a vibrational mode of the
excited state that is close to the ground state
equilibrium distance. Transitions from the 0
vibronic level of the ground state into the 0 / 1
/ 2 /… vibronic level of the first excited state
are known as 0-0 / 0-1 / 0-2 /… transitions. The
relative intensity of these is controlled by the
overlap integral between the electronic ground
state / vibronic ground state (0 vibronic state)
wavefunction and excited state / vibronic state
0, 1, 2,… wavefunctions. The overlap integral can
be separated into an electronic and a vibronic
integral; the square of the vibronic integral is
known as the Franck-Condon factor of the
transition. A detailed discussion is in [T6, Ch.
17]. If the equilibrium bond lengths in the
ground- and first excited state were equal, the
0-0 transition would have unit Franck-Condon
factor, and all higher vibronic transitions would
have Franck-Condon factor zero, i.e. they would
be forbidden. In reality, bond lengths for
excited states are longer than for the ground
state, and higher vibronic transitions become
allowed, the so- called vibronic satellites.
Usually, however, the 0-0 transition will be the
most intense.

36
normalised fluorescence (x)
1 1

normalised absorbance (o)


0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
300 350 400 450 500 550 600
Wavelength [nm]

In the excited state, there will be a rapid


-12
(order 10 s), radiationless relaxation of the
excited state into its lowest vibronic level,
this is known as internal conversion. From there,
the photon may be re- emitted after typically 1
to 10 ns with a transition to the lowest vibronic
level of the ground state (0-0 transition), the
first vibrational state (0-1 transition), second
vibrational state (0-2 transition), etc. Vibronic
spacing is similar in ground- and excited state,
therefore these often appear like mirror images.
Fig. I.7 shows absorption- and fluorescence
spectra of perylene as an example. Perylene is a
molecule with pronounced vibronic satellites.
Less rigid molecules, e.g. para- phenylenes,
often do not show resolved vibronic sidebands,
but a single, broadened peak.
Fig. I.7: Absorption and emission spectra of
perylene dissolved in cyclohexane. Both spectra
are normalized to unit maximum peak. Note that
when spectra are represented against a wavelength
scale, absorption spectra appear on the left at

37
shorter wavelengths (higher energies), emission
spectra on the right (lower energies). When
represented against an energy or wavenumber
scale, the situation is reversed.

In Fig I.7, the 0-0 absorption and 0-0


fluorescence peak at (virtually) the same
wavelength, as it is expected from the previous
discussion. In many cases, however, 0-0 emission
is shifted by a few nm (order 5 to 10) to longer
wavelength (lower energy). This phenomenon is
known as Stokes shift. Stokes shift results from
the interaction of molecules with their
environment; we had so far considered isolated
molecules. No details here.
Either the onset of absorption, or the
intersection of absorption and emission spectra
in a normalised plot like Fig. I.7, is often
called the ‘optical bandgap’. In the absence of
Stokes shift, of course the intersection of
absorption- and emission spectra occurs at the
absorption / emission maximum.

Perylene has a fluorescence quantum yield ηFL =


0.94 in cyclohexane solution, and many modern
conjugated polymers exceed 50 % quantum yield
even in the solid state. However, some conjugated
polymers display rather low fluorescence quantum
yields.

38
In organic light emitting devices, light is
generated not by the absorption of a photon, but
by the combination of an electron- and a hole
polaron: An electron in the LUMO and an electron
vacancy (hole) in the HOMO can combine to form an
exciton. The fluorescence emitted from an
electrically generated exciton is called
electroluminescence (EL). However, there is a
fundamental difference between the formation of
excitons by absorption of light, and by
combination of polarons.

When light is absorbed, only singlet excitons are


formed, and all of them are in principle capable
of fluorescence. The respective transition is
denoted as S ÆS transition (singlet ground state
0 1

st
to singlet 1 excited state). Since the photon
carries a unit h of orbital angular momentum L,
0 1
S and S have orbital angular momentum quantum
numbers l differing by 1 (one). Inversely, the
S ÅS
0 1
fluorescence transition does fulfil the
selection rule ∆l = 1. The photon’s orbital
angular momentum provides the required angular
S ÆS
0 1
momentum for the transition; conversely,
the S ÅS transition provides the orbital angular
0 1

momentum needed for photon emission. Such a


transition is called dipole allowed.

When excitons are generated electrically, naively


only 1 in 4 excitons will be a singlet exciton
1
(S state). 3 out of 4 will be triplet excitons,

39
0
corresponding to the T state. This is simply
because there is 3 ways to combine 2 spins into a
triplet state and only one way to combine them as
0 0
a singlet. Both S and T have the same angular
momentum quantum number l, hence ∆l = 0 for the
S ÅT
0 0
transition: The triplet- to- singlet
transition is dipole forbidden. Triplets can not
emit fluorescent light, because the necessary
angular momentum h can not be supplied from the
orbital angular momentum difference between
excited and ground state. Obviously, it is bad
news for organic electroluminescence that
apparently only 1 in 4 electrically generated
excitons can emit light.

0
However, T does carry angular momentum in its
spin: Remember triplet excitons have total spin S
= 1, singlets 0. In some molecules, triplet spin
angular momentum can be transferred to a photon’s
orbital angular momentum. The conversion of spin
into orbital angular momentum is facilitated by
the interaction of the magnetic fields that both
electron spin and orbital angular momentum
generate. These fields interact with each other,
and interaction is proportional to the product of
orbital and spin angular momentum, L.S. This
interaction is therefore known as spin- orbit or
0
LS coupling. The emission of light from T via LS
coupling is known as phosphorescence. To have
strong LS coupling, we need to incorporate atoms
into our molecules that have filled atomic shells
with high orbital quantum number l. This will

40
generally be ‘heavy’ atoms (i.e., heavier than
carbon). Phosphorus may work – the clue is in the
name – but modern phosphors used in organic EL
are heavy metal organometallic complexes using
e.g. Iridium.

I.3.c Controlling the bandgap


In the previous chapter, we have seen how the
characteristic vibronic structure of organic
spectra arises. However, the overall location of
absorption- and emission of a semiconductor are
controlled by the size of its bandgap. The
perceived colour of light, in turn, is controlled
by the location of the emission band within the
visible spectrum. In particular for full- colour
displays, as well as for photovoltaics, the
control of the bandgap is a key requirement.
Synthetic chemistry is extremely versatile in
manipulating molecular architecture in a way that
allows bandgap tuning throughout the visible
spectrum. Bandgaps reduced or increased by
molecular engineering are often referred to as
redshifted and blueshifted, respectively. We will
discuss the two key approaches to bandgap
control: The ‘steric’ approach and the
‘electronic’ approach. In the latter approach, it
is in fact possible to manipulate Ip or Ea only,
leaving the respective other unchanged. It thus
becomes possible to design hole- or electron
transporting materials.

The steric control of bandgap

41
When π orbital overlap is improved by forcing
rings into a coplanar arrangement by chemical
bonds, bandgap will be reduced. Example:
Oligomers of para-phenylene (PPP), fluorene
(PFO), and methylated ladder-type poly(para-
phenylene) (mLPPP):
Table I.4
Number of Oligo-PP Oligofluore Oligo-mLPP
Benzene [eV] ne [eV]
Rings [eV]
3 4.44 3.70
4 4.25
5 4.15 3.18
6 4.03 3.56
7 3.00
8 3.43
9
10 3.35

Table I.4: Absorption maxima for oligo(para-


phenylene)s, Oligofluorenes, and oligo(ladder-
para- phenylene)s. Data compiled from [J Grimme,
M Kreyenschmidt, F Uckert, K Müllen, U Scherf,
Adv. Mater. 7, 292 (1995)] and [D Klaerner, R D
Miller, Macromolecules 31, 2007 (1998)].

42
It is evident that at a given number of rings,
the more planarized the backbone, the lower the
bandgap.

The electronic control of the bandgap.


By introducing either electron- withdrawing or
electron-donating chemical groups into a
conjugated molecule, electron affinity and
ionization potential will be affected, hence, the
bandgap will change. Such groups can be
introduced in two ways, namely as sidechains or
in the mainchain. We will discuss the following
examples: Alkoxy- sidechains attached to PPV
rings (MEH- PPV), CN- groups attached to alkoxy-
PPV vinylene bonds (CN- PPV, a case somewhat
intermediate between sidechain- and mainchain
modification) and fluorene copolymers (mainchain
modification).

MEH- PPV is an example for alkoxy- substituted


PPVs. Sidechains make it soluble in organic
solvents such as THF or chloroform. Sidechains
also somewhat isolate chain backbones from each
other in the solid film, which increases quantum
yield over unsubstituted PPV. An interesting
molecular engineering motif that was introduced
in the design of MEH- PPV is the use of a
branched sidechain. The branch constitutes a
chiral centre, MEH- PPV chains thus contain a
racemic mixture of right- and left handed
monomers. However, backbone conjugation is not
affected by the branch in the sidechain. This

43
gives MEH- PPV the photophysical properties of a
homopolymer, but crystallisation is suppressed as
in a random copolymer. As crystallisation often
is detrimental to fluorescence efficiency, this
was a welcome step forward in the development of
conjugated polymers. The alkoxy- linkage of MEH-
PPVs sidechains to the phenyl backbone ring also
changes the electronic structure of the backbone.
Alkoxy links have a tendency to donate electrons
to the backbone, which changes the shape and
location of the HOMO. As a result, emission is
redshifted compared to PPV, from green to orange.

The case of CN- PPV is somewhat intermediate


between sidechain- and mainchain modification. In
addition to the alkoxy- sidechains as in MEH-
PPV, highly electron- withdrawing cyano groups
are attached to the vinylene bonds. This leaves
the conjugated backbone highly electron
deficient, thus considerably increasing the
electron affinity (by about 0.6 eV [R1]). Due to
their high electron affinity, CN- PPVs are n-
type semiconductors. CN- PPVs typically are red
emitters.

Another approach to bandgap control is


copolymerization of different conjugated units
into the polymer backbone. Copolymers between
alkane- and alkoxy substituted PPV- type polymers
are discussed in [R1]. Here, we will focus on
copolymers of fluorene. Polyfluorenes display a
‘blue’ bandgap that is almost indifferent to the

44
type of sidechain attached. Note that sidechains
are attached pairwise at a position which itself
is not part of the conjugated backbone, and these
pairs are widely spaced, so that there is neither
steric nor electronic impact of the sidechains on
the backbone properties. However, strictly
alternating copolymers of fluorenes with
comonomers with different electronic properties
have been prepared, such as F8BT (alternating
copolymer between dioctyl fluorene with
benzothiadiazole) and F8T2 (alternating copolymer
of dioctyl fluorene with two thiophenes). For
both F8BT and F8T2, the resulting bandgap is
reduced, and they both emit in the green- yellow
region. The reduction of the bandgap has
different reasons: In the case of F8BT, the
benzothiadiazole comonomer has a higher electron
affinity Ea than fluorene, thus leading to a
polymer with higher Ea. In the case of F8T2, the
two thiophene groups have a lower ionization
potential Ip than fluorene, thus leading to a
polymer with lower Ip. Copolymerization thus does
not only allow control of the bandgap, but of
both Ip and Ea in a predictable manner, and a
large number of fluorene copolymers have been
synthesized and studied. For a review, see [R13].
Both polymers have found interesting
applications: Some of the most efficient organic
EL devices have been built from blends of a
minority amount of F8BT as electron injecting /
transporting material in hole injecting /
transporting polyfluorene host material. F8T2, on

45
the other hand, is an excellent material for p-
type OFETs.

I.4 Charge carrier injection


The injection of charge carriers is an issue of
immense practical importance for semiconductor
devices: Transistors require the injection of one
type of carrier from an electrode, and rather
fast transport of that carrier. Light emitting
devices require the injection of carriers of both
types from different electrodes. Photovoltaic
devices need to separate excitons and transport
the resulting carriers to opposite electrodes. It
is thus paramount that we discuss the factors
controlling carrier injection and transport.

Carrier injection from a metal electrode into a


semiconductor is controlled by the work function
Φ of the metal relative to the electron affinity
Ea of the semiconductor for electron injection,
and relative to the ionisation potential Ip of
the semiconductor for hole injection. A level
diagram as in Fig. I.13 is often used to
illustrate carrier injection e.g. into an organic
light emitting device (OLED). In a level diagram,
electrons minimise their energy by moving
downwards to lower energy; holes minimise their
energy by moving uphill because that implies an
electron going downwards.

46
Location Location

V a c u u m
0 level
Vacuum te
level

Ea
2.5 Cathode
2.8
Ca

4.2
Al
4.7 Ip
ITO 5.2 ITO
PPV
Energy [eV] th

Fig. I.13: Energy levels for a PPV layer


sandwiched between unlike electrodes. Indium tin
oxide (ITO) is a transparent metallic material
that is commonly used as anode for OLEDs. Left:
No bias voltage applied. Right: A voltage is
applied in forward bias.

Let us clearly state the assumptions that have to


be made when Fig. I.13 is used to discuss carrier
injection. Firstly, we assume the near complete
absence of dopant- induced charge carriers in the
semiconductor. In electrical engineering, undoped
materials are often called ‘insulators’ rather
than ‘semiconductors’ even when they have a small
bandgap. The absence of dopant- induced charge
carriers implies that bands remain straight at
all times - bands may tilt, but they will not
bend (show curvature). Thus, Fig. I.13 describes
a metal- insulator junction. Secondly, we assume
that the levels drawn in Fig. I.13 – metal work

47
functions, ionisation potential, and electron
affinity – are the same at the metal- organic
semiconductor interface as they are in the
isolated metal, or bulk semiconductor. Note that
due to the molecular nature of organic
semiconductors, there are no surface ‘dangling
bonds’ which do usually distort bulk energy
levels in inorganic semiconductors. However, in
metal- organic semiconductor contacts, interface
dipole moments often are present (sometimes,
deliberately engineered) which will modify
barriers. Also, the work function of a metal
coated with an organic semiconductor can be
different from that of the same metal with
respect to vacuum. An example is the gold-
pentacene interface, which displays a
surprisingly large energy barrier. These issues
are discussed e.g. in [N Koch, A Elschner, J
Schwartz, A Kahn, Appl. Phys. Lett. 82, 2281
(2003)]. In the presence of such effects, levels
in Fig. I.13 have to be re- drawn at a different
energy.

After these cautional remarks, let us inspect the


salient features of Fig. I.13, that illustrates
the level scheme at the example of PPV sandwiched
between ITO- and Al (or Ca) electrodes. On the
left, we see that a hole would have to move
downwards by 0.5 eV on injection from ITO to PPV.
However, holes will voluntarily move uphill
rather than downwards. The necessary step into
the ‘wrong’ (energetic) direction is called an

48
injection barrier, here of 0.5 eV for holes from
ITO into PPV. Electron injection from Ca or Al
into PPV needs to overcome a barrier of 0.3 or
1.7 eV for electron, respectively. On the other
hand, for electron injection from ITO, there
would be a large barrier of 2.2 eV. Thus, the use
of electrodes made from unlike metals defines a
forward and reverse bias for the OLED. To
minimise barriers, a high workfunction anode and
a low workfunction cathode are required. Table
I.7 lists the work functions of some metals, as
well as the highly doped semiconductor ITO, and
the synthetic metal PEDOT/PSS.

Table I.7

Metal Work function Φ [eV]

Cs 1.81
Ca 2.8
Mg 3.64
Al 4.25
Ag 4.3
Au 4.7
Cu 4.4
ITO ≈ 4.7

PEDOT/PSS ≈ 5.0…5.2

49
Table I.7: The work function of a number of
metals, most data from [T4, Table 18.1, Ch. 18].
High work function metals are known as ‘noble’
metals. Low work function metals are highly
reactive and usually need to be protected from
ambient atmosphere to avoid oxidation.

The right- hand part of Fig. I.13 shows the same


device (assuming a Ca cathode) under a forward
bias. Carriers can now overcome injection
barriers by tunnelling from the electrodes so far
into the device that energy at that location of
the tilted band is as low or lower than in the
electrode. Also, carriers can be injected
thermally. For barriers of 0.5 eV as shown here,
the current density in a device in forward bias
will usually be controlled by the injection of
carriers across the barrier (injection limited
current), rather than by the transport of
carriers across a device. The tunnelling current
density j(Vbias) is described by the equation of
Fowler and Nordheim (Fowler- Nordheim (F-N)
tunnelling):

3/ 2
(eq.I.4.1) C ( Vbias )2 exp[−B d∆V
j FN = ∆V ]
d Vbias

*
with B = 8π√(2m )/(2.96eh) with an effective
electron mass m*. For a detailed discussion, see
e.g. [T1, Ch. 12]. Note that jFN depends on
applied voltage Vbias and film thickness d only in

50
the combination Vbias/d, i.e. jFN scales with the
applied field E. Thermally activated or
thermionic injection is described by the
Richardson- Schottky (R-S) equation:

2
(eq. I.4.2) j RS = AT exp[−(∆V − Vm (E)) /kT ]

2 3
with A = 4πemkb /h , and Vm(E) describing the
field- dependend lowering of the injection
barrier by the attraction of the injected carrier
to its mirror charge; Vm(E) = √(eE/4πεε0).

Generally, R-S injection will dominate at low


fields, and F-N tunnelling at large fields that
are practically relevant for organic devices.
Experimental j(V) results on MEH- PPV diodes can
qualitatively be described by the F-N eq.n [I D
Parker, J. Appl. Phys. 75, 1656 (1993)], however,
the absolute current density is several orders of
magnitude lower than described by the F-N eq.n.
This is due to a large backflow of carriers from
the semiconductor back into the metal.

In the case of ∆V Æ 0, i.e. in the case of a


work function that is matched to the respective
semiconductor level (5.2 eV for hole injection
into PPV), the metal- semiconductor contact is
termed ohmic. For ohmic contacts the F-N and R-S
equations do no longer apply. Instead, current
density will be controlled by the transport of

51
charge carriers across the film. Carrier
transport is characterised by charge carrier
mobility µ, which plays a crucial role for
organic transistors. Practically, a contact can
be considered ohmic whenever carrier transport is
a more restrictive limit on current density than
carrier injection.

Having an ohmic contact means the problem of


injection is solved – ohmic contacts are
desirable. Considerable effort has therefore been
devoted to increase the workfunction of the
transparent ITO anode by a variety of
physicochemical treatment cycles [J S Kim, M
Granstrom, R H Friend, N Johansson, W R Salaneck,
R Daik, W J Feast, F Cacialli, J. Appl. Phys. 84,
6859 (1998)]. Recently, it has become common to
coat ITO with a thin film of the high work
function synthetic metal PEDOT/PSS [L
Groenendaal, F Jonas, D Freitag, H Pielartzik, J
R Reynolds, Adv. Mater. 12, 481 (2000)] (Φ = 5.2
eV). As cathodes, low workfunction materials such
as Ca are commonly used. These require protection
from ambient atmosphere, otherwise they would
rapidly degrade. This can be provided by
encapsulation, or by capping with a more stable
metal such as Al. Note that in such an electrode,
Fermi levels of Al and Ca will equilibrate, but
the work function will remain that of the metal
through which the electron attempts to leave,
which is Ca when we consider electron injection
into the organic semiconductor.

52
II.1 Synthetic Metals
According to a general theorem proven by Peierls,
there cannot be a one- dimensional metal (Peierls
transition). This seems to exclude the
possibility of metallic polymers.

However, inorganic semiconductors can be made


quasimetallic if a high level of doping is
introduced. In 1977, Heeger, MacDiarmid,
Shirakawa, and coworkers discovered a very
similar semiconductor Æ (quasi)metal transition
in the organic semiconductor trans- polyacetylene
(PA) [H Shirakawa, E J Louis, A G MacDiarmid, C K
Chiang, A J Heeger, Chem. Comm. No. 13, 578
(1977)]. Pure polyacetylene is a semiconductor
with a bandgap in the visible. However, when PA
was exposed to halogen vapours, these doped the
semiconductor and conductivity increased
significantly. Iodine vapours were most
effective, increasing conductivity by several
orders of magnitude up to 38 S/cm. Heeger,
MacDiarmid, and Shirakawa were awarded the 2000
Chemistry Nobel Prize for their discovery.
Besides PA, examples for synthetic metals are
polyaniline (PANi), polypyrrole (PPy), and
poly(3,4-ethylene dioxythiophene) (PEDOT or
PDOT).

Doping in synthetic metals is somewhat different


from doping in inorganic semiconductors, were
heteroatoms of different chemical valencies are

53
introduced into a host crystal lattice. In
synthetic metals, doping is by a chemical
reaction between semiconducting polymer and
dopant. Both redox- and acid/base reactions
between semiconductor and dopant have been used
successfully; typical examples are PA oxidation
with chloride or iodine, and acid doping of PAni
with camphorsulfonic acid, PPy with phosphoric
acid, and PEDOT with poly(styrene sulfonic acid)
(PSS). Quite high dopant concentrations of (1 to
50)% are typically used. Iodine oxidises
polyacetylene, thus removing π electrons from the
HOMO. This opens up mobile vacancies (holes) in
the previously completely full HOMO (‘valence
band’). Conductivity will depend on doping level;
for the highest doping levels PA can be as
5
conductive as Platinum (≈10 S/cm at room
temperature). Generally, synthetic metals become
more metal- like in their behaviour as dopant
concentration is increased. Above a critical
dopant level, most synthetic metals remain
conductive in the limit T Æ 0, and display NTC
behaviour at room temperature. Fig. II.1
summarises the behaviour of a number of synthetic
metals. Multiple entries for the same material
correspond to samples with different dopant
levels.

(Fig. 1 from A B Kaiser, Adv. Mater. 13, 927


(2001))

54
Fig. II.1: Room- temperature conductivity and
temperature behaviour of conductivity for several
synthetic metals at different temperatures. SWCN
= single- wall carbon nanotubes. Full squares:
NTC at RT and σ > 0 at T Æ 0 (‘proper metal’);
hollow squares: NTC at RT, but σ Æ 0 at T Æ 0;
full circles: PTC at RT but σ > 0 at T Æ 0;
hollow circles: PTC at RT and σ Æ 0 at T Æ 0
(‘proper semiconductor’). From A B Kaiser, Adv.
Mater. 13, 927 (2001).

To add to the ambiguity in classifying synthetic


metals as ‘proper’ metals or not, highly doped PA
and PAni show non- zero conductivity in the limit
of zero temperature and NTC behaviour around room
temperature, but PTC behaviour at low
temperatures, with a conductivity maximum in the
region (100…250) K. For a detailed discussion,
see [R18].

However, the focus of synthetic metals research


never was to resolve ambiguities in
classification, but the desire to arrive at
practical materials that can be used e.g. as
antistatic coatings, or electrodes. Like many
other mainchain- conjugated polymers, PA is
intractable and insoluble. The main obstacle
towards practical applications of synthetic
metals is to prepare highly conductive organic
materials in a versatile formulation.

55
II.2.a Water- based synthetic metals
Nowadays, mainly PANi and PEDOT are used as
synthetic metals. This is because both of them
are available as aqueous dispersions for
spincoating or ink jet printing. This makes them
easy to process into coatings, or for multilayer
film applications in conjunction with organic
semiconductors: Soluble organic semiconductors
are usually processed from organic solvents, in
which the water- based synthetic metals are not
soluble. The advent of water- based synthetic
metals can be seen as a second breakthrough in
addition to the initial discovery of metallic
polyacetylene. Soluble synthetic metals have
considerably larger potential for applications
than insoluble polyacetylene.

PANi is a very complex material, with several


Redox states. For a discussion, see [T2, Ch.7.9].
Here, we will discuss PEDOT, which typically is
acid- doped with poly(styrene sulphonic acid)
(PSS) to form the highly conductive PEDOT/PSS
complex. PEDOT in that case is synthesized from
EDOT monomer in an aqueous medium that already
contains PSS; in this way PEDOT/PSS become truly
inseparable:

56
Na 2 S 2 O 8
O O
H 2O
S n

SO 3 H

* n
*

SO 3 - SO 3 H SO 3 H SO 3 H SO 3 - SO 3 H

O
O O O O O

* S S S *
C S n
+ S S C +

O O O O O
O

Fig. II.2: The synthesis of PEDOT from EDOT in


aqueous medium in the presence of poly(styrene
sulphonic acid) (PSS) and Na2S2O8. Note how PSS
acts as proton donor (i.e., acid), and PEDOT as
proton acceptor (base).

Na2S2O8 acts as oxidizing agent for the coupling


of EDOT monomers. Note how the chemical bonding
pattern of EDOT/PEDOT changes under acid doping:
A C=C π bond opens up and the C bonds to an H
+

donated by the acid. As a result, there is:

57
- A net positive charge on the PEDOT chain that
will strongly attract the negative charge left on
the acid. Since this happens at many points along
the chain, PEDOT and PSS become closely
intertwined and float as a fine dispersion or
colloid in the aqueous solution, rather than
precipitate. They can not separate, and will not
dissolve in organic solvents.

- An unpaired π electron remains on the main


chain that is highly mobile along the chain.

PEDOT/PSS aqueous dispersion is commercially


available under the tradename ‘Baytron P’. From
Baytron P, thin, highly transparent, conductive
surface coatings can be prepared by spincasting
or dip- coating onto almost any surface. The
resulting conductivity is in the order 1 to 10
S/cm. A typical sheet resistance for spincast
PEDOT/PSS coatings is 1 MΩ/square.

II.2.b Applications of synthetic metals


PEDOT/PSS was originally developed as antistatic
coating for photographic films. Large- scale
processing of photographic film (‘development’)
requires the winding of films over reels in the
dark. This leads to static charging, and
discharge sparks may expose the film. A PEDOT/PSS
coating allows charges to disperse and thus
prevents the build- up of high voltages. For this
applications, rather low conductivities are

58
8 2
sufficient. Today, > 10 m of photographic film
is coated with PEDOT/PSS every year.

It should be obvious, however, that an elegant


and versatile off- the- shelve product like
water- based PEDOT/PSS will find a number of
other applications, in particular as it is sold
in high quality at a moderate price. Antistatic
coating of plastic films for packaging
microelectronics components is one of them. Also,
PEDOT/PSS has been used as counter- electrode for
anodised capacitors [F Larmat, J R Reynolds,
Synth. Met. 79, 229 (1996)].

PEDOT/PSS is now also increasingly used in


organic electronics. PEDOT/PSS displays a high
work function in the order (5 … 5.2) eV, which
according to I.4 qualifies it as good electrode
for hole injection into a semiconductor. It is
therefor used e.g. as contact electrodes and ink-
jet printed ‘wire’ in organic field effect
transistor circuits, and as anode (or anode
coating) in organic electroluminescent (EL)
devices. Also it is a useful electrode for
organic photovoltaics, were somewhat higher sheet
resistance is not a problem due to the generally
rather low current densities. We will discuss the
use of PEDOT/PSS repeatedly in the following
chapters. The excellent review [R6] is highly
recommended.

59
Recently, flexible and transparent PEDOT/PSS-
coated polyester sheets with sheet resistance in
the order 1kΩ/square have become available under
TM
the tradename ‘Agfa Orgacon ’. These sheets can
be patterned by an etching procedure which is
performed in a very similar way to conventional
photoresist patterning, although the chemistry is
rather different. This is the substrate onto
which future organic electronics and displays
will be prepared.

60
II.2 Organic field effect transistors
The transistor is the mainstay of solid state
electronics, and represents one of the seminal
inventions of the 20th century. All transistors
have three terminals. A current between two of
these three terminals can be controlled by the
input of the third terminal. Transistors can be
categorised into two ‘families’ according to how
this control is exercised: Those where the
control is through a current at the third
terminal, and those where the control is through
a voltage at the third terminal.

In current- controlled transistors, the three


terminals are called base (B), collector (C), and
emitter (E). B controls the current between C and
E. Such transistors are realised by doping the
same semiconductor (typically, silicon) with n-
and p- type dopants in different regions, either
in npn- or pnp configuration, and are also known
as bipolar transistors. Current- controlled
transistors are typically applied in voltage- or
current amplifiers. However, organic
semiconductors do not lend themselves for the
manufacture of bipolar transistors: Most organic
semiconductors will even without doping transport
only holes, or only electrons. An intrinsically
p- type material can not be transformed into an
electron transporter by n- doping.

61
Voltage- controlled transistors are also known as
field effect transistors (FETs). The three
terminals are called source (S), drain (D), and
gate (G). A voltage at G controls the S- D
current. The basic principle of the field effect
transistor dates back to 1930 [J E Lilienfeldt,
US Patent 1745175, 1930]. FETs are simpler in
design than bipolar transistors, require only one
type of carrier transport material (n or p), and
as we will see, not necessarily need to be doped
at all. Field effect transistors can be
manufactured in high integration densities and
are typically applied for logic gates in digital
electronics.

Due to the possible operation with only one type


of carrier, organic transistors invariably are
field effect transistors (organic FETs or OFETs;
sometimes also known as organic thin film
transistors, OTFTs), and only these shall be
discussed here. One of the major driving forces
behind organic FET (OFET) research is the idea to
make low- performance integrated circuits
completely from plastics on plastic substrates,
that are cheap enough to be discarded after
single use (‘disposable electronics’). A target
application is e.g. an electronic pricetag on a
food wrapper that a supermarket checkout can read
remotely. Such a tag must be cheaper than the
cost of a check- out assistant pulling a barcode
across a scanner. It is now regarded as realistic
that this cost barrier can eventually be

62
overcome. Also, higher added- value applications
are envisaged in combination with light emitting
organic semiconductors. Active- matrix addressed
display screens require two FETs and a capacitor
to address each pixel in a screen. Ultimately,
organic semiconductor technology aims to make the
transistors as well as the light emitting
material from organics.

II.2.a Description of organic FET operation


Fig. II.3 shows the principle design of an
organic thin film transistor (OTFT) with
electrical connections suitable for basic
measurements. A voltage between source (S) and
drain (D) called drain or source- drain voltage
VD attempts to drive a drain current ID through
the semiconducting transistor channel. However, a
drain current will only flow if there are mobile
charge carriers in the channel. The channel
semiconductor in OFETs is not chemically doped;
in fact, great care has to be taken to keep
impurity levels in OFET- grade semiconductors as
low as possible to avoid unintentional doping.

63
Vd

Source (Metal) Drain (Metal)


Semiconductor
Channel
Gate Insulator
Gate (Metal)

Substrate

Vg

Fig.II.3: A field effect transistor in the


‘bottom gate’, ‘top contact’ architecture.
Transistors may also be built with ‘bottom
contacts’ (S, D are placed directly on the
insulator), or completely the other way round
(‘top gate’ architecture).

A gate voltage VG will pull carriers out of the


source into the semiconducting channel. However,
since the gate dielectric is an insulator, these
carriers cannot reach the gate metal. Instead,
they will accumulate in the semiconductor near
the channel / insulator interface where they
represent mobile carriers known as an
accumulation layer. The channel semiconductor

64
thus gets doped by applying VG. However, other
than in chemical doping, this is quickly
reversible: Switching VG off will switch the
doping off. Hence, VG can switch the channel
conductivity. This behaviour is quantified by the
so- called on / off ratio, which in some cases
6
can be as high as 10 . Accordingly, ID at a given
VD will change with VG: The FET is an electronic
switch. The accumulation layer in an OFET is
generally very thin, about 5 nm or less [M A
Alam, A Dodabalapur, M R Pinto, IEEE Transactions
on Electron Devices, 44, 1332 (1997)]. This is
much thinner than the typical thickness of the
semiconducting film.

Note that an OFET is switched on by applying VD


and VG of equal polarity opposite to the sign of
the mobile carriers. Thus, OFETs can easily be
used to determine the type of carriers a
particular material sustains: If positive VG, VD
switch the transistor on, carriers are negative
(electrons), if negative VG, VD are required,
carriers are holes.

To characterise the gate- and source / drain


voltage dependent drain current in a field effect
transistor beyond the ‘on / off’ description, two
types of measurements may be carried out:

65
The output characteristic of a FET is the family
of curves ID vs VD with VG fixed during every
single VD scan (‘drain sweep’).

The transfer characteristic is the family of ID


vs VG curves with VD fixed during every single VG
scan (‘gate sweep’).

In principle, both of these contain the same


physical information, as long as a sufficient
number of gate voltages (drain voltages) are
scanned for the output (transfer)
characteristics. Practically, however, output
measurements will usually scan ID vs VD at a
number of gate voltages, while transfer
characteristics often will be taken at only two
fixed VD: One VD will be chosen to be as small as
practically possible; much smaller than the
maximum VG used for the scan; while the second VD
will be chosen to be larger than the maximum VG
used in the VG scan. These two regimes are known
as the linear and saturation regime,
respectively. This terminology will become clear
with the quantitative description of OFET
behaviour.

OFET operation can be described quantitatively if


a number of assumptions are made; leading to an
‘ideal’ scenario. These assumptions are: Channel
length shall be considerably longer than device
thickness (long channel limit), source and drain

66
make contact to the channel semiconductor without
contact resistance or injection barrier, the
semiconductor shall be trap- free and undoped,
and carrier mobility shall be independent of
gate- and drain voltages. Note that in reality,
few if any of these assumptions will be met!
Then, the following equation applies [T1, Ch.
14.2.2.1.2]:

Z
(eq.II.2.1.a) ID = Ciµ(VG − VT )VD for VD << VG ≥ VT
L

Z 2 for VD ≥ VG
(eq.II.2.1.b) ID = ISat = Ciµ(VG − VT )
2L

With Z/L the channel width/length, VT a threshold


voltage, µ the charge carrier mobility, and Ci
the capacitance per unit area of the insulator.
Ci is given by Ci = εε0/ti, with ε the insulator
dielectric constant and ti the insulator
thickness. Eqs. II.2.1 do not apply for the
subthreshold regime, VG < VT, which we will
not discuss here.

The dielectric constant ε describes by how much a


dielectric shields an applied field; the better
this shielding the more charges can accumulate on
the plates of a capacitor at a given voltage. It
is instructive to consider the product of Ci and
εε 0VG
VG that enters eq. II.2.1: CiVG = = εε 0 EG = DG (we
d
ignore VT for the moment), with EG the gate

67
electric field and DG the gate electric
displacement. DG has the dimensions of charge /
area, and corresponds to a surface charge density
QS of equal magnitude – this is the accumulation
layer. From this consideration, we see it is the
electric displacement DG that switches the
transistor on, not the gate voltage or gate
electric field directly.

While the precise derivation of eq. II.2.1 is


somewhat technical, it can easily be understood
qualitatively if for the moment, we leave the
threshold voltage VT aside. Z/L is an obvious
geometry factor. The conductivity of the channel
will be proportional to charge / area QS = CiVG.
Also, for VD << VG, ID rises linearly with VD
according to Ohm’s law – consequently, ID ∝
Z
/L
CiVGVD. Obviously, ID will also be proportional to
the velocity with which carriers move under a
given field, which is described by charge carrier
mobility µ - if we ignore VT, we now got eq.
II.2.1.a.

However, as VD rises into the same order as VG,


the accumulation layer will taper towards the
drain, because the voltage between gate and drain
is less than between gate and source.
Consequently, accumulation layer resistance
rises. When VD reaches VG, there is no field
between gate and drain, and we have accumulation
layer pinch- off. This is illustrated in Fig.

68
II.3. At VD = VG, the accumulation layer becomes
triangular and is on average half as ‘strong’ as
for VD Æ 0; this explains the factor 1/2 in eq.
II.2.1.b. When VD exceeds VG, the pinch- off point
moves away from the drain into the transistor
channel, and the carrier- depleted part of the
channel between pinch- off point and drain will
display very high resistance. Consequently, the
transistor displays drain current saturation for
VD ≥ VG, with ID,sat ∝ VG . The only way to increase
2

ID beyond the saturation is to increase VG.

Fig. II.3: Accumulation layer (black) in a field


effect transistor. Top: VG applied, VD = 0. The
accumulation layer is fully developed. Middle: VG
applied, VD = VG. The accumulation layer is
triangular and pinches off at the drain. Bottom:

69
VG applied, VD ≈ 2VG. The accumulation layer
pinches off in the middle of the channel.

This leaves only one important contribution to


eq. II.2.1 unexplained, that is the threshold
voltage VT. The theory of VT is intricate; also,
it is often found that VT changes in transistors
as a result of prolonged operation; this is
termed gate bias stress. Gate bias stress has
been studied e.g. by Katz et al. [H E Katz, X M
Hong, A Dodabalapur, R Sarpeshkar, J. Appl. Phys.
91, 1572 (2002)], who conclude that it may result
from both slowly orientating dipoles in the
insulator, or the presence of slowly mobile ions
(electret behaviour). We here will simply take VT
as an empirical constant.

Fig. II.4. shows examples for experimentally


obtained OFET output and transfer
characteristics.

[II.4.a = Fig. 3a of [H Sirringhaus, R J Wilson,


R H Friend, M Inbasekaran, W Wu, E P Woo, M
Grell, D D C Bradley, Appl. Phys. Lett. 77, 406
(2000)].

70
II.4.b

1.2
|ID |1/2 [(microAmp) 1/2 ]
1

0.8

0.6

0.4

0.2

0
0 10 20 30 40 50 60 70
|V G | [V]

[II.4.c = Fig. 3b of K Fujita, T Yasuda, T


Tsutsui, Appl. Phys. Lett. 82, 4373 (2003)]

Fig. II.4: a.) Output characteristic of an OFET


with F8T2 as active material. From [H
Sirringhaus, R J Wilson, R H Friend, M
Inbasekaran, W Wu, E P Woo, M Grell, D D C
Bradley, Appl. Phys. Lett. 77, 406 (2000)]. b.)
Plot of √|ID| vs |VG| with ID taken at VD = VG from
II.6.a. c.) Saturated (VD = -60V) transfer
characteristic of an OFET with pentacene as
active material. The same ID is plotted in two
ways: logID (left abscissa, open squares) and
√ID (right abscissa, full circles). From [K
Fujita, T Yasuda, T Tsutsui, Appl. Phys. Lett.
82, 4373 (2003)].

71
Charge carrier mobility can be determined from
experimental characteristics with the help of
eq.s II.2.1. A very robust method of doing so is
to plot I D vs. VG , with ID in the saturation

regime, VD > VG. This should result in a


Z
Ciµ )
1/2
straight line with a slope ( and intercept
2L
VT. Such a plot can be constructed in several
ways, both from output- and transfer
characteristics. Saturation current level can be
read for different gate voltages from the
saturated (flat) part of an output
characteristic. Alternatively, drain and gate
voltage may be connected to the same source /
measure unit; on ramping up voltage we scan ID(V
= VG = VD), i.e. saturation current at different
gate voltages, but always taken at VG = VD. In
such a measurement, we can directly see how ID

2
rises parabolically with V (ID VG/D ) [D M
Taylor, H L Gomes, A E Underhill, S Edge, P I
Clemenson, J. Phys. D: Appl. Phys. 24, 2032
(1991)]. The most straightforward method,
however, is the measurement of a saturated
transfer characteristic, i.e. a gate voltage scan
with a fixed drain voltage VD that is larger in
modulus than the gate voltage at all times. In
this way, it is ensured that drain current is
always saturated. Plotting √ID vs VG with VD >
VG gives direct access to µ and VT via slope and
intercept. Another important FET characteristic
is the on / off ratio, defined as I(VG = V, VD =
V)/I(VG = Voff, VD = V) with an operational voltage

72
V that corresponds to the voltage available in
the respective application, and Voff the gate
voltage that minimises the drain current; usually
Voff ≈ 0. On / off ratios typically are large
3 7
numbers roughly in the region 10 …10 , and can be
extracted most conveniently from a saturated
transfer characteristic when this is plotted with
the ID axis on a logarithmic scale.

We note that ‘saturated’ mobility often differs


from mobility at small drain voltage. The
transconductance gm is defined in the linear
regime of the output characteristic (VD << VG) as
the slope of ID with VG, and is related to
mobility via eq. II.2.3:

 ∂I  ZC
(eq. II.2.3) gm =  D  = i µVD
 ∂VG V <<V L
D G

gm may be extracted either from the linear (non-


saturated) regime of output characteristics, or
from a transfer characteristic taken with a drain
voltage smaller than the gate voltage. Usually, µ
will differ when measured in different regimes.

‘Speed’ of OFET circuitry


Depending on application, electronics may have to
switch at very fast speeds. We will now discuss
the factors that limit the speed of organic
ciruitry.

73
We had seen that short channel length L will
result in higher drain currents, but L is even
more important for switching speed. The maximum
speed at which a transistor circuit can operate
is limited by the time τ it takes carriers to
cross the FET channel. That is the time the
accumulation layer takes to be emptied of charges
through the drain after VG has been switched off.
At frequencies f > f0 = 1/τ, the output signal
(ID) drifts out of phase with the input signal
(VG), and switching amplitude decreases. τ is
given by eq.II.2.4:

(eq.II.2.4) 1 = f 0 ≈ µVD
τ L2

L enters the equation squared: Obviously the


transit time of carriers through the channel will
be proportional to L. Also, the lateral electric
field that pulls carriers across the channel is
proportional to 1/L. τ is known as the RC time of
the transistor, and can be calculated
alternatively from the channel resistance and
capacitance, with the same result eq.II.2.4.
Assuming VD = 10 V and µ = 10
-2 2
cm /Vs, f0 is 1 kHz
for L = 100µm, but 100 kHz for L = 10 µm. In
contrast, inorganic computer electronics work
with GHz frequencies. It is obvious that organic

74
electronics has to compete on price rather than
performance.

It should be noted with care that eq.II.2.4


represents an upper limit for f0. There may be
other capacitances than the accumulation layer in
the transistor that charge (discharge) when VG is
switched on (off). These parasitic (stray)
capacitances may for example arise from an
overlap of source / drain electrodes with the
gate electrode.

Switching speeds of organic transistor circuits


can be studied with the help of ring oscillators.
A ring oscillator consists of an odd number of
logic ‘NOT’ gates, which can be made of two FETs.
The NOT operation converts LOW (HIGH) input
voltage into HIGH (LOW) output voltage. The
output of each NOT gate is fed into the input of
the next, with the output of the last of an odd
number of NOT gates being fed back into the input
of the first NOT gate. Thus, a ‘ring’ is
completed that has no self- consistent state, see
Fig. II.5. Instead, output will oscillate between
HIGH and LOW with a frequency f that is related
to f0, the switching frequency of an individual
transistor. f will be smaller than f0, as more
than one transistor has to switch, and will
increase with the number of NOT gates or stages
of the oscillator.

75
NOT NOT

NOT

Fig. II.5: A ring oscillator (schematically). If


we assume the output of the bottom NOT gate were
LOW, the output of the top right NOT gate should
be HIGH, the output of the top left NOT gate
should be LOW, leading to HIGH output of the
bottom NOT gate – which is inconsistent with the
original assumption. The ring oscillator has no
self- consistent state and will oscillate between
HIGH and LOW with frequency f.

When the oscillating output is displayed on an


oscilloscope, f can be extracted. Often, the
inverse of f divided by (2 x number of NOT gates
= total number of transistors) is reported as
propagation delay per stage. In the absence of
stray capacitances, this equals τ = 1/f0. Ring
oscillators allow a dynamic measurement of
carrier mobility. In any case, f is a practically
relevant measure of circuit speed. Fast OFET ring
oscillators with f > 100 kHz have been
demonstrated by the group at Siemens [W Fix, A
Ullmann, J Ficker, W Clemens, Appl. Phys. Lett.
81, 1735 (2002)]. Ring oscillators are not purely
a diagnostic tool, but have important
applications. The group at 3M have demonstrated a
radiofrequency identification tag (rf ID tag)

76
that uses a pentacene- based organic ring
oscillator with a frequency of about 200 Hz [P F
Baude, D A Ender, M A Haase, T W Kelley, D V
Muyres, S D Theiss, Appl. Phys. Lett. 82, 3964
(2003)]. The rf ID tag has an antenna (LC
circuit) that absorbs an incoming radio signal
(frequency 13.6 MHz), and re- emits a signal of
the same frequency, but with an amplitude
modulation. The amplitude modulation is
facilitated by the ring oscillator periodically
shorting / re- engaging the antenna circuit.

II.2.b Requirements on OFET materials


‘Good’ OFETs should display high drain current at
low drain and gate voltages, without reliance in
optimized geometry factor; high on / off ratios,
which means drain current at VG = 0 should be
extremely low; and fast τ. This implies strong
requirements on all materials used in OFETs –
semiconductor, insulator, and metals. In the
field of organic semiconductors, the one property
that has been addressed most is charge carrier
mobility, and that is the one we will discuss in
detail. However, let us just list a few other
material requirements:

- Source metals should make ohmic (barrier- free)


contacts to the semiconductor.

- Gate insulators should be thin without


displaying pinholes (tricky for solution

77
processed organics!) or any other current
leakage, and should have a high dielectric
constant.

- organic semiconductors must be free of


unintentional dopants down to very low
concentrations, otherwise the transistor will not
switch off properly.

- organic semiconductors must be free of charge


carrier traps. Trapped carriers are no longer
mobile, which means part of the accumulation
layer DG cannot contribute to the drain current.

- The interface between insulator and


semiconductor is particularly important. Often,
inorganic oxides are used as gate insulators
(SiO2, Al2O3, Ta2O5). These are a found to improve
when they are modified by a self- assembling
monolayer.

Charge carrier mobility in organic


semiconductors.
Eq.s II.2.1 underscores the importance of high
carrier mobilities for good drain current at
moderate gate and drain voltages, and eq. II.2.4
shows the importance of high mobility for fast
transistor switching speeds. Hence, much effort
in synthetic chemistry and physical chemistry is
geared towards high mobility materials. In this

78
section, we will discuss several aspects of
charge carrier mobility in organic
semiconductors.

Carrier mobilities can very strongly depend on


morphology, which in turn may change as a result
of different preparation conditions. Mobility
also usually rises with charge carrier density.
Therefore, there is no such thing as ‘the’
mobility of a material with given chemical
formula. In fact, Fig. 14.13 in [T1] shows for a
few materials the evolution of carrier mobility
with time as a result of improved preparation
technique. Hence, a wide range of data may be
quoted for the same material. This is meant as a
caution before showing a few examples in the
following table:

Table III.1

+ -
Materia Morphology µ h /e
l 2
[cm /Vs]
+
Pentace Large single crystal, 400 h
ne vapour deposited, c-axis,
low T
+
Pentace spincoated 0.001 h
ne
+
Pentace Evaporated onto 0.002…0. h
ne substrate, substrate at 03
room temperature (differe

79
nt
studies)
+
Pentace Evaporated onto 0.62…5 h
ne substrate, substrate at
120 C
+
F8T2 Aligned liquid crystal h
Parallel to alignment 0.01…0.0
Orthogonal to alignment 2

≈ 0.0015
-7 -4 +
PPV Several 10 …10 h
morphologies/methods
+
PAT Regioregular, face-on 0.001 h
substrate
+
PAT Regioregular, edge-on 0.1 h
substrate
+
6T polycrystalline 0.02 h
+
6T Single crystal 0.1 h
+
DH6T 6T with (α,ω)-dihexyl 0.05 h
endchains
-5 -
C60 Under air < 10 e
-
C60 In ultrahigh vacuum 0.08 e
-
F8BT Spincast from toluene ≈10 e
-3

Table II.1: Mobilities at room temperature. More


data can be found in [T3], table 8.1, for organic
crystals, and in [T1], Ch. 14 tables 14.1, 14.2,
14.3., and in [C D Dimitrakopoulos, D J Mascaro,
IBM J. Res. & Dev. 45, 11 (2001)].

80
We first note that typical organic mobilities are
much lower than the mobilities of carriers in
crystalline inorganic semiconductors. In
crystalline inorganic semiconductors, transport
is coherent or band- like with carrier mobilities
in the region of several hundred or even thousand
µ
2
cm /Vs (typically higher for Also, will
decrease with increasing temperature. This is the
hallmark of band- like transport, which in
exceptional cases can be observed in single
crystals of organic semiconductors at low
temperatures, as well – e.g., Karl et al. found µ
≈ 400 cm /Vs in Pentacene at low temperature [N
2

Karl, J Marktanner, R Stehle, W Warta, Synth.


Met. 41-43, 2473 (1991)]. However, even in single
crystals, transport will undergo a transition
from band- like to incoherent hopping transport
when thermal fluctuations (phonons) become
appreciable in magnitude compared to the
crystal’s binding forces. Organic crystals have
considerably weaker binding forces than inorganic
semiconductors (van der Waals forces vs. covalent
bonds), and will generally undergo the band- to-
hopping transport transition below room
temperature. Practical organic semiconductor
devices at ambient temperatures will therefor
always operate in the slower, hopping- type
transport regime. The fastest possible mobilities
2
in that regime will be in the order (1…10) cm /Vs
even for the best ordered (single crystalline)
materials, and mobility will increase with

81
temperature. Amorphous organic semiconductors
will display hopping- type transport at all
temperatures. A detailed discussion of mobility
limits is in [C D Dimitrakopoulos, D J Mascaro,
IBM J. Res. & Dev. 45, 11 (2001)].

The minimum mobility required for OFET


applications will depend on application as well
as on device geometry, however, materials with
-3 2
mobilities less than 10 cm /Vs will generally be
considered as unsuitable for organic transistor
applications. This excludes phenylene- vinylene
polymers (PPV and its derivatives) from
transistor applications.

The high mobilities found for some single


crystals, or for certain thin films carefully
deposited onto specific substrates that are not
useful for device fabrication, are not always of
value for applications. A strong sensitivity to
oxygen is not acceptable, because encapsulation
is not an option for low cost devices. As a rule
of thumb, thiophene or thiophene- containing
materials often display high mobilities, both for
low molecular weight materials (6T, DH6T) or
polymers (PAT, PTV, F8T2). However, poly(alkyl
thiophene)s are generally sensitive to oxygen;
low molecular weight thiophenes or thiophene
copolymers like F8T2 are not that sensitive.

82
Low molecular weight (vapour- deposited)
thiophenes are less susceptible to oxygen doping,
however, their performance is now eclipsed by
pentacene that has been prepared with room-
2
temperature mobility of 5 cm /Vs [T W Kelley, D V
Muyres, P F Baude, T P Smith, T D Jones, Mat.
Res. Soc. Symp. Proc. 71, L6.5.1 (2003)]. This is
a mobility as good as that of amorphous silicon,
and is believed to be close to the theoretical
limit for hopping- type transport.

As an alternative to thiophene- based polymers,


derivatives of poly(triaryl amine) (PTAA) have
been developed, most prominently by the group at
Avecia [J Veres, S Ogier, S Leeming, B Brown, D
Cupertino, Mat. Res. Symp. Proc. 708, BB8.7
(2002)]. These display somewhat lower mobilities
than the best (i.e., regioregular) poly(alkyl
thiophene)s, but can be operated under air.
Fluorene- thiophene copolymers (F8T2) also
exhibit good mobilities and moderate tolerance to
oxygen.

The Bässler disorder model [T1, Ch.12] gives some


insight into the design criteria for high
mobility materials in the hopping regime.
Bässler’s equation contains the energetic
disorder parameter σ , which describes how much
2

the frontier orbitals (HOMOs for hole


transporters) of individual molecules deviate
from the average HOMO (more precisely, σ is the

83
standard deviation of the assumed Gaussian
density of transport states). Higher σ
2
leads to
reduced mobility, it is thus paramount to
minimise energetic disorder by designing
molecules to have little scatter in their HOMO
levels. Disorder is introduced by conformational
degrees of freedom, such as possible rotations
around single bonds; and molecular dipoles. Note
that pentacene has no dipoles and no degree of
freedom in its molecular architecture, which goes
some way towards explaining its record hopping
mobility. Also, evaporated pentacene is
crystalline, thus further reducing energetic and
positional disorder. For some reason, in the case
of pentacene, the inevitable grain boundaries do
not seem to compromise transport.

Introductory reviews on OFETs are [R7-R9].

II.3 Organic light emitting devices


The huge optical display market is currently
dominated by cathode ray tubes (CRTs) and liquid
crystal displays (LCDs). Both technologies are
rather dated and have severe drawbacks, e.g. CRTs
are bulky and consume large amounts of energy,
and LCDs are passive devices that require
backlighting and often suffer from poor viewing
angles, and switching speeds not fast enough for
the display of moving images. State- of- the- art
LCD screens have these problems under control,
but this comes at the expense of considerable

84
complexity in device manufacture, which means a
high price.

Replacing current displays with organic light


emitting devices (OLEDs) is tempting commercially
as well as technologically. Here, we shall
discuss the history and current state- of- the-
art of OLEDs.

The first electroluminescent (EL) polymer devices


were based on PVK, as reported in a series of
papers by Partridge [R H Partridge, Polymer 24,
733; Polymer 24, 739; Polymer 24, 748; Polymer
24, 755, 1983]. However, these devices were based
on sidechain- conjugated polymer with very poor
electron injection and transport, and did not
attract much interest. In 1987, Tang and van
Slyke reported much more efficient organic EL
based on evaporated low molecular weight
materials [C W Tang, S A van Slyke, Appl. Phys.
Lett. 51, 913 (1987)]. Finally, the discovery of
EL from a mainchain conjugated polymer [J H
Burroughes, D D C Bradley, A R Brown, R N Marks,
K Mackay, R H Friend, P L Burns, A B Holmes,
Nature 347, 539 (1990)] generated outstanding
interest. Throughout the 1990s, the drive to
develop these discoveries into marketable
technology provided most of the momentum in
organic semiconductor research.

II.3.a Overview over basic processes

85
In fluorescence, excitons are created by the
absorption of light, while in EL excitons are
created by electron and hole polaron ‘capture’.
Polarons first have to be injected from the
electrodes, and migrate towards each other. They
then form an exciton that sometimes can decay
under the emission of light. Fig. II.6 shows the
basic architecture of an organic light emitting
device (OLED). The variety of electrical and
photophysical processes involved are summarized
in the chart Fig. II.7.

As a direct consequence of the vertical device


architecture, one of the key requirements on OFET
materials can be relaxed considerably for OLED
semiconductors: Assuming that carrier traps can
be avoided, charge carrier mobility is by far not
as important an issue for OLEDs as it is for
OFETs. This is simply because of the much shorter
distances L carriers need to travel across a
device to the other electrode. In the planar OFET
geometry, typical channel lengths are several µm,
while in the vertical OLED architecture, layer
thickness is in the order 100 nm. Apart from
specialities, such as organic lasers or OLEDs
that emit short pulses, we can therefore use
conjugated materials with much lower carrier
mobilities in OLEDs than in OFETs. The PPV family
of OLED materials is a case in hand.

86
Electric drive circuit

Cathode Organic
Semiconductor

Anode

Substrate

Light out

Fig. II.6: OLED architecture

Anode Cathode
Hole Electron
Electron injection injection Hole
leakage leakage
current current
Coulomb Capture

exciton formation
"1/4" "3/4"
Intersystem crossing
Singlet Triplet
S=0 S=1
Triplet- triplet
annihilation
Radiative
decay (allowed) Radiative
non- decay
radiative non-
radiative (weakly
decay allowed)
decay
Ground state
Fig. II.7: A chart describing the formation and
decay of excitons in organic EL devices. Adapted

87
from [D D C Bradley, Current Opinion in Solid
State & Materials Research 1, 789 (1996)]].

However, to operate OLEDs as efficiently as


possible, all the processes shown in Fig. III.11
need to be optimised. This introduces a number of
requirements on materials and device engineering
that are not an issue for OFET materials. We will
discuss them in the following.

II.3.b Bipolar carrier injection


The basic physics of carrier injection from
electrodes into organic semiconductors were
discussed in I.4. The key feature that sets apart
an OLED from an OFET is that an OFET can operate
with one type of carriers, while an OLED always
requires both electrons and holes to be injected,
otherwise no excitons can form. Ideally, we wish
to have ohmic injection of holes at the anode,
and electrons at the cathode simultaneously. As a
consequence, OLEDs will use unlike metals for
anode and cathodes, namely a high work function
anode and a low work function cathode, while
OFETs usually use the same metal for source and
drain electrodes. However, it will still be
difficult to achieve ohmic (i.e., barrier free)
injection at both electrodes for a given organic
semiconductor. Even if that could be achieved,
the bandgap and thus emission colour of the
device would be defined by the work function
difference of the metals used, and we could never

88
make full colour displays. Materials chemistry
alone cannot overcome this hurdle.

The breakthrough towards efficient organic EL


devices came from a device engineering, rather
than materials development, approach. Tang and
van Slyke from the Kodak group have manufactured
a bilayer device consisting of a low ionisation
potential, hole transporting diamine layer and a
high electron affinity, electron transporting
Alq3 layer, which is also an efficient green
emitter [C W Tang, S A van Slyke, Appl. Phys.
Lett. 51, 913 (1987)]. Their idea has been widely
adapted and modified since. Fig. II.8 shows the
level diagram for a (fictitious) double- layer
device consisting of a hole- transporting layer
(HTL) and electron- transporting layer (ETL).

89
Vacuum level
0

2.7

3.2
3.7
2.5 eV
Mg
2.5 eV
4.7
ITO
5.2
5.7
Energy [eV]

Fig. II.8: Level diagram of a fictitious double


layer device, using ITO anode and Mg cathode.

Both layers are assumed to have a bandgap (Ip -


Ea) = 2.5 eV, however, the HTL has lower Ip than
the ETL, and the ETL has higher Ea than the HTL.
It is immediately obvious that a single layer
device using either HTL- or ETL alone would
necessarily have one large (1 eV) injection
barrier. In the double layer architecture both
barriers are moderate (0.5 eV) if the device is
addressed in forward bias (ITO as anode, Mg as
cathode).

90
In addition to the injection barrier, both holes
and electrons will encounter an internal barrier
at the HTL/ETL interface, which is also called a
heterojunction. Note, however, that
heterojunctions in organic devices are quite
different from those in inorganic devices due to
the lack of (deliberate) doping. The internal
barrier is not detrimental to device performance.
Instead, it can help to improve the balance
between electron- and hole currents. Assuming a
slightly smaller injection barrier for holes than
for electrons, or higher hole- than electron
mobility, even in a bilayer device we would
expect a carrier imbalance with a larger hole-
than electron current. However, since holes will
encounter an internal barrier, they will not
simply cross the device and leave at the cathode
as a ‘blind’ leakage current. Instead, they will
accumulate at the interface, where they represent
a positive space charge. The effect of the field
resulting from that space charge is to improve
charge carrier balance: Firstly, it will impede
the further injection of majority carriers
(holes) from the anode, and secondly, it will
enhance the injection of minority carriers
(electrons) from the cathode. Also, excitons will
form at the internal interface, far away from the
electrodes. Cathodes in particular have been
associated with exciton ‘quenching’ (i.e.,
radiationless exciton decay); this is avoided by
placing exciton formation in the centre of the
device rather than close to the cathode.

91
As Tang and van Slyke used small molecules,
bilayers could readily be manufactured by
subsequent evaporation. This approach has been
extended to sophisticated multilayer
architectures, e.g. by the group of Kido in
Yamagata. They have demonstrated some of the
brightest and most efficient OLEDs to date
2
(140,000 Cd/m and 7.1% external quantum
efficiency) [J Kido, T Matsumoto, Appl. Phys.
Lett. 73, 2866 (1998)]. An interesting approach
to highly efficient, low- voltage operated
devices is the use of doped charge- injection
layers, that transform the characteristic of
carrier injection from a tunneling junction
(‘metal- insulator contact’) to a Schottky
junction (‘metal- semiconductor contact’). Often,
thin layers of LiF are evaporated as n- dopant
between the electron transporting organic
semiconductor and the metal cathode. An example
where both hole- and electron injection into an
OLED are aided by dopants is in [J Blochwitz, M
Pfeiffer, M Hofmann, K Leo, Synth. Met. 127, 169
(2002)]. A concept closely related is that of the
light- emitting electrochemical cell, where
mobile carriers are introduced in the form of a
salt that dissociates into mobile ions.

With polymeric organic semiconductors, vapour


deposition is not an option, devices have to be
prepared by spincasting instead. Multilayer
architectures are harder to realise with

92
spincasting than with vapour deposition, because
of the need for ‘orthogonal’ solubilities. To
sidestep solubility problems, in principle a
precursor route may be employed, where the first
layer is prepared from a soluble precursor
polymer that then is converted in- situ into a
conjugated, and completely insoluble polymer.
This has been successfully employed for hole-
transporting PPV / electron transporting CN- PPV
double layer polymer OLEDs [H Becker, S E Burns,
R H Friend, Phys. Rev. B 55, 1 (1997)]. However,
the precursor route requires lengthy in- situ
thermal conversion under high vacuum and has
generally fallen out of favour with the advent of
soluble conjugated polymers.

Recently, a very favourable approach has emerged


that combines the ease of injection into a double
layer device with the simplicity of solution
processing. In that approach, a single layer of a
blend of a hole- transporting and an electron-
transporting conjugated polymer, namely
poly(dioctyl fluorene) (PFO) and F8BT, is
spincast in one single preparation step. As
spincasting implies the very rapid formation of a
solid film from solution, the two polymers have
little time to phase separate and a solid film
may result wherein both polymers remain
intimately mixed. Such a mixture has been termed
a ‘bulk heterojunction’, and the preparation and
morphology control of hole / electron
transporting blends is the focus of much current

93
research, mainly with a view to photovoltaic
applications of organic semiconductors [J J
Dittmer, E A Marseglia, R H Friend, Adv. Mater.
12, 1270 (2000)]. Holes are injected and
transported into the (majority component) PFO,
but can be transferred easily to F8BT, as it has
similar ionisation potential. However, F8BT has
poor hole mobility due to hole- specific traps.
Instead, it has rather high electron affinity and
displays comparatively good (albeit dispersive)
electron transport [A J Campbell, D D C Bradley,
H Antoniadis, Appl. Phys. Lett. 79, 2133 (2001)].
Thus electrons are mobile on the F8BT chain until
they encounter a trapped hole. With some further
device improvements, highly efficient (4.1 cd/A)
and low onset voltage (≈ 3V) OLEDs have been
prepared from such blends [J Morgado, R H Friend,
F Cacialli, Appl. Phys. Lett. 80, 2436 (2002)].

II.3.c Exciton formation


When both hole- and electron polarons have been
injected into a device, and these drift towards
each other under the applied voltage, they will
combine into excitons that may emit light. The
physics of this process was discussed in I.4.b.

At first sight, it appears that exciton formation


in multilayer architectures is hindered by the
internal barrier that carriers of either type
encounter at the HTL/ETL interface. However, this
is generally not the case. Excitons in organic

94
semiconductors generally display exciton binding
energies Eb of a few tenths of an eV [J L Bredas,
J Cornil, A J Heeger, Adv. Mater. 8, 447 (1996)].
When a carrier has to overcome an internal
barrier to form an exciton, this may require a
certain amount of energy; however, on exciton
formation, Eb is instantly ‘refunded’ –
effectively, the internal barrier is reduced by
Eb. Thus, majority carriers remain stuck at an
internal barrier and redistribute the internal
field in the favourable way discussed above,
until a minority carrier arrives at the
interface. As soon as a minority carrier is
available, exciton formation is then helped by
the effective barrier reduction Eb. High Eb also
stabilizes excitons against dissociation and non-
radiative decay. In ‘bulk heterojunction’ blends,
one carrier has to transfer from one chain to
another to form an exciton. This will be the type
of carrier for which the energy level offset of
either the ionisation potentials (|∆Ip|) or the
electron affinities (|∆Ea|) is smaller. The
smaller of the two offsets (min (|∆Ip|, |∆Ea|))
defines the energetic cost of carrier transfer.
Two very different scenarios emerge for the case
min (|∆Ip|, |∆Ea|) < Eb as opposed to min
(|∆Ip|, |∆Ea|) > Eb. In the former case,
formation of excitons from polarons will be
favoured, while in the latter case, the
dissociation of existing excitons into polarons
will be preferred. In the case of F8 / F8BT
blends, that had been introduced previously,

95
exciton formation is clearly favoured, and such
blends are useful for OLED applications. In other
hole / electron transport material blends,
exciton dissociation is favoured, which makes
such blends attractive for use in photovoltaic
devices. Examples include blends of F8BT with a
poly(fluorene-alt-triarylamine) [J J M Halls, A C
Arias, J D MacKenzie, W Wu, M Inbasekaran, E P
Woo, R H Friend, Adv. Mater. 12, 498 (2000)], and
blends of poly(alkyl thiophene) and perylene
tetracarboxyl diimide, [J J Dittmer, E A
Marseglia, R H Friend, Adv. Mater. 12, 1270
(2000)].

While measurements of |∆Ip|, |∆Ea|, and Eb with


sufficient precision to predict exciton formation
or dissociation are usually not available, there
is a simple experimental approach to decide which
is the case. When at least one of the blend
components is highly fluorescent on its own, but
fluorescence is quenched in the blend, this
indicates that in that materials combination,
excitons are split, and it may be useful for
photovoltaics, but not for electroluminescence.
II.3.d Optimising OLED efficiency
The discussion so far outlines the strategy
towards OLED devices with balanced carrier
injection and quantitative exciton formation that
can be driven at low voltage. The (formidable)
challenge that then remains is to maximise the
amount of light generated from the excitons. It
is obvious that we require a material with a high

96
luminescence quantum yield. However, the
formation of normally non- emissive triplet
excitons presents an unwanted limit on OLED
efficiency, and several approaches to overcome
this have been explored.

Generally, ηEL and ηFL are related via

σS σT
(eq. III.4.1) ηEL = η
σ S σ T + 3 PL

with σS/T the polaron capture cross- section for


singlet- and triplet exciton formation,
respectively. The naive assumption σS = σT leads
to ηEL = /4 ηPL. This issue was first raised in
1

I.3.a. We will now see how this limitation to


device efficiency can be tackled.

Enhanced singlet exciton formation


While some experimental studies based on
comparisons of EL and PL quantum efficiencies
appeared to confirm ηEL = /4 ηPL [M A Baldo, D F
1

O’Brien, M E Thompson, S R Forrest, Phys. Rev. B


60, 14422 (1999)], other studies are adamant they
found high EL quantum efficiencies consistent
with a singlet : triplet formation ratio ≈ 1:1 [Y
Cao, I D Parker, G Yu, C Zhang, A J Heeger,
Nature 397 (6718), 414 (1999)], implying σS ≈ 3σT.
To determine singlet / triplet formation ratios

97
directly rather than inferring them from EL / PL
efficiencies, Vardeny et al. [M Wohlgenannt, K
Tandon, S Mazumdar, S Ramasesha, Z V Vardeny,
Nature 409 (6819), 494 (2001)] carried out a
systematic magnetic resonance study on a number
of organic semiconductors with bandgaps in the
visible. They found that σS/σT was indeed
generally larger than 1, namely between 2 and 5
for different materials. σS/σT ≈ (2 … 5)
corresponds to ηEL ≈ (0.4 … 0.6) ηPL, instead of
0.25 ηPL. Vardeny et al. then studied an oligomer
series [M Wohlgenannt, X M Jiang, Z V Vardeny, R
A J Janssen, Phys. Rev. Lett. 88, art. no. 197401
(2002)], finding σS/σT to increase with
conjugation length. This implies larger σS/σT for
polymeric than for low molecular weight organic
semiconductors – a finding that potentially can
influence the future direction of an entire
industry.

Indeed, Friend et al. found a marked violation of


the naïve 1 singlet / 3 triplet rule for a
polymeric organic semiconductor, but not for a
low molecular weight analogue [J S Wilson, A S
Dhoot, A J A B Seeley, M S Khan, A Köhler, R H
Friend, Nature 413 (6858), 828 (2001)]. They
interpret this as the result of the relatively
long quantum coherence in a polymeric organic
semiconductor. Oppositely charged polarons ‘feel’
their Coulomb attraction with or without quantum
coherence. However, in polymers we have quantum

98
coherence over one conjugation length;
consequently polarons can ‘sense’ each others’
spin from a relatively long distance. Thus, if
singlet formation is preferred over triplet
formation, polymers have a chance of avoiding
triplet formation while electron and hole are
still widely separated, i.e. only weakly bound
electrostatically. In small molecules, quantum
coherence ends at the end of a molecule. Once a
hole and an electron are on the same molecule, it
is too late to avoid exciton formation,
regardless of spin statistics.

Electrophosphorescence
As alternative to the enhanced singlet formation
cross- section in polymers, in particular the
low- molecular weight OLED community has
developed the concept of ‘harvesting’ triplets
for light emission by using phosphorescence.

In a typical electrophosphorescent device, a wide


bandgap host semiconductor is ‘doped’ with a
small percentage of a phosphorescent emitter. The
excitation is transferred from the ‘host’ to the
‘guest’ via excitonic energy transfer. Forrest et
al. at Princeton have developed a range of green,
yellow, orange and red organoiridium complexes [S
Lamansky, P Djurovich, D Murphy, F Abdel-Razzaq,
H E Lee, C Adachi, P E Burrows, S R Forrest, M E
Thompson, JACS 123, 4304 (2001)], which are
exemplified by the particularly efficient red
phosphor btp2Ir(acac). When doped into a wide

99
bandgap host, electrophosphorescence with > 80%
internal quantum efficiency and 60 lm/W is
observed [C Adachi, M A Baldo, M E Thompson, S R
Forrest, J. Appl. Phys. 90, 5048]. Using pure
btp2Ir(acac) without host matrix had resulted in
less efficient devices due to triplet- triplet
annihilation [C Adachi, M A Baldo, S R Forrest, S
Lamansky, M E Thompson, R C Kwong, Appl. Phys.
Lett. 78, 1622 (2001)]. In particular in the red,
electrophosphorescence is an attractive approach.
Due to the response characteristics of the human
eye, red dyes must show very narrow emission
peaks, otherwise colour purity will be
compromised. Iridium- based phosphors display
considerably narrower emission bands than typical
fluorescent dyes, and are thus particularly
useful as red emitters. These phosphors also
display relatively short triplet lifetimes (4µs),
which means they are fit for use in video rate
display systems, and do avoid problems associated
with triplet- triplet annihilation at high
brightness, which had been encountered with
longer lifetime phosphors.

Another phenomenon specific to


electrophosphorescent devices is that relatively
long- lived excitons may diffuse for rather long
distances during their lifetime.
Electrophosphorescent devices therefore often
have exciton- blocking layers to prevent triplet
excitons to diffuse to the electrodes where they
would be quenched. When controlled carefully in

100
sophisticated multilayer architectures, exciton
diffusion may be used to achieve efficient white
emission [B W D'Andrade, M E Thompson, S R
Forrest, Adv. Mater. 14, 147 (2002)]. However,
the manufacture of such devices is highly
complex.

Electrophosphorescence is ambitious for blue


emission due to the need for a high bandgap host
semiconductor. In some cases, thermally assisted
‘uphill’ or endothermic transfer from a smaller
bandgap host to a higher bandgap emitter has been
observed between an organic host and a
phosphorescent organoiridium complex [C Adachi, R
C Kwong, P Djurovich, V Adamovich, M A Baldo, M E
Thompson, S R Forrest, Appl. Phys. Lett. 79, 2082
(2001)]. A signature of this phenomenon is that
it can be frozen out at low temperature, as it
requires thermal energy. Recently, despite the
problems, rather efficient exothermal blue
electrophosphorescence (≈ 9 lm/W) was
demonstrated [R J Holmes, S R Forrest, Y J Tung,
R C Kwong, J J Brown, S Garon, M E Thompson,
Appl. Phys. Lett. 82, 2422 (2003)].

Electrophosphorescence has also been observed


from a ladder- type poly(p-phenylene), that is a
typical fluorescent polymer [J M Lupton, A
Pogantsch, T Piok, E J W List, S Patil, U Scherf,
Phys. Rev. Lett. 89 (16), art. no. 167401
(2002)]. This observation has been rationalised
as the result of a very low concentration of

101
covalently bound heavy metal catalyst residue;
namely Palladium at about 80 ppm. This highlights
firstly, that extremely low impurity
concentrations can have profound effects on the
properties of organic semiconductors, and
secondly, that such effects in some cases may
lead to favourable behaviour, as well – in this
case, triplet harvesting.

102
References – Textbooks [T] and review papers [R]

[T1] G Hadziioannou, P F van Hutten, (eds.)


‘Semiconducting Polymers – Chemistry, Physics, and
Engineering’, Wiley-VCH, Weinheim, 19999 (ISBN 3-527-
29507-0) A collection of essays from many outstanding
scientists in the field, each addressing a specific
aspect in depth (and sometimes, beyond). Several
individual essays are referred to throughout the course.
Heavy on electroluminescence, transistors, photovoltaics,
and synt hesis of respective materials

[T2] M C Petty, M R Bryce, D Bloor (eds.), ‘Introduction


to Molecular Electronics’, Edward Arnold, London, 1995
(ISBN 0-340-58009-7. Again, a collection of essays, this
time covering aspects of molecular electronics which by
and large are NOT electroluminescence, transistors, or
photovoltaics, and not synthesis.

[T3] J D Wright, Molecular Crystals (2nd ed.) (Cambridge


University Press, Cambridge, 1995, ISBN 0-521-47730-1).
An excellent introduction into (organic) molecular
crystals. Many of the phenomena encountered in
semiconducting polymers have parallels to those in
molecular crystals; cynics might say organic
semiconductor research is a recent extension of molecular
crystals research.

[T4] Same as [T1] – sorry.

[T6] P W Atkins, ‘Physical Chemistry’, Oxford University


Press, 1998 (ISBN 0198501021). The most widely used
introduction into physical chemistry. Very broad
coverage.

[R1] A Kraft, A C Grimsdale, A B Holmes, Angew. Chem.


Int. Ed., 37, 403 (1998) [Kraft]. A broad discussion of
conjugated polymers for display purposes. Emphasis is on
Chemistry.

[R6] L Groenendaal et al., Adv Mater. 12, 481 (2000). An


excellent review on the chemistry, physics, and
applications of probably the best synthetic metal money
can buy, PEDOT/PSS. From the manufacturers’ research
team. (Elib)

[R7] G Horowitz, Adv. Mater. 10, 365 (1998). Elib

[R8] G Horowitz, J Mater. Chem. 9, 2021 (1999). Elib

[R9] H E Katz, J. Mater. Chem. 7, 369 (1997) Elib

103
Either of R7-R9 will serve you nicely as an introduction
into organic field effect transistors.

[R13] M Bernius, M Inbasekaran, E Woo, W Wu, L Wujkowski,


J. Mater. Sci.: Mater. El., 11, 111 (2000). The team from
DOW that has established fluorene homo- and copolymers as
viable alternative to the PPV family describes their
synthesis and use as hole transporting and emissive
polymers spanning the whole colour spectrum.
[R18] A B Kaiser, Adv. Mater. 13, 927 (2001) A review of
the physics of synthetic metals – more theory, less
applications, than [R6].

104

Вам также может понравиться