You are on page 1of 32

Review

Hall of Fame Article


www.advmatinterfaces.de

The Impact of Dipolar Layers on the Electronic Properties


of Organic/Inorganic Hybrid Interfaces
Egbert Zojer,* Thomas C. Taucher, and Oliver T. Hofmann

all devices that need to interface with out-


The presence of dipolar layers determines the functionality of most technologi- side circuits. The following discussion will
cally relevant interfaces. The present contribution reviews how periodic dipole evolve around the electronic properties of
assemblies modify the properties of such interfaces through so-called collec- such interfaces, where crucial parameters
are the energetic alignment between the
tive electrostatic effects. They impact the ionization energies and electron affin- electronic states in the materials consti-
ities of thin films, change the work function of metallic and semiconducting tuting the interface and adsorbate-induced
substrates, and determine the alignment of electronic states at interfaces. variations of the work function. In par-
Dipolar layers originate either from the assembly of polar molecules or they ticular, we will discuss how these proper-
arise from interfacial charge rearrangements triggered by the deposition of ties are fundamentally modified by dipolar
layers present at the interface.
an adsorbate layer. Such charge rearrangements result from the omnipresent
Therefore, after briefly addressing the
Pauli pushback caused by exchange interaction, from covalent bonds, or from role of interfaces in organic (opto)electronic
charge transfer following the deposition of particularly electron rich (donors) or devices in Section 2, Section 3 will focus on
electron poor molecules (acceptors). A peculiarity of charge-transfer interfaces the fundamental impact of regular arrange-
is that they enter the realm of Fermi-level pinning, where the sample work ments of dipoles on the electrostatic energy
function becomes independent of the substrate and is solely determined by of a charge carrier (at this stage mostly
without referring to actual materials). This
the electronic properties of the adsorbate. Beyond changing work functions discussion will form the basis for all aspects
and injection barriers, the presence of polar layers also modifies various other described later in this paper. In Section 4, we
physical observables, like core-level binding energies or tunneling currents in will apply the developed concepts to under-
monolayer junctions. All these aspects suggest that polar layers can also be stand the electronic properties of inter-
exploited for electrostatically designing the electronic properties of materials. faces and how they are impacted by polar
layers. Section 5 will then be dedicated to
discussing the origin of polar layers at inter-
faces, which either arise from the adsorp-
1. Introduction tion of polar molecules or from interfacial charge rearrangements
due to the deposition of an organic adsorbate layer. When such
The characteristics of essentially all organic electronic devices are interfacial charge transfer occurs, one also often enters the realm
dominated by the interfaces between their constituents. In organic of Fermi-level pinning, which results in several unusual inter-
electronics, such interfaces occur in different “flavors”: between face properties. For example, the work function of the combined
organic semiconductors (OSCs) and dielectrics (e.g., in thin-film system becomes independent of the work function of the sub-
transistors), between different OSC layers in multilayer stacks strate. Then, the presence of dipolar layers at the interface may no
of organic light-emitting devices and solar cells, between OSCs longer influences its electronic properties. Conversely, changes
and their inorganic counterparts in hybrid devices (e.g., hybrid in the adsorbate geometry at the interface may massively impact
solar cells), and between OSCs and (often metallic) electrodes in the system work function, even when the affected molecules a
priori do not possess a dipole moment of their own. Section 6
Prof. E. Zojer, T. C. Taucher, Dr. O. T. Hofmann will describe how dipolar layers at interfaces impact physical
Institute of Solid State Physics observables like charge-carrier injection barriers, core-level
NAWI Graz
binding energies, and the electrical conductivity of monolayers.
Graz University of Technology
Petersgasse 16, 8010 Graz, Austria As a last aspect related to the presence of dipolar layers, we will
E-mail: egbert.zojer@tugraz.at briefly summarize how they can be used to generate new mate-
The ORCID identification number(s) for the author(s) of this article rials with unprecedented properties exploiting the concept of
can be found under https://doi.org/10.1002/admi.201900581. “electrostatic design.”
© 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co.
KGaA, Weinheim. This is an open access article under the terms of the
Creative Commons Attribution License, which permits use, distribution
and reproduction in any medium, provided the original work is properly 2. Interfaces in Organic Electronic Devices
cited.
Prior to an in-depth discussion of the role of dipolar layers for
DOI: 10.1002/admi.201900581 the electronic properties of interfaces, it is useful to briefly

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (1 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

remember the diverse role interfaces play in organic elec-


tronics:[1,2] As far as the growth of organic semiconductor Egbert Zojer received his
layers is concerned, the interaction with the substrate controls Ph.D. in Physics (1999)
the growth mode of the organic layer, its texturing, and the from the Graz University
size of the formed crystallites.[3–6] In many instances, inter- of Technology, where he
faces even trigger the formation of surface-induced phases, was appointed associate
which possess distinctly different structures than the corre- professor in 2002. On leave
sponding bulk crystals.[7,8] In electronic devices, charge car- from Graz, he has worked for
riers can be trapped at interfaces. Historically, this significantly three years at the University
delayed the development of n-type organic transistors built on of Arizona and at the Georgia
SiO2 dielectrics.[9] More generally, charge-carrier trapping and Institute of Technology with
the resulting potential drop across the dielectric layer change Jean-Luc Brédas (2002–2005).
parameters like the threshold voltage and can cause hysteresis His research has always
effects.[10–14] In organic heterojunction solar cells, the energy- involved organic semiconductors and he has been dealing
level offset at interfaces between regions consisting of electron with topics like optical and electron spectroscopy, organic
rich (donating) and electron poor (accepting) molecules pro- devices, and materials modelling. In recent years, the focus
vides the driving force for overcoming the comparably large of his activities has been on understanding the properties
exciton binding energies.[15–17] Particularly complex organic of organic/inorganic hybrid interfaces, mostly employing
heterostructures are found in organic light-emitting devices, quantum-mechanical simulations.
where multiple interfaces between different organic semi-
conducting materials are exploited (often in conjunction with
molecular doping),[18] to facilitate electron and hole injection, Thomas C. Taucher received
carrier transport, and minority carrier blocking.[19,20] Finally, at his M.Sc. in Physics (2016)
the interface between OSC materials and electrodes, the ener- from the Graz University of
getic mismatch between the respective frontier electronic states Technology in Austria. He
results in the formation of energy barriers[1,21–27] that hinder is currently a Ph.D. student
charge-carrier injection and give rise to (injection-limited) con- in the group of Egbert Zojer
tact resistances.[12,28–30] working on the simulation of
Most of the abovementioned interface properties can be con- metal/organic interfaces.
trolled by dipolar layers. For example, the level offset between
donor and acceptor regions in organic (as well as perovskite)
solar cells can be tuned by incorporating polar interfacial
layers.[31–33] A particularly intensively investigated topic is the
modification of electrode work functions and, correspondingly,
charge-carrier injection barriers through polar adsorbate layers.
Since the seminal works of Campbell et al.,[34,35] self-­assembled Oliver T. Hofmann obtained
monolayers (SAMs) with polar groups at their periphery his M.Sc. in Chemistry
(so-called tail-group substituted SAMs) have been grown on
­ (2007) and Ph.D. in Physics
various electrode materials with the intention of tuning elec- (2010) at the Graz University
trode work functions and of reducing contact resistances in of Technology. Afterward,
(opto)electronic devices.[2,36–47] Notably, when applying such he acquired a Max-Planck
SAMs, it has always been difficult to determine whether their fellowship and an Erwin-
beneficial effect resulted primarily from their impact on the Schrödinger grant to join
growth of the OSC film,[48] or whether induced changes of the the groups of Nobert Koch
electronic properties of the interface played the decisive role. (HU Berlin) and Matthias
Therefore, recently also SAMs with dipoles embedded into the Scheffler (FHI Berlin), where
molecular backbones[49,50] have been successfully employed he worked on improved
in organic devices.[49,51] These SAMs allow changing contact methods to computationally describe metal/organic and
resistances by several orders of magnitude without impacting semiconductor/organic interfaces. In 2014, he returned to
the growth of the OSC, thus demonstrating the crucial role of his alma mater as group leader, where he now focusses on
the layer’s polarity.[51] the first-principles prediction of the atomistic structure of
SAMs with polar tail groups have also been grown on the hybrid interfaces.
dielectric layers of organic thin-film transistors in order to tune
the turn-on voltages of the devices.[52,53] Originally, the SAMs’
impact on the device performance has been associated with the observed changes of turn-on voltage, which amount to sev-
dipole moments of the adsorbed molecules and with the elec- eral ten volts.[11,55,56] Therefore, in the meantime it has been
tric field inside the adsorbate layer.[52–54] However, as we will largely accepted that the crucial dipole at these interfaces is not
argue below, the shifts in the electrostatic energy landscape the one associated with the polar substituents of the adsorbed
induced by realistic dipole moments are inconsistent with the ­molecules. Rather, the shift in turn-on voltage is caused by the

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (2 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

much larger dipole that forms across the entire dielectric due properties of interfaces. For that, it is useful to compare how
to trapping of charge carriers at the SAM and their counter- isolated dipoles, ordered arrays of point dipoles, and sheets of
charges in the gate electrodes.[11,13,14,57] point charges impact the electrostatic energy of electrons in
These observations show that polar layers play a crucial role their surrounding (see also refs. [64,65]).
for devices. Their exact origin and their detailed properties can, The electrostatic energy of an electron at rest (i.e., without
however, sometimes be comparably complex. This has trig- kinetic energy) is typically referred to as the vacuum level.
gered a multitude of dedicated investigations on polar layers When an electron comes close to a single isolated dipole, the
beyond employing them in devices and measuring the resulting electrostatic energy becomes strongly perturbed, as shown
changes of the device characteristics. Indeed, when aiming at in Figure  1a. Naturally, the perturbation depends on whether
a fundamental understanding of intrinsic interface properties, the electron is close to the “positive” or “negative” end of the
one typically relies on the key techniques of surface science
like photoelectron spectroscopy, scanning probe techniques,
and scattering approaches. Such investigations are often per-
formed under ultrahigh-vacuum conditions, either required by
the employed techniques or chosen to avoid contaminations and
to realize well-defined structures. A drawback of investigating
interfaces in ultrahigh vacuum is that such conditions are usu-
ally not applied when growing and operating practically relevant
devices. In other words, there typically exists a “pressure gap”
between surface-science investigations of interfaces and their
impact on (opto)electronic devices of practical relevance. Never-
theless, the surface-science investigations are extremely useful,
as well-defined and “clean” conditions are crucial for achieving
a fundamental understanding of interfaces at an atomistic level.
Therefore, in the following we will mostly rely on the insights
generated by such experiments and related simulations. Notably,
independent of the “pressure gap,” all fundamental aspects dis-
cussed below will also hold under typical device fabrication and
operation conditions-only the magnitude of the encountered
dipole layers might differ between atomically clean and “tech-
nical” interfaces (for the reasons discussed in Section 5.2.1).
To conclude this section, we would like to make a final note
regarding the relation between the interfacial level alignment
determined in surface-science investigations and the efficiency
of carrier injection processes in devices: Surface-science studies
typically yield the so-called nominal charge-carrier injection
barriers determined by the offset of the frontier states in the
OSC and the Fermi-level characterizing the electronic proper-
ties of the electrode. The effective barrier experienced by charge
carriers in an operating device is, however, significantly lower
due to the Schottky effect arising from a combination of the
electric field at the interface caused by the externally applied
bias in combination with the field due to the interaction of the
injected carrier with the polarization charge it causes in the
conductive substrate.[58–61] This not only yields a massive reduc-
tion of the contact resistance, but also results in a pronounced
dependence of the contact resistance on device geometry and Figure 1.  Change in the electrostatic energy of an electron induced by
operation conditions.[62,63] It also renders an explicit considera- assemblies of dipoles arranged in the following way: a) point dipole;
tion of barrier shaping fields crucial when trying to understand b) infinitely extended 2D sheet of point dipoles on a quadratic grid (with
lattice constant a) in the x–y plane. The direction of the dipoles is perpen-
carrier injection for actual device structures. dicular to the sheet; c) series of two infinitely extended sheets of point
dipoles at a distance of three times a; d) sheets of oppositely charged
point charges (corresponding to extended dipoles in contrast to point
dipoles) arranged on a quadratic grid with lattice constant a. The distance
3. Impact of Dipole Layers on the Electrostatic between the sheets amounts to two times a. The plotting range in all
Energy of Electrons panels corresponds to 6∙a  × 6∙a. Such a relative length scale is useful,
as the shape of the energy landscape is determined by the ratio between
3.1. Densely Packed Dipole Layers
the distance from the dipole sheet and the interdipole distance. In the
plot we also refrain from quantifying the changes in electrostatic energy,
The purpose of this section is to develop a fundamental under- as these scale linearly with the magnitude of the point dipoles and point
standing of the effect of ordered polar layers on the electronic charges.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (3 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

point dipole. Hence, when the charge is close to the dipole, its other dipoles. This field, which can be calculated employing
energy is significantly impacted by its position. However, at the Topping model,[73] induces a polarization that reduces the
increasing distances the perturbation decays rapidly, such that net dipole moment of each molecule.[74,75] Such a behavior
the exact position of the electron no longer plays a role. Conse- has been observed both in experiments[69,76–80] and in simula-
quently, there is only a single, unique energy associated with tions,[64,67,76,81–86] considering a variety of different substrates
the vacuum level. and adsorbate layers. The magnitude of these depolarization
The situation is fundamentally different when considering effects depends on the polarizability of the electrons close to
extended assemblies of dipoles. This is shown in Figure 1b the polar element, as well as on the packing density of the
for the example of an infinitely extended 2D sheet of point dipoles. In this context, it should be emphasized that the rel-
dipoles (there, for the sake of simplicity, the dipole moments evant polarizability is not the one of the full, isolated mole­
are arranged in z-direction). As a consequence of the superposi- cule. First, even when applying a homogeneous external
tion of all electrostatic potentials, i.e., due to so-called collective field, the polarizability of a molecular monolayer significantly
(or cooperative) electrostatic effects, a step in the electrostatic deviates from that of a molecule.[87] Second, the field gener-
energy appears at the position of the dipole sheet, zd.[64,66–70] ated by an extended assembly of dipoles falls off very quickly
This means that the energy of an electron changes when it (for tightly packed molecules, this can be within a few Å).[64]
is moved from left of the sheet (from z  <  zd) to right of the Thus, the field is highly inhomogeneous and typically not the
sheet (to z > zd). In other words, the step in electrostatic energy whole molecule experiences the field from the other dipoles.
results in different vacuum levels left and right of the dipoles. As a consequence, depolarization is better described by an
Therefore, the infinitely extended 2D sheet of dipoles splits effective dielectric constant, εeff, such that Equation (1) can
space into a region left and a region right of the dipole layer be expressed in terms of the dipole moment of the isolated
with a rigid shift in the relative energetics between the two molecule µ0[83]
regions. This has several implications: qe µ 0
∆VL = − (2)
i. From the Poisson equation it follows that the magnitude of
ε 0 ε eff A
that step, ΔVL, is proportional to the dipole moment per area, The magnitude of depolarization effects can be sizable: For
µ/A example, calculations for densely packed biphenylthiolate
films bearing CN and NH2 substituents yield values of
qe µ
∆VL = − (1) εeff around 3.[83]
ε0 A iii. As indicated already above, there is virtually no electric field
or potential gradient outside a densely packed assembly of
Here qe is the charge of an electron and ε0 is the vacuum per- dipoles, even at distances shorter than the interdipole dis-
mittivity. Only the component of the dipole moment perpen- tance.[64] Consistently, the drop in electrostatic energy in
dicular to the plane of dipoles determines ΔVL, rather than Figure 1b is quite abrupt. Figure 2a shows this situation for
its absolute value. When there are also parallel components, an actual dipolar SAM consisting of cyano-biphenylthiolates
the situation becomes more involved, as then the electro- on a Au(111) surface: For tightly packed molecules at full
static energy also varies as a function of x and y. In practice, coverage (left panel), the CN groups significantly change
such a situation cannot occur for a dipole layer on top of a the vacuum level, but the step in energy is confined to a very
metal surface, as there the components of the dipoles par- narrow region close to the substituents. Consequently, such
allel to the surface (and the resulting electric field) are fully a densely packed dipole layer will not polarize the substrate
screened by the conduction electrons of the metal. On sem- as long as it is located sufficiently far away from it. In such a
iconductors and dielectrics, the screening of the µx and µy case, there is also no need to invoke mirror charges or mirror
components will generally be incomplete. This results in lat- dipoles in order to describe ΔVL. Conversely, at 1/16th cover-
eral gradients of the electrostatic energy, with the overall shift age, there is virtually no long-range shift in the electrostatic
possibly as large as the bandgap. In thermodynamic equilib- energy induced by the cyano groups. Instead, significant
rium, larger shifts are prevented by additional charge rear- variations in the electrostatic energy are found in vacuum
rangements in analogy to what has been predicted for polar several Å right of the SAM, as well as in the empty region
multilayer films.[71,72] Alternatively, lateral potential gradients between the molecules, where they reach down to the metal
can be prevented or at least mitigated by large-scale surface substrate.
reconstructions. A further conceivable scenario limiting the iv. A consequence of the abrupt potential drop for densely
lateral variation in electrostatic energy is the formation of packed layers is that dipoles in consecutive layers do not
domains with differently oriented polarization. To avoid such influence each other, provided that the distance between
complications, for the future discussion we will assume that them is sufficiently large. This is shown in Figure 1c for
all dipole moments are oriented exclusively in z direction. two consecutive dipole layers positioned at three times
ii. Importantly, screening does not only impact the x and y com- the distance between the dipoles in each layer. In such a
ponents of the dipoles. Quite generally, when realizing dipo- situation, the net shift of the electrostatic energy is sim-
lar layers via the assembly of polar molecules, the dipole per ply the sum of the shifts caused by each of the two dipole
molecule in the assembly is different from the dipole of the layers, an effect that has already been observed by Adam
isolated molecule. The reason for that is that each molecule et al. in 1934.[64,89] The main consequence of this aspect
within the layer feels an electric field originating from all is that w
­ ell-separated layers of dipoles neither polarize nor

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (4 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

depolarize each other. In practice, this allows to functional- In the above discussion, we have considered point dipoles or
ize molecules at different positions (e.g., use different head dipoles associated with polar substituents of comparably small
and tail groups), with the net effect of the resulting dipoles spatial extent (e.g., the CN groups in Figure 2a). The situation
being simply additive.[90] changes, when the dipoles are spatially extended. This occurs,
for example, when they originate from the charge transfer
between a substrate and an adsorbate layer several Å above the
topmost substrate atoms (see Section 5.2). The simplest model
representing that situation consists of two parallel sheets of
point charges of opposite polarity. The field between the charged
sheets is essentially homogenous, provided that the distance
between the charges within the sheets is in the same range as
or smaller than the distance between the sheets, as is shown
in Figure 1c. Outside the charged double layer, there is virtu-
ally no field and the electrostatic energy becomes constant. This
is analogous to the situation of a charged plate capacitor. Also
when the adsorbed molecules undergoing interfacial charge
transfer are significantly larger than the adsorption height (i.e.,
the distance to the substrate), an analogous situation prevails
provided that the adsorbate layer is densely packed, that all
molecules are equally charged, and that the charge is rather
homogeneously distributed within the adsorbed molecule (e.g.,
when it is transferred to a delocalized π-orbital). For the afore-
mentioned cases, the linear drop of the electrostatic energy
between the charged sheets amounts to

qe σ d qe σ 0 d
∆VL = − =− (3)
ε0 εrε0

Here, d is the distance between the sheets and σ is the surface


charge density. In cases in which there is a medium between
the charged sheets, this can be accounted for in two ways: One
can include all polarization charges already into σ (e.g., when
implicitly accounting for polarization effects in a simulation).
Alternatively, one can determine σ0, the surface charge density
in the adsorbate disregarding the polarization charges in the
dielectric. In a second step, one then accounts for the polari-
zation of the dielectric medium via its dielectric constant εr.
Such a description is particularly useful in more “macroscopic”

Figure 2. a) DFT calculated electrostatic energy of a CN-substituted


biphenylthiolate SAM relative to the substrate Fermi level, EF, in the plane
of the molecules for dense packing (left) and at reduced coverage (right).
In the latter case, for the sake of the argument, the geometry has been
obtained by generating a 4 × 4 supercell and removing all but one mole­
cule (i.e., the geometry at reduced coverage has not been optimized).
Isolines are plotted every 0.1 eV. b) Transfer characteristics (drain current,
ID, as a function of gate-source voltage, VGS) of pentacene-based thin-film
transistors containing acidic SAMs at the interface between the dielectric
and the organic semiconductor. The drain currents are plotted on a linear
(left axis) and logarithmic scale (right axis). The three panels correspond
to three different thicknesses of the dielectric (100 nm for the top panel,
150 nm for the middle panel, and 250 nm for the bottom panel). The red
data points have been measured for the as-fabricated devices (where the
acidic SAMs generate an interfacial layer of trapped charges due to the
acid residues), while blue data points have been obtained after exposing
the devices to NH3 gas (resulting in a compensation of the interfacial
space-charge layer). The shift of the turn-on voltage between the two situ-
ations (depicted by the green arrow) is due to the dipole layer formed
across the interface in the as-fabricated devices. It scales linearly with the
thickness of the dielectric. For further details see ref. [88] (b) Adapted with
permission.[88] Copyright 2011, Wiley-VCH Verlag GmbH & Co. KGaA.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (5 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

cases, for example, when the dipole layer originates from the 3.2. Finite Extent, Defective, and Inhomogeneous Dipolar Layers
accumulation of charges on two sides of the gate dielectric in
an organic thin-film transistor. Notably, the above considerations apply only to infinitely
There are two fundamental differences between the two extended, defect-free, densely packed dipolar layers. Deviations
limiting cases of dipole layers consisting of point dipoles and from such perfect structures have several important implications.
arising from charged double layers: For point dipole layers, the
step in electrostatic energy is very abrupt (see above discussion),
while for a charged double layer the energy changes gradually 3.2.1. Dipolar Layers of Finite Extent
(for a homogeneous medium linearly) between the charged
sheets. Moreover, for point dipole layers, the change in electro- While the change in electrostatic energy due to a dipole layer of
static energy is determined by a single parameter, namely the infinite extent can be described by Equations (1)–(3), for dipole
dipole density. Conversely, for charged double layers, besides sheets of finite extent these equations apply only provided that
the charge density also the distance between the charged sheets the distances to the layer are comparably short. For large dis-
plays an important role. This has several ­consequences: For tances the drop in the electrostatic energy is proportional to
electron donating or electron accepting adsorbate layers under- 1/z2, as the situation approaches that of a point dipole.[21] This
going interfacial charge transfer, establishing thermodynamic is shown in Figure 3a, where we plot the distance dependence of
equilibrium usually requires a certain system-dependent shift the electrostatic energy for square dipole clusters of varying size.
in electrostatic energy (see Section 5.2). In such a situation, At very short distances, the energy above the center of the cluster
increasing the charge transfer distance results in a reduction is close to the value of the infinitely extended dipole sheet. The
of the amount of transferred charge and vice versa. Conversely, shift, however, drops significantly at distances on the order of the
when the amount of accumulated charge is fixed, changing lateral extent of the sheet, as can be seen in the linear–log plot
the charge transfer distance results in an increase of the shift in the top panel of Figure 3a. Then, the electrostatic energy dis-
in electrostatic energy. This effect has, for example, been plays the expected quadratic drop with z, which results in a linear
observed in organic thin-film transistors containing reactive evolution with a slope of −2 in the log–log plot (bottom panel of
(acidic) monolayers at the interface between the dielectric and Figure 3a). Another important aspect to keep in mind for finite-
the organic semiconductor in the transistor channel.[57] There, extent dipole sheets is that at their boundaries inhomogeneous
the proton transfer to the semiconductor generates a fixed fields appear. This effect is comprehensively discussed in ref. [64].
density of trapped interfacial charges. They can be compen-
sated by mobile charges in the organic semiconductor (due to
the proton-transfer doping).[88] Alternatively, when the mobile 3.2.2. Defective Dipole Layers
charges are removed by depleting the channel, the trapped
interface charges induce an accumulation of an equivalent Conceptually similar considerations as for dipole clusters
amount of counter-charges in the gate-electrode on the other of finite extent apply to infinite sheets containing finite-size
side of the dielectric. The resulting shift in the electrostatic defects. Electric fields and their corresponding potential varia-
energy across the dielectric linearly translates into a shift of the tions “leak” through the hole in the dipole layer on length-scales
device characteristics (see Figure 2b).[11,88] The magnitude of that are larger than the extent of the defect. This is shown for a
the change in the turn-on voltage of the transistor is then pro- 2D dipole sheet with a hole of 6 × 6 dipoles in Figure 3b. Here,
portional to the thickness of the dielectric, as is indicated by the it should be mentioned that the plot in Figure 3b assumes all
green arrows in Figure 2b. The net effect can be enormous, as point dipoles to be of equivalent magnitude. In actual, defec-
exemplified for the transistor with the thickest dielectric consid- tive polar layers this is not necessarily the case, as the above-
ered in ref. [88] (d = 250 nm), where the shift amounts to more described depolarization effects are weaker for dipoles close
than 100 V. to the holes, as well as for dipoles close to the edges of island
Finally, it should be noted that yet another type of interfacial structures. A more detailed discussion of the resulting “electro-
charge distribution will become relevant for the following dis- statics of nonideal polar monolayers” can be found in ref. [64].
cussion: For extended semiconductor substrates the charge in
the semiconductor does not have to be concentrated into a thin
(mono)layer, but can spread over a wider region. Within the abrupt 3.2.3. Inhomogeneous Dipole Layers
junction approximation, that is typically invoked for inorganic
1 Significant deviations from the idealized picture also occur
semiconductors,[91] this region is proportional to (where N
N when the dipoles at the interface are far apart from each other.
is the amount of immobile charge carriers, e.g., due to ionized As a hypothetical model scenario, we already described the elec-
dopants). This results in a linear increase of the magnitude of trostatic energy distribution for a polar low-coverage SAM in
the electric field from the edge of the depletion (accumulation) the right panel of Figure 2a. In practice, the molecules in such
region toward the interface and, correspondingly, in a quad- low-coverage SAMs do, however, not remain upright. They
ratic change of the electrostatic energy. This situation is typi- rather “fall over,” which qualitatively changes their electronic
cally encountered in semiconductor pn- or Schottky-junctions. properties.[93] Thus, in the following we will use flat-lying mole­
A detailed mathematical description of this band bending can cules, for which the dipole moment originates from ­interfacial
be found, for example, in solid-state physics and semiconductor charge transfer, as model systems for describing inhomoge-
(device) textbooks (e.g., ref. [91]). neous dipole layers. In such cases, low dipole densities are

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (6 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

encountered, for example, when a surface is incompletely


covered and the molecules repel each other.[94,95] Another pos-
sibility is the adsorption of extended molecules in which the
charge transfer occurs to only a small region of the adsorbate
(e.g., in metal-phthalocyanines, if the interfacial charge transfer
occurs to only the central metal atom). In such cases, the low
dipole density causes a comparably small shift of the vacuum
level. Conversely, at the position of the adsorbed molecule,
the local dipole has a much larger impact on the electrostatic
energy, reminiscent of the situation of an isolated point dipole.
This is illustrated in Figure 3c by calculations for tetrafluoro­
benzoquinone on Ag(111):[92] For full monolayer coverage (red
line), the electrostatic energy experiences the abrupt step dis-
cussed above for densely packed dipoles. It is essentially the
same at the position of the molecule, z TFBQ,C and far from the
interface. Conversely, at 25% coverage (dashed blue) the shift
in electrostatic energy at the position of the molecule, z TFBQ,C, is
essentially twice as large as the shift of the vacuum level above
the surface. A similar situation is obtained, when the surface
coverage is complete, but only a fraction of the molecules expe-
rience interfacial charge transfer. This occurs, for example, for
very weakly coupled interfaces (see the integer charge transfer
scenario discussed in Section 5.2.4), or at interfaces at which a
fraction of the molecules is chemically modified, e.g., through
reactions with a residual gas.[92]

3.2.4. Practical Implications of Defective and Inhomogeneous


Layers

To assess the practical impact of finite-extent, defective, or


inhomogeneous dipole layers, one has to identify which pro-
cesses and lengthscales are relevant for a given experiment.
For example, the nominal barrier for thermionic or tunneling
injection of carriers into (opto)electronic devices is determined
by the electrostatic energy right above the electrode at the posi-
tion where the injection occurs. In that case, one is in a “near-
surface” situation that is very sensitive to the local electrostatic
energy. Similar considerations apply to Kelvin-probe (KP) force

6 × 6 dipoles are missing. The plotting range corresponds to 40∙a in the


direction parallel and 6∙a in the direction perpendicular to the dipole sheet.
c) Change of the electrostatic potential energy, ΔU, upon deposition of the
strong electron acceptor tetrafluorobenzoquinone (TFBQ) on Ag(111), as
a function of the vertical distance from the molecular adsorbate. The red
line shows ΔU for the adsorption of a full monolayer of TFBQ. In that
case, the vacuum level above the sample (determined by ΔU at high z) is
almost the same as ΔU at the position of the molecule (at z − z TFBQ,C = 0).
Figure 3.  a) Linear–log and log–log plot of the electrostatic energy of a The blue, dashed line shows ΔU for a reduced coverage of 25%. There,
2D square cluster of finite extent consisting of point dipoles (calculated the potential at z − z TFBQ,C = 0 has essentially the same value as in the
above the center of the cluster) as a function of the distance from the full monolayer situation (indicating that the molecular levels are shifted
cluster. All dipole moments in the sheet are assumed to be oriented per- just as much relative to the metal). However, due to the loose packing,
pendicular to the 2D plane and of the same magnitude. The length scale for larger values of z, ΔU decays to ≈50% of its maximum value. The full
is given in multiples of the interdipole distance, a, and the shift in electro- blue line corresponds to the adsorption of a mixed monolayer of TFBQ
static energy is plotted relative to the situation obtained for an infinitely and para-tetrafluorobenzodiole (TFBD), a weakly physisorbed molecule.
extended dipole layer. The considered clusters consist of 10 × 10 (red), Qualitatively, the situation is equivalent to that for the neat TFBQ film
40 × 40 (blue), 160 × 160 (yellow), and 640 × 640 (green) point dipoles. The at low coverage (maximum of ΔU at z − z TFBQ,C = 0 and decay to smaller
vertical arrows denote the z-distances from the clusters that equal their values for larger distances from the surface). A quantitative discussion is
x,y-extent. b) Spatially resolved change in the electrostatic energy induced given in ref. [92]. Reproduced under the terms of the Creative Commons
by an infinitely extended 2D sheet of point dipoles in which the central CC-BY license.[92] Copyright 2015, American Chemical Society.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (7 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

microscopy measurements.[96–99] Also in X-ray photoelectron the energy of an electron at rest infinitely far away from the
spectroscopy (XPS) experiments, the relevant quantity is the local metal. Its position is independent of the path along which the
electrostatic energy around the atom at which the core electron is electron is removed. In contrast to the situation at infinity,
excited.[100,101] the vacuum level near the metal surface does depend on the
The “near surface” quantities are likewise relevant, when per- actual orientation of the surface through which the electron is
forming conventional KP measurements, or when determining ejected.[21] For example, for tungsten the vacuum level meas-
the secondary cutoff (SECO) in photoelectron spectroscopy exper- ured for the (110) surface is 0.62 eV higher than that of the
iments. For the SECO this is because the dipoles create a poten- (100) surface.[108] This is a consequence of the spill-out of elec-
tial barrier near the surface, which needs to be overcome. This trons from the metal surface, which extends over several Å[109]
barrier (rather than the electrostatic energy at the detector) deter- and creates a dipole layer that shifts the energy landscape (and
mines the lowest kinetic energy electrons that are able to leave the with it the vacuum level) above the surface. The magnitude
sample. These define the energetic position of the SECO.[102] of that spill-out depends on the packing density of the metal
As KP and photoemission average over comparably large atoms in the top layer and, thus, on the exposed surface. Due to
areas, they are, however, rather insensitive to local lateral varia- the finite lateral extent of the metal substrate, upon increasing
tions in the electrostatic energy. Therefore, only a single cutoff the distance from its surface the surface vacuum level gradu-
(corresponding to the average electrostatic energy at the surface) ally merges into the vacuum level at infinity, as discussed in
is detected, even when lateral inhomogeneities extend over sev- Section 3.2.1 (see Figure 3a). Notably, the quantity relevant
eral nm.[103] Only for rather extreme lateral energy variations, a for the following discussion of the interface energetics is the
second cutoff can be observed.[104] For such a situation (in this vacuum level directly above the metal. In the following, the
case encountered for a patterned surface with a periodicity of sev- energy difference between this energy level and the Fermi level
eral µm), Schultz et al. have shown recently that one of the two will be referred to as the work function of the substrate, Φsub.
cutoffs corresponds to the area-averaged value of the electrostatic
energy, while the second one is determined by its maximum.[105]
Interestingly, in some cases defects and deviations from 4.1.2. Inorganic Semiconductor Substrates
the perfect periodicity in polar adsorbates are even crucial for
adsorbed polar layers to have an impact: For example, it has The key quantities for semiconductors are the electron affinities
been argued that in Schottky junctions, in which the metal and (EA) and ionization energies (IE) of the material, i.e., the ener-
the semiconductor are separated by a polar organic layer con- gies gained when adding an electron and paid when removing
taining pinholes, the dependence of the current through the an electron from the semiconductor. Within a single-particle
pinholes on the dipole orientation arises from stray fields in picture (and for the discussion throughout this paper), they
the defective SAM.[64] Also the impact of adsorbed polar mono­ are approximated by the positions of the valence and conduc-
layers on the electric current in silicon devices[66,106] has been tion band edges. In practice, (almost) all semiconducting sub-
associated with charge transfer triggered by residual fields.[64] strates are doped. Therefore, they exhibit a Fermi level that will
For further details on the role of molecular monolayers within ­typically deviate from the mid-gap position. As for metallic sub-
metal/semiconductor junctions, we refer the interested reader strates, the energetic distance between the Fermi level and the
to the recent dedicated review by Vilan and Cahen.[107] vacuum level defines the substrate’s work function, Φ. Notably,
the positions of the Fermi level and the band edges relative
to the vacuum level above the sample can again be modified
4. Electronic Structure of Inorganic/Organic by the presence of surface dipoles. Such dipoles occur, e.g.,
due to ­different surface reconstructions and terminations. A
Hybrid Interfaces
common example are hydrogen overlayers, which saturate
Prior to discussing the possible origins of dipolar layers at inter- the dangling bonds at a cleaved surface.[110] Moreover, signifi-
faces, it is useful to briefly review the fundamental aspects of the cant electric fields can be present within thin films of polar
electronic structure of organic–inorganic heterostructures, and semiconducting substrates,[111–115] where the field originates
how they are impacted by dipolar layers. Here, as a first step, from the asymmetry of the unit cell. Notably, the presence of
it is useful to discuss the electronic properties of the individual surface states due to dangling bonds at the semiconductor sur-
components that will eventually form the heterostructure. face or due to (geometric) surface reconstructions can change
the density of mobile carriers in the semiconductor close to the
interface. The resulting net charge density (e.g., due to uncom-
4.1. Properties of the Sub-Systems Forming the Interface pensated ionized dopants) locally causes an electric field, which
gives rise to surface band bending ΔΦBB (cf., discussion in
4.1.1. Metal Substrates Section 3).[91] Therefore, akin to the role of surface dipoles for
the work function of a metal, the band bending plays a crucial
The most relevant parameter describing the electronic proper- role for the effective work function of the substrate Φsub also
ties of a metal in the present context is its work function, Φ (see for semiconductors, as indicated in Figure 4b. Importantly, the
Figure 4a). It represents the energy to add/remove an electron presence of band bending leads to different energetic positions
to/from the metal and can be described as the energy differ- of the band edges in the bulk viz-a-viz the surface: Ionization
ence between the Fermi level of the metal and the vacuum energy and electron affinity in the bulk (defined as the ener-
level. As discussed in Section 3, the vacuum level is defined as gies of the band edges relative to the vacuum level outside the

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (8 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 4.  Schematics of the electronic structure of a) a metal substrate, b) an inorganic semiconductor substrate including surface band bending,
c) an isolated organic molecule, and organic monolayers consisting of molecules with d) zero or e) nonzero net dipole moments. The various indicated
quantities are explained in detail in the main text. Shaded orange areas correspond to states in the metal substrate and in the semiconductor, while
blue areas correspond to orbitals or bands in the organic molecules and monolayers. Dark shading refers to occupied and light shading to unoccupied
states. The horizontal arrow in (b) symbolizes the dipole layer associated with band bending.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (9 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

sample) are increased (in the case of upward band bending, as thermodynamic equilibrium charge-transfer processes occur is
shown in Figure 4b) or reduced (for downward band bending) determined by the adiabatic quantities.
by ΔΦBB. Conversely, at the very surface the positions of the When the molecules form a (hypothetical) free-standing
bands relative to the vacuum level remain unchanged by ΔΦBB monolayer (or a molecular crystal) several effects occur, which
(as indicated by EASC and IESC in Figure 4b). The local shift modify the quantities discussed above (see Figure 4d)
of the bands, however, modifies their distance from the Fermi
level, which is consistent with the change in the free-carrier i. In addition to the geometry relaxation effects, also dielectric
density close to the surface. screening within the monolayer stabilizes the charged spe-
In the present context it is important to emphasize that the cies. Thus, screening increases the electron affinity and re-
semiconductors used as transparent electrodes for organic duces the ionization energy.[21] Electrons and holes within the
optoelectronic applications are typically highly doped wide- monolayer get dressed and form quasi-particles, which drag a
bandgap semiconductors, like indium-doped tin oxide (ITO), polarization cloud along. The resulting changes in energy are
ZnO, or GaN. Due to their large ionization energies, they are typically on the order of 1–2 eV for OSC bulk materials, both
often (involuntary) n-type doped.[55,116–118] This leads to states experimentally and in simulations.[125–129]
in the bandgap close to the band edges, albeit with a very low ii. Due to the interaction between neighboring molecules,
density. Often, the actual situation in such wide-bandgap semi- bands of finite width form. As the IEs and EAs are defined
conductors is not well described with a single kind of dopant. based on band edges, this again reduces IEs and increases
Rather, a more realistic scenario is to invoke shallow donors EAs. The effects described in (i) and (ii) also decrease the
(e.g., due to hydrogen uptake[116,117]) and deep donor levels fundamental gap of the monolayer, FGml, compared to the
(e.g., from surface oxygen vacancies, which are prevalent in isolated molecule.
transparent conductive oxides).[55,119,120] As a consequence, the iii. An additional aspect that is often overlooked but absolutely
band bending cannot assume arbitrary values (limited only by crucial for interfaces is that also in molecular layers, surface
the magnitude of the bandgap), as one would expect for con- dipoles exist.[130] These arise mostly from polar bonds in the
ventional semiconductors. Instead, band bending due to ion- molecular periphery, which can be due to polar substituents,
ized shallow donors only occurs until it shifts the levels of peripheral heteroatoms, or even CH bonds (which are also
the deep donors into resonance with the Fermi level.[121,122] At polar). When ordered molecular assemblies are formed and
this point, the deep donors become electronically active, effec- when there is a nonvanishing dipole component perpendicu-
tively increasing the amount of available charge carriers. In lar to the surface of the layer, a step in electrostatic energy oc-
practice, the result of this effect is that the observed values for curs, which shifts the molecular states relative to the vacuum
band bending tend to be much smaller than one would expect level. In the following, it will be referred to as the surface
in the absence of the deep donor states.[121,123] Accordingly, dipole of the monolayer, SDml (cf., Figure 4d). Notably, this
also the band-bending region becomes much narrower than surface dipole shifts the occupied and unoccupied states in
what would be predicted when only considering the density the same direction and, thus, does not change the fundamen-
of shallow donors. Although this means that the abrupt junc- tal gap of the monolayer. As SDml depends on the orientation
tion approximation may often not be well justified for these of the molecules in the monolayer, changing that orientation
semiconductor materials, in lieu of a better model it is still fre- also changes the monolayer ionization energies and electron
quently used in simulations. affinities, IEml and EAml. In passing, we note that a finite
SDml also exists for disordered layers, since the surface di-
poles at the periphery of organic molecules sometimes point
4.1.3. Organic Semiconductors in the same direction largely independent of the molecular
orientation (e.g., for pure hydrocarbons, all CH dipoles
For discussing the electronic properties of molecular organic point outward). Thus, even after averaging over random ori-
semiconductors, it is useful to start with the underlying mole­ entations, a local net dipole at the surface of the organic layer
cular properties, like the molecular ionization energy, IEmol, and often persists.
the molecular electron affinity, EAmol (see Figure 4c). Their dif- The impact of surface dipoles in organic molecules
ference defines the fundamental gap, FGmol. This quantity is has been shown, for example, for the ionization energy
not to be confused with the excitation gap measured in optical of a α,ω-dihexyl-sexithiophene, which can be changed by
absorption experiments, which is smaller by the exciton binding ≈0.6 eV solely by v­arying the orientation of the molecules
energy. In molecules the exciton binding energy is sizable due to in the mono­layer.[130] Furthermore, calculations by Heimel
weak electrostatic screening and it remains large also in organic et al.[131] on highly ordered poly(3-hexylthiophene) mono­
solids (amounting to several tenths of an eV up to ≈1.6 eV).[124] layers suggest that varying the orientation of the polymer
Generally, one has to distinguish between vertical and adi- chains changes the monolayer IEs and EAs by 0.4 eV, while
abatic energies, where the latter includes the energy gained by for partially fluorinated side chains the magnitude of effect
the geometry relaxation of the charged systems. C ­ onsequently, not only more than quadruples, but the shift also changes its
adiabatic IEmol’s are smaller and adiabatic EAmol’s are larger sign. For layers consisting of inequivalent molecules with dif-
than vertical ones.[21] Whether one has to consider vertical ferent terminations, changing their ratio even allows tuning
or adiabatic values depends on the process of interest. For the ionization energy of the monolayers.[132] In all examples
example, for quasi-instantaneous photoionization processes the discussed in this paragraph, the dipoles at the left and right
relevant quantities are the vertical ones. Conversely, whether in “surface” of the monolayer are the same. Thus, the net dipole

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (10 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

moment of the monolayer vanishes and the vacuum level left The net work function of the combined system in the Schottky–
and right of the layer has the same value (see Figure 4d). Mott limit, ΦSM, becomes
The situation changes qualitatively when the SDml to ΦSM = Φsub + ∆VL ml (5)
the left and to the right of the monolayer are no longer
identical. This is, for example, the case for “classical” self- Consequently, for the adsorption of a layer with no net dipole
assembled monolayers, in which the docking group (on moment (cf., Figure 5a), ΦSM and Φsub are identical, i.e., the
the left side of the schematic monolayer in Figure 4e) is work-function change ΔΦ is zero.
chemically different from the tail group substituent (on the When switching on the interaction between substrate and
right side).[21,35,50,65,68,70,133–136] Then the adsorbate layer cor- adsorbate,
responds to two successive dipole layers, whose effects no
longer cancel and, thus, induce a shift in the vacuum level, i. Screening effects are massively increased. This results in
ΔVLml, between the left (VLleft) and right (VLright) side of the a sizable bandgap renormalization for adsorption on met-
monolayer (Figure 4e). Importantly, in such a situation the als[141–145] as well as on insulators.[146] Consequently, the elec-
monolayer ionization energies and electron affinities are no tronic parameters of the adsorbate layer are again modified,
longer unambiguously defined. One now needs to differen- leading to the interface-specific quantities EAint and IEint.
tiate between left- and right-sided quantities, as removing an Note that the bandgap renormalization decreases roughly
electron from the monolayer to the left and to the right costs with one over the distance of the adsorbate layer from the
different amounts of energy. The same rationale applies to mirror-image plane of the metal substrate, as can be shown
the gain in energy upon adding an electron (see Figure 4e). by electrostatic image-charge self-energy corrections assum-
ing perfect screening by the metal surface.[144,145,147,148]
Finally, the role of the Fermi level of the organic semicon- ii. Eventual chemical interaction between substrate and adsorb-
ductor layer shall be discussed briefly. For an (admittedly hypo- ate results in the formation of hybrid bands, which increases
thetical) intrinsic OSC layer it is in the middle of the funda- the effective bandwidths of the states derived from the or-
mental gap also at finite temperatures, due to the very narrow ganic semiconductor. The magnitude of that effect strongly
valence and conduction bands. Because of the large bandgaps depends on the type of interaction, as will be discussed below.
typically encountered in OSC materials, the position of the iii. Changes in the geometry of the adsorbate (e.g., due to interfa-
Fermi level will, however, vary significantly, even when only cial charge rearrangements or due to molecular distortions trig-
minute amounts of charge are transferred or when the layer gered by the formation of bonds between substrate and reactive
is only weakly contaminated by electrically active impurities. groups in the adsorbate—see below) can also modify the ioni-
Therefore, we will refrain from basing the following consid- zation energies and electron affinities of the adsorbate layer as
erations on the Fermi level of the OSC and its equilibration well as the step in electrostatic energy across the layer, ΔVLint.
with the substrate Fermi level.[137] Nevertheless, it should be iv. Most importantly, there will always be interfacial charge rear-
mentioned that the related charge-neutrality level has been rangements, whose origin will be discussed in the following
employed successfully for understanding interfacial level align- section. While the effects described in (i)–(iii) are usually di-
ment in the framework of the induced density of interface rectly attributed to the molecule (thus replacing ΔVLml with
states (IDIS) model.[138–140] This works because in the IDIS the corresponding quantity at the interface, ΔVLint), the con-
model, one does not rely on the density of states (DOS) of sequences of interfacial charge rearrangements for the en-
the undisturbed semiconductor, but considers the formation ergetic level alignment are often collected into the so-called
of hybrid states due to the interaction between substrate and bond dipole (BD). Strictly speaking, for the present consider-
adsorbate. ations, BD also contains changes in electrostatic energy asso-
ciated with geometric rearrangements of the metal substrate.
Here, mere relaxations of the substrate geometry typically
4.2. Electronic Structure of Interfaces have only a comparably minor impact. Conversely, surface
reconstructions, such as the appearance of adatoms can play
4.2.1. Metal–Organic Interfaces a more dominant role.[68,149–152] Note that BD, by definition,
includes all image-charge and image-dipole effects, both due
A useful gedankenexperiment for understanding the prop- to the bonding as well as, in principle, due to molecular di-
erties of a metal–organic interface is to start from the indi- pole moments. We, however, emphasize that provided that
vidual components discussed above (the metal substrate and ΔVLint arises from densely packed dipoles at sufficient dis-
the free-standing monolayer). As a first step, one then aligns tance from the substrate surface (e.g., dipoles due to polar
their vacuum levels, where for polar monolayers like the one tail-group substituents in self-assembled monolayers), there
depicted in Figure 4e, the vacuum level facing the substrate is no such dielectric screening of the molecular dipoles by the
(i.e., the “left” vacuum level) is the one to be used. Assuming metal substrate (cf., Section 2).
that the two constituents do not otherwise influence each other The key consequence of the bond dipole is that it causes a
(i.e., in the Schottky–Mott limit, SM, see Figure 5a,b), the level shift in the energy landscape, as discussed in Section 2. This
alignment expressed by the electron- and the hole-injection bar- is schematically shown in Figure 5c,d. Thus, the work func-
riers, EIBSM and HIBSM is simply given by tion of the combined interfaces is given as

EIBSM = Φsub − EA ml
left
and HIBSM = IEml
left
− Φsub (4) Φ = Φsub + ∆Φ = Φsub + BD + ∆VL int (6)

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (11 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 5.  Energy level alignment at interfaces between an organic semiconductor and a–d) a metal or e) an inorganic semiconductor. The situations
depicted in panels (a) and (b) correspond to the Schottky–Mott limit (vacuum-level alignment), while panels (c)–(e) show the equilibrium situation
including the bond dipole, BD, and in the case of the semiconductor substrate also a possible change in the band bending ΔBB. Notably, for the metal
substrate (panels (c) and (d)), we here assume a situation in which no electrons are transferred into the unoccupied states of the adsorbate layer (i.e.,
in which the electron affinity of the adsorbate layer is not sufficiently large). The case including charge transfer and the resulting Fermi-level pinning
situation will be discussed in detail for the adsorption on metals in Section 5.4. Conversely, for the semiconductor substrate (panel (e)) such charge
transfer is assumed as a prerequisite for the discussed (significant) increase of the band bending in the inorganic semiconductor. To obtain that charge
transfer, we consider the adsorption of a strongly electron accepting layer with a large electron affinity achieved in part by inverted surface dipoles of
the monolayer compared to the situation depicted in Figure 4d (and assumed in Figure 5a,c). In all schematics, the adsorption of organic monolayers
is considered, such that band-bending effects within the adsorbate layer can be neglected. The various indicated quantities are explained in detail in
the main text. Shaded orange areas correspond to states in the metal substrate and in the semiconductor, while blue areas correspond to states or
bands in the organic monolayers. Dark shading refers to occupied and light shading to unoccupied states.

where we adopt the sign convention that a bond dipole that from ΔVLint. In practice, Equation (6) can also be used as
reduces Φ is counted as negative.[21] Note that Equation (6) the equation defining the bond dipole. This is, e.g., done
assumes that the bond dipole drops essentially between in density-­functional theory simulations.[65,70,90] The situa-
the substrate and the adsorbate, i.e., it is spatially separated tion becomes more complex, when there is no clear spatial

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (12 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

separation between the charge rearrangements causing BD If one was rather interested in the work-function change due to
and the groups bearing the molecular dipoles. This might the interface as a whole (i.e., relative to a semiconductor with flat
occur, for example, for the adsorption of flat-lying polar bands), ΔΦ′, one also needs to consider the band bending prior
molecules. In such cases the two dipoles from Equation (7) to deposition of the adsorbate, ΔΦBB, and Equation (8) becomes
will couple and influence each other.[153] This can massively
change the adsorbate-induced work-function modification, ∆Φ′ = ∆ΦBB + ∆BB + BD + ∆VL int (9)
as will be discussed in detail below in the context of Fermi-
level pinning.
Overall, this yields the following expressions for electron- The overall situation (here assuming a strongly electron
and hole-injection barriers at metal–organic interfaces[65,70,90] accepting adsorbate layer without an intrinsic dipole) is
depicted in Figure 5e. A more in-depth discussion of the elec-
EIB = Φsub − EA int
left
+ BD and HIB = IEint
left
− Φsub − BD (7) tronic structure of the interfaces between organic semicon-
ductors and ZnO, as a prototypical example for an inorganic
Importantly, for the adsorption of polar molecules, again semiconductor substrate, can be found in ref. [115].
the electron affinity and ionization energy on the side of the
monolayer facing the interface (i.e., the “left” quantities) are
the relevant ones. 4.2.3. Organic Monolayers versus Multilayers and Doped Films
The above considerations focus on adsorbing a single organic
monolayer. Therefore, band bending within the OSC has so
4.2.2. Interfaces between Organic and Inorganic Semiconductors
far not been considered for establishing the energetics of the
interface. In order to understand (potential) band bending in
For such interfaces, the situation is complicated by the band
organic multilayers, it is useful to first consider the situation
bending within the inorganic semiconductor discussed in
for Schottky contacts between metals and inorganic semicon-
Section 4.1.2. Notably, as sketched in Figure 5e, neither the
ductors.[91] There, one finds that the width of the depletion
original surface band bending nor its modification through an
layer, over which the band bending occurs, is proportional to
adsorbate change the alignment between the frontier states in
1
the organic semiconductor and the band edges of the inorganic (where N is the space charge density in the depletion layer,
semiconductor right at the interface. Moreover, the adsorption N
of physisorbed or covalently bonded molecules causes only very which, within the abrupt junction approximation and in the sat-
local charge rearrangements, which do not notably affect the uration regime of doping, corresponds to the doping density).
surface band bending,[85,154] unless the surface chemistry is sub- This implies that for very small doping concentrations the band
stantially altered (e.g., by the removal of hydroxyl groups).[155] For bending occurs over comparably large distances. Consequently,
the adsorption of organic electron acceptors on n-type or electron for thin films that are smaller than the depletion layer, the band
donors on p-type materials, however, the semiconducting nature bending is essentially “cut off,” i.e., only the part of the band
of the substrate can have a tremendous impact. Provided that bending that fits into the spatial extend of the thin film is real-
the adsorbate depletes charge carriers (i.e., electrons or holes) at ized. Following these arguments, also for nominally undoped
the surface and in the bulk, this results in an additional band organic thin films, which are not excessively defective or where
bending, ΔBB. Importantly, the physics that leads to the forma- the defects are not electronically active, the shift of the elec-
tion of the BD (between the substrate and the adsorbate) and tronic states that occurs within a few ten nm from the electrode
the band bending ΔBB (within the substrate) are related, as (as the typical thickness of OSC layers in devices) becomes neg-
both occur due to charge transfer and the ensuing shift in the ligibly small.
electrostatic energy. Hence, both terms must have the same This suggests that band bending in organic semiconductors
sign. Moreover, in a combined system there is no clear defini- beyond the first monolayer is primarily relevant for less ideal
tion where the substrate ends and the adsorbate beings. Hence, situations, e.g., for (intentionally or unintentionally) doped
there is some ambiguity as to how to partition the total interface organic semiconductors[156,157] or when dealing with mate-
dipole between these two terms.[123] For the present discussion, rials having defect- or disorder-induced states in the gap.[158]
and as shown in Figure 5e, it is most useful to define ΔBB as the For example, Ishii et al.[159] observed depletion-layer widths
change of the substrate bands relative to the Fermi level (implic- of ≈100 nm for C60/metal interfaces, while they found flat
itly assuming that there is no hybridization between the molec- bands when using N,N′-bis(3-methylphenyl)-N,N′-diphenyl-
ular and the substrate states, which would make this distinction [1,1′-biphenyl]-4,4′-diamine (TPD) as semiconductor. These
impossible). flat bands they attributed to the high purity of the used TPD.
The adsorbate-related change of the semiconductor work Band bending for nominally pristine organic semiconductors
function then arises from the modification of the band bending has also been observed, when studying various interfaces with
due to the interfacial charge transfer, ΔBB, the bond dipole at the conjugated polymers[160–165] and when separating the metal
interface, BD, and potentially a shift in the vacuum energy due from the organic semiconductor using an insulating alkali-
to intrinsic dipoles of the adsorbate layer in the configuration it halide layer.[166] In such cases, this is typically explained by the
assumes at the interface, ΔVLint. The resulting equation reads filling of electronic states in molecules at some distance from
the interface.[161,162,166,167] Energetically, these states are found
∆Φ = ∆BB + BD + ∆VL int (8) “tailing” into the nominal gap and they are typically attributed

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (13 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

to structural imperfections and defects.[161,168–171] Notably, this material, and a charge accumulation with the corresponding
band bending due to the accumulation of additional carriers in band bending in the second organic layer (see Figure  6).[137]
the OSC beyond the first monolayer is fundamentally different Notably, in these cases, the evolution of the electrostatic
from the “conventional” band bending described above for potential depends strongly on which of the organic materials is
(doped) inorganic semiconductors. There, the necessary space- in direct contact with the substrate surface, which is a possible
charge region typically does not arise from accumulating addi- explanation, why the level alignment in organic heterojunctions
tional carriers but is rather the result of a depletion of mobile changes if the order in which the organic materials are depos-
carriers leaving behind uncompensated charged dopants. ited is changed.[26,174–177]
An instructive semi-quantitative analysis of the origin and Finally, it should be noted that “conventional” band bending
implication of the accumulation-triggered band bending at can be realized via intentional redox-doping of the organic
interfaces has been presented by Oehzelt et al.[172] They self-
consistently modeled the formation of an accumulation-­
generated space-charge layer and the resulting band bending
for different shapes and widths of the density of states in the
OSC material. Consistent with the band bending arising from
charge transfer into states tailing into the nominal gap, the
extent of the band bending in their model crucially depended
on the assumed shape of the DOS. For actual interfaces the
latter is determined by the prevalence of defects and the degree
of interaction between substrate and adsorbate. Overall, this
suggests that for multilayer organic films band bending (with
a spatially extended region of electrostatic energy variation)
is favored for weakly interacting systems and when there is
a small but non-negligible density of states in the nominally
forbidden gap.[172] Conversely, for highly pure, defect-free
organic semiconductors and for strongly coupled interfaces at
which the density of states is broadened significantly,[138–140]
the interfacial potential change is restricted essentially to the
first adsorbate layer.[172] In passing we note that an effect not
considered in the model by Oehzelt et al. that might modify
the actual situation is that charge transfer beyond the first
monolayer goes into molecules, for which the larger distance
from the substrate results in a reduced screening. This gives
rise to smaller electron affinities and larger ionization ener-
gies, which increases the energetic cost for charge transfer
beyond the first monolayer. Notably, such distance-dependent
polarization effects have also been suggested as an alternative
explanation for the dependence of the ionization energy on the
thickness of the adsorbate.[173]
When band bending in an OSC layer occurs, the total
interface dipole can be decomposed into a bond dipole at the
immediate metal–organic interface and the energetic shift
caused by the band bending.[167] This situation is analogous
to that discussed above for inorganic semiconductor sub- Figure 6. Charge accumulation (bottom panel) and level alignment
strates. Like there, one needs to distinguish between the ener- (top panel) of an organic/organic heterojunction in contact with a low
getic offsets between electronic states directly at the interface work-function electrode (EF = 2 eV). The substrate ends at z = 0, the first
and those far away from it. Again, the former offset remains organic material extends from 0–10 nm, and the second organic material
unaffected by band bending, while the latter is significantly from 10–20 nm. In the top panel, the orange boxes denote the valence
bands, the blue boxes the conduction bands, and the solid black line the
modified.[172]
evolution of the vacuum level linked to the electrostatic potential. Both
An interesting special case here are organic/organic organic materials are electron acceptors. The second organic material is a
heterojunctions. It has been argued[137] that there, the level stronger electron acceptor, indicated by its lower conduction band. In ther-
alignment may not be determined directly by interfacial modynamic equilibrium, little charge is transferred to the weak acceptor,
charge rearrangements. Rather, the heterojunction is in while a larger amount of charge is transferred to the strong acceptor.
thermodynamic ­equilibrium with a substrate that acts as charge This gives rise to the more extended space charge region in that material
reservoir. In appropriate material combinations (e.g., when first (see blue regions in the bottom panel). The corresponding potential evo-
lution (top panel) shows a quasi-linear increase within the weak acceptor
an inert material and then a strong electron acceptor is depos- and levels off with a strong parabolic evolution in the strong acceptor.
ited on a metal), charge flows from the substrate—across the Reproduced under the terms of a Creative Commons Attribution 4.0 Inter-
“first” organic material—into the second organic material. As a national License.[137] Copyright 2015, The Authors, some rights reserved;
result, there is a homogenous potential drop in the first organic exclusive licensee American Association for the Advancement of Science.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (14 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

s­ emiconductor,[18,178,179] which results in a sharply reduced configurations with a net dipole moment, typically self-assembly
depletion-layer widths.[180,181] An in-depth discussion of doped processes in combination with anisotropic bonding to the
adsorbate layers, however, goes beyond the scope of the pre- ­substrate are employed. The latter is needed to overcome the
sent review and we refer the reader to recent, more specialized repulsion between neighboring parallel dipoles and often relies
papers.[182–185] on the formation of strong (covalent) bonds between the sub-
strate and functional docking groups attached to the adsorbed
molecules.[35,47,49,50,133–136,186,187] Covalently bonded SAMs[188]
5. Dipole Layers and Interfacial comprise, for example, thiolate-[186,189–194] and selenolate-docked
molecules[195–198] on noble metals, or phosphonates[186,188,199–201]
Charge-Rearrangements
and silanes on oxides.[13,14,52,53,57,188,201] A review of the proper-
The next aspects that need to be discussed are the possible ties of such layers also beyond the aspects discussed here can,
origins of the interfacial dipole layers, which cause the shift for example, be found in refs. [36,202–204].
of the electrostatic energy. So far, we have described primarily For cleverly designed structures, also weaker interactions,
the two sources of more macroscopic origin, band bending and e.g., arising from van der Waals, vdW, forces can be sufficient
charge trapping. to align molecules such that their intrinsic dipoles are par-
In the present section we will focus on the origins of much allel.[78–80,205–207] This requires a molecular shape, where one
more local, microscopic dipoles with an extent of at the most particular molecular orientation has a larger adsorption energy
a few Å. As summarized in Figure  7, such dipoles can origi- than all others. Such a situation can, for example, be realized
nate from polar functional groups that are part of the adsorbed for shuttle-cock shaped phthalocyanine-based systems or for
molecules, or they arise from interfacial charge transfer pro- porphyrins and phthalocyanines bearing axial polar substitu-
cesses like the omnipresent Pauli pushback, charge rearrange- ents.[78–80,205–207] An example for such a situation is shown in
ments due to the formation of covalent bonds, or because of Figure  8, where for the Cl-up orientation of GaCl-phthalocy-
the adsorption of strongly electron donating or accepting mole- anine (GaCl-Pc) the vdW-contact area between substrate and
cules. In the latter case one has to distinguish between so-called adsorbate is maximized. For the less favorable Cl-down orien-
fractional and integer charge transfer, both of which commonly tation, the vdW attraction to the surface results in a signifi-
result in Fermi-level pinning. cant distortion of the molecular backbone. Due to the larger
average distance between the atoms in the substrate and those
constituting the adsorbate it also causes a significantly reduced
5.1. Adsorption of Polar Molecules interaction energy.[208] In spite of the small van der Waals
energy per atom, this effect is significant due to the resulting
One possible source of dipolar layers at interfaces is the large van der Waals contact areas comprising a sizable number
adsorption of molecules containing polar bonds and/or polar of atoms per molecule. In passing we note that for a priori less
functional groups. To align them in an ordered fashion in planar sub-phthalocyanines bearing polar ClB groups, the
van der Waals attraction to the surface is even strong enough
to essentially planarize the molecules with profound conse-
quences also for the electronic properties of the adsorbate.[209]
Interestingly, besides polar molecular systems, also poly-
mers bearing polar groups have been successfully applied as
surface modifiers to produce a variety of low work function
electrodes.[210]
A particularly intriguing example of polar, van der Waals
aligned adsorbate layers is encountered at the interface
between polar phthalocyanines and graphite: Upon
depositing OTi-phthalocyanine (OTi-Pc),[78] ClAl-phthalo-
cyanine (ClAl-Pc),[79] or OV-phthalocyanine (OV-Pc),[79]
Fukagawa et al. observed the buildup of a sizable surface
dipole during the ­formation of the first molecular monolayer.
An equivalent observation has been made by Blumenfeld
et al. for the adsorption of OV-naphtalocyanine on the same
substrate.[206] The dipole arises from a p ­ arallel alignment of
the polar OTi/ClAl/OV groups in the m ­ olecules in direct
contact with the graphite surface.[78] As shown in Figure  9,
this results in a linear shift of the vacuum level to higher
energies relative to the Fermi level with increasing ClAl-
Pc coverage (up to an effective coverage of ≈0.36 nm; see
white ellipse). When continuing the growth, the net dipole
moment of the adsorbate disappears due to the antiparallel
Figure 7.  Overview of the origins of interfacial dipole layers at inorganic– alignment of the molecules in the second layer (up to a nom-
organic interface. inal coverage of 0.7 nm).

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (15 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 9.  Map of the HeI UPS spectral intensities for ClAl-Pc d
­ eposited
onto graphite as a function of coverage. The feature found around
−1.2 eV at low coverage corresponds to the HOMO-derived band, while
the dotted line starting at 4.5 eV represents the vacuum level. As the
energy scale is aligned to the Fermi level, the latter also represents
the work function of the system. Adapted with permission.[79] Copyright
2011, American Physical Society.

to decompose on Ca[211] and partial decyanation has been


observed for tetracyanoquinodimethane (TCNQ) derived mole­
cules on Cu(100).[212] For the adsorption of polar molecules a
partial decomposition upon adsorption might then result in
the loss of the polar group. This has been observed, e.g., for
phthalocyanine-derived molecules bearing chlorine-containing
polar groups on Cu(111) surfaces: A final adsorption state of
the abovementioned GaCl-Pc/Cu(111) interface is a system
in which the Cl atom has been separated from the molecule
(see Figure 8c), which results in a significant reduction of the
molecular dipole. This also modifies the adsorption-induced
work-function change (albeit in a more complex way than
expected based solely on the molecular dipoles, as will be
discussed in Section 5.4).[208] Dechlorination processes have
also been observed for chloro-boron-subphthalocyanine (ClB-
subPc) on Cu(111), although there it has been argued that only
the molecules landing in Cl-down orientation undergo the
dechlorination.[213]

5.2. Interfacial Charge Transfer

Figure 8.  DFT-simulated structures of GaCl-Pc on the Cu(111) surface for A further possible origin of interfacial polar layers besides
a) the Cl-up and b) the Cl-down configuration, as well as c) for the dechlo- intrinsic molecular dipoles are charge rearrangements upon
rinated molecule. In the Cl-up case, the vdW-contact between substrate interface formation.[21,27] In this context, it should be men-
and adsorbate is maximized. The vdW attraction to the surface results
tioned that describing these charge rearrangements as dipoles
in a significant distortion of the molecular backbone. Reproduced under
the terms of the Creative Commons CC-BY license.[208] Copyright 2015, is a coarse simplification, as intramolecular charge distribu-
American Chemical Society. tions as well as interfacial charge rearrangements are typically
more complex than a simple separation of positive and negative
An aspect to keep in mind when molecules are deposited charges. Thus, when quantitatively deriving shifts in the inter-
on comparably reactive surfaces is that their chemical integ- face energetics, it might not be sufficient to apply Equation (1),
rity on the surface is not necessarily guaranteed. For example, but one will typically have to solve Poisson’s equation with the
tris(8-hydroxyquinolino)aluminum (Alq3) has been reported actual (plane-averaged) charge redistributions as source term.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (16 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Nevertheless, for more qualitative arguments, approximating The magnitude of the pushback can be sizable. Using n-tetra-
the actual situation as a dipole layer typically eases the discus- tetracontane on Au(111) as a prototypical physisorbed system,
sion without introducing serious artefacts. Mizushima et al. observed a work-function reduction due to
Charge rearrangements involving a direct coupling between pushback in excess of 1 eV.[215] They also provided a more detailed
the substrate and the adsorbate go hand in hand with the description of the interaction, involving, e.g., an adsorbate-induced
­formation of hybrid states, where one can distinguish between modification of the Shockley surface state of the metal in combi-
three different processes, which often occur simultaneously. nation with system-specific interfacial charge rearrangements.[215]
These are rationalized through electron sharing between substrate
and adsorbate involving bonding and antibonding CH states in
5.2.1. Pauli Pushback (Cushion Effect) the metal–molecule bond[216] and by including the hybridization
between metal d-bands and unoccupied molecular orbitals.[217]
When the bonding and antibonding of these hybrid states are In the context of Pauli pushback, one also has to keep in
simultaneously occupied, this results in a repulsion of the elec- mind that it is massively reduced, when the metal substrate is
tron clouds of substrate and adsorbate. In other words, due to not clean (e.g., covered by a thin hydrocarbon layer) or when
the Pauli principle, the electron clouds are pushed back into it is intentionally or unintentionally covered with a thin non-
the respective sub-systems.[21,214] This is shown in Figure 10 for metallic layer (e.g., an oxide).[24,218] In that case, the majority of
benzene on a Au(111) surface. Pauli pushback is particularly the pushback already happens when the original contamination
pronounced for metal substrates, as its magnitude depends on occurs and no further significant work-function reduction due
the (relative) polarizability of substrate and adsorbate.[214] As far to OSC growth is to be expected. On more practical grounds,
as the electronic properties of the interface are concerned, the such considerations also apply when growing OSC layers under
pushback of the electron cloud spilling out of a metal results high vacuum instead of ultrahigh vacuum conditions or when
in a significant reduction of the metal work function.[21] Strictly depositing the OSC from solution. There, the processing con-
speaking, this effect is not the consequence of forming a new ditions result in electrode contamination already prior to the
dipolar layer at the surface, but rather results from reducing the OSC growth. The resulting “technical” electrodes have a signifi-
already existing metal surface dipole discussed in Section 3. cantly reduced work function to start with, but when growing
OSC layers on top of them also the work-function reduction
due to pushback will be much smaller. The situation is funda-
mentally different when growing OSC molecules that strongly
bind to the substrate (like thiolates on Au), where it is possible
that contaminants are displaced as a result of bond formation
between substrate and docking groups.[193,219] In that case an
atomically clean metal substrate will be the ideal reference
system for analyzing the situation, even if the films have not
been grown under ultrahigh vacuum conditions.

5.2.2. Chemical Bonding to the Substrate

When, in contrast to the above situation, more bonding than


antibonding hybrid orbitals are occupied, a chemical bond
is formed between the substrate and the adsorbate. This is
typically associated with pronounced charge rearrangements,
which, as discussed above, give rise to a shift in the electro-
static energy landscape in analogy to the formation of a dipole
layer. Such chemical bonds are, for example, formed when
thiolates bond to coinage metal surfaces during the formation
of a self-assembled monolayer comprised of upright-standing
molecules.[186,189–194] Similar considerations apply to seleno-
Figure 10.  Visualization of Pauli pushback for a benzene molecule lying flat late-,[195–198] phosphonate-,[186,188,199–201] and silane-docked
on a Au(111) surface calculated employing dispersion-corrected density-
molecules.[13,14,52,53,57,188,201] Local charge rearrangements influ-
functional theory. Panel (a) shows the electron density of the metal (ρslab)
plus the charge rearrangements Δρ induced by the adsorption of benzene, encing the interface dipole also occur when flat-lying molecules
due to wave function orthogonalization (see main text: isovalue:0.00025 e− bearing reactive substituents adsorb on metals. Examples for
bohr−3)). Δρ is calculated as the density of the combined system minus the such reactive substituents are the CN groups in 2,3,5,6-tetra-
density of the isolated subsystem, i.e., Δρ = ρcombined − ρslab − ρbenzene. The fluoro-7,7,8,8-tetracyanoquinodimethane (F4TCNQ) adsorbing
displayed quantity ρslab + Δρ is, thus, equal to ρcombined − ρbenzene; hence, on coinage metals[152,220–223] or on g ­raphene/Ir(111),[224] the
in contrast to the commonly employed plots, like the one in Figure 11, the
carbonyl groups in 5,7,12,14-pentacenetetrone (P4O) bonded to
charge density of the pristine metal slab is not subtracted. Blue regions
correspond to positive and pink regions to negative values. To better visu- Cu(111) and Ag(111),[225] or the carboxylic groups upon inter-
alize the situation for the metal, in panel (b) only the positive values less face formation between perylenetetracarboxylic dianhydride
than 2.95 Å above the topmost Au layer are shown. (PTCDA) and Ag(111).[226–229]

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (17 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

5.2.3. Fractional Charge Transfer to Electron Donating or Electron substrates.[220–223,236] In this context, it should be mentioned
Accepting Molecules that it has been suggested recently that for TCNQ on Ag(111)
adatoms might be present on the surface, which fundamen-
When the molecules adsorbing on a substrate have strongly tally changes the adsorption geometry.[152] However, even in
electron donating or electron accepting character, the hybrid
orbitals formed at the interface are highly polarized, i.e., they
are quite strongly localized on one subsystem (the substrate
or the adsorbate). By filling, e.g., primarily the orbitals with
a dominant weight on the adsorbate, a net transfer of a frac-
tional number of electrons per molecule is realized. In passing
we note that the actual amount of transferred charge cannot
be uniquely quantified, as a “molecular charge” is no physical
observable in an extended system (just like an atomic charge
in a molecule). Nevertheless, approximate values can be deter-
mined employing various partitioning schemes.[230]
The main difference between charge transfer to adsorbed
donors and acceptors and the formation of covalent bonds
discussed in the previous paragraph is that here charge is not
rearranged only locally but is typically transferred to/from the
extended π-system of the adsorbate molecules. For donors,
the above-described fractional charge transfer (FCT) leads to
work-function reductions with the magnitude of the effect sig-
nificantly exceeding Pauli pushback. Typical examples are the
interfaces between tetrathiafulvalene[94,231–233] or viologen (deriva-
tives)[232,234] and a Au(111) substrate. For the methylviologen-
modified Au substrate, a work-function decrease as large as
2.2 eV has been observed. This resulted in a reduction of the
electron-injection barrier into a subsequently deposited OSC
layer by 0.6–0.8 eV,[234] compared to an interface where only Pauli
pushback is active. For strong acceptors, the electron transfer
to the adsorbate can surpass the inevitable pushback, such
that the net effect is a work-function increase due to molecular
adsorption. Such a behavior has been observed for many inter-
faces comprising adsorbate layers consisting, e.g., of F4TCNQ,
P4O, or PTCDA.[220–223,225–229,235–237] The adsorption of (compa-
rably weaker) acceptors can lead to the observation that electron
transfer to the adsorbate exactly cancels Pauli pushback with
the consequence that the work function remains almost unaf-
fected by the adsorbate. This, for example, happens at the inter-
face between hexaazatriphenylene-hexacarbonitrile (HATCN)
and Ag(111) up to monolayer coverage.[238] Notably, even the
deposition of such comparably weak acceptors as an interfacial
layer significantly reduces the hole injection barrier into subse-
quently deposited OSC layers.[239] The reason for that is that in
the absence of the interfacial layer, only pushback would occur
at the interface, which would significantly reduce the substrate
work function.
Whether a molecular adsorbate is a sufficiently strong donor
or acceptor to trigger the above-described electron transfer
mostly depends on its adiabatic ionization energy or electron
affinity. These quantities sometimes are also associated with
­so-called “polaronic” levels.[240] Particularly large IEs/EAs are Figure 11. a,b): Isodensity plots of the DFT-calculated charge density
realized, when the charge transfer triggers the conversion of rearrangements upon adsorbing an F4TCNQ layer on the Ag(111) sur-
an antiaromatic or quinoidal structure into an aromatic one,[225] face. Electrons flow from the dark grey to the light grey areas. c) Molecular
as such an aromatic stabilization is known to energetically orbital population analysis[242,243] (obtained by integrating the orbital-­
projected density of states up to the Fermi level) for F4TCNQ adsorbed
­stabilize the charged species.[241]
on Ag(111). The bands derived from the isolated molecular monolayer
For a detailed discussion of the charge-transfer mechanisms were taken as reference for the projection. The full (open) circles corre-
in the focus of this section, it is useful to take a closer look at spond to molecular orbitals occupied (unoccupied) prior to adsorption.
the situation encountered for F4TCNQ on flat coinage metal Adapted with permission.[221] Copyright 2009, American Physical Society.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (18 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

that case, the simulations on the flat substrate are particu- between substrate and adsorbate. This can be described by the
larly instructive for illustrating the fundamental aspects of the model discussed in the previous section for interfaces at which
charge-transfer mechanisms present at strongly coupled metal– substrate and adsorbate states hybridize. In the case of weakly
organic interfaces. coupled interfaces, there are no polarized hybrid states, whose
The isodensity plots of the charge-rearrangements occur- filling would result in a fractional net transfer of charge. There-
ring for F4TCNQ on Ag(111) are shown in Figure  11a,b. They fore, in such situations instead of fractionally charging every
reveal several interesting aspects: There is clear charge-density molecule (FCT), a fraction of the molecules receives an integer
accumulation in the π-system of the molecule (areas shaded in charge, i.e., one observes integer charge transfer (ICT).[24,158]
light gray), with the nodal pattern of the charge rearrangements Depending on the intermolecular coupling, the charge might
strongly reminiscent of the lowest unoccupied orbital (LUMO) of also be delocalized over a few molecules.[248] The FCT and ICT
F4TCNQ. This means that hybrid orbitals formed between the scenarios are schematically compared in Figure  12a, where in
molecular LUMO and the sp-band of Ag become occupied. These the case of FCT translational periodicity is preserved, while it
hybrid orbitals are mostly localized on the F4TCNQ molecule (cf., is broken in the ICT case. Notably, the ICT charge distribu-
above discussion). The filling of LUMO-derived states in F4TCNQ tion should not be viewed as static, but the additional integer
upon adsorption is also supported by the molecular orbital popu- charges can tunnel between adsorbate molecules, unless they
lation analysis, which testifies to a nearly 90% effective filling of are stabilized on certain sites due to defects or by lattice relaxa-
the LUMO upon adsorption (see red circle in Figure 11c). tions of the substrate.[249]
At the same time, there is charge density depletion (areas Traditionally, ICT dominates i) for interfaces formed
shaded in dark grey) in the region of the σ-system, which can under ambient conditions by solution-processing of the
be associated with a relaxation of the electronic states within organic adsorbates, ii) when using metal substrates passi-
the molecule in reaction to the filling of the LUMO. Notably, vated by a native oxide layer (e.g., Al) or by an organic inter-
in the same spatial region also Pauli pushback due to the overlap layer, iii) when depositing molecules on conductive oxides, or
between the σ-electrons of F4TCNQ and the electron density iv) when using a conducting polymer like poly(3,4-ethylenedi-
tail from the Ag surface occurs. Considering, however, that in oxythiophene) polystyrene sulfonate (PEDOT/PSS) as a sub-
similar plots for the adsorption of strong donors, one observes strate.[177,240,252–256] More recently, ICT has also been observed
electron accumulation in the σ-region,[232] we deem the relaxa- in simulations (see Figure 12b)[250] and experiments,[257] when
tion effect to dominate. Particularly strong electron depletion is intentionally separating a metallic substrate from an electron
observed locally in the region of the σ-states associated with the accepting adsorbate layer by a dielectric layer with a thickness
CN groups. This transfer of electrons from the molecule to of only a few monolayers (provided it does not prevent charge
the substrate occurs from deeper lying σ-states localized on the transfer altogether).[258] A particularly interesting situation has
CN groups, as can be concluded from the reduced effective been computationally predicted for F4TCNQ on the (1010 ) sur-
occupation of these states in the adsorbate (see blue ellipse and face of ZnO. There, FCT due to the bonding of the CN groups
rectangle in Figure 11c). Due to the vicinity of the CN groups to the Zn atoms at the surface (affecting all molecules in
to the Ag surface, these states hybridize so strongly with the Figure 12d) coexists with ICT into the π-system of only a frac-
metal d-bands that a significant fraction of the resulting hybrid tion of the adsorbed molecules (molecule 1 in Figure 12d).[251]
states come to lie above the Fermi level.[220,221] As a consequence Whether charge transfer occurs at all for a specific material
of this superposition of charge forward and backward donation, combination, again depends on the adiabatic ionization ener-
in spite of the LUMO-derived band being nearly completely gies or electron affinities of the adsorbate layer, but also on
filled, the net charge transfer between the metal substrate and the structure and density of the molecular overlayer, as well as
the F4TCNQ layer amounts to only ≈0.56 e− per molecule (as on the thickness of the dielectric layer.[250] Whether the charge
estimated from the cumulative charge rearrangements between localizes or not (i.e., whether FCT or ICT occurs) depends
the sub-systems, as described in detail in ref. [221]). on the degree of hybridization and the strength of the lattice
Finally, it should be mentioned that in recent years also relaxation.
mixed donor/acceptor adsorbates on metal surfaces have been The consequences of an ICT situation are manifold: For
investigated.[244,245] In such cases the question arises, whether instance, the coverage-dependence of the adsorbate-induced
intermolecular or molecule-to-substrate charge transfer domi- work-function change distinctly differs between ICT and FCT
nates.[27] Answering that question is tightly linked to the issue, scenarios (see Figure 12c).[250] Notably, in the ICT case the
whether the mixed adsorbate layers phase segregate[246] or mix maximum adsorbate-induced work-function change is observed
homogeneously.[247] A detailed discussion of the properties of at comparably low coverages (0.4 monolayers in the example
mixed donor/acceptor adsorbate layers, however, goes beyond shown in Figure 12c). This can be explained by the fact that a
the scope of the present manuscript. Therefore, we refer the charged molecule raises the electrostatic potential in its vicinity.
interested reader, for example, to the discussion of this issue in Consequently, the LUMOs of neighboring molecules are often
the review by Otero et al.[27] shifted above the substrate’s Fermi level, removing the driving
force for further charge transfer. Therefore, in the case of ICT
one must not associate the maximum in the work-function
5.2.4. The Integer Charge-Transfer (ICT) Situation change with the closure of the monolayer.
Moreover, as bond lengths and charge distributions vary
Establishing thermodynamic equilibrium typically requires a between neutral and charged molecules, X-ray photoelectron or
noninteger number of electrons per molecule to be transferred infrared spectra of monolayers experiencing an ICT situation

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (19 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 12.  a) Schematic comparison between the ICT and FCT scenarios for charge transfer at organic/inorganic interfaces. b) Calculated electron-
density difference upon adsorption of tetracyanoethene (TCNE) on Cu/NaCl. For the sake of clarity, only positive values are shown for molecules within
the unit cell and the substrate atoms are omitted. c) Evolution of the TCNE-induced work-function modification as a function of coverage for the ICT
(red) and FCT scenarios (black). The FCT scenario is realized by performing the simulations using a GGA functional suffering from a many-electron
self-interaction error, which overdelocalizes charges.[250] d) Charge rearrangements within a surface supercell containing four F4TCNQ molecules on
n-doped ZnO, averaged along the line of sight. Charge flows from the red to the blue areas. (b,c) Adapted with permission.[250] Copyright 2015, American
Chemical Society. (d) Adapted with permission.[251] Copyright 2019, American Chemical Society.

should correspond to the superposition of the spectra of neu- 5.3. Characterizing the Charge-Transfer Situation
tral and charged species. As in the FCT case only one (frac-
tionally charged) species prevails, the distinction between FCT A parameter that is commonly used in literature to characterize
and ICT scenarios should be comparably straightforward in the implications of charge-transfer for the electronic properties
experimental studies, provided that the spectral resolution is of the interface is the “slope parameter,” S. It originates from
sufficient. Complications might arise from the timescales of Schottky–Mott theory and for a given (organic) semiconductor
the specific experiments due to the dynamic nature of the ICT connects the work function of the bare substrate, Φsub, with the
situation with transferred charges hopping between adsorbed work function after adsorbate deposition, Φ[22,24,138,158,240,252,263,264]
molecules.
Alternatively, the coexistence of neutral and charged mole­ dΦ
S= (10)
cules on the surface in the ICT case should be detectable by dΦSub
scanning tunneling spectroscopy. This is frequently the case
for metal atoms separated from a metal substrate by a thin For the case of vacuum-level alignment (the Schottky–Mott
dielectric layer, e.g., Au on Cu/NaCl.[259] Recently, this was limit), S becomes 1. This also applies, when the same bond
also observed for organic adsorbates, specifically pentacene dipole forms on different substrates. A possible scenario for
on a Cu/NaCl[260] and on a Ag/MgO substrate[261] (although the latter could be the prevalence of substrate-independent
in the former case, the pentacene was deliberately charged Pauli pushback. Conversely, as soon as the adiabatic ioniza-
with the scanning tunnelling microscope tip). In simulations, tion energy of the adsorbate layer drops below the substrate
such an unambiguous distinction between the FCT and ICT work function (or the adiabatic electron affinity rises above),
situations is a priori not possible. This is a consequence of additional charge has to be transferred between substrate and
charge over(de)localization resulting from the many-electron adsorbate. The resulting additional dipole layer then reestab-
self-interaction error, respectively its overcompensation, as lishes thermodynamic equilibrium. Provided that the electron
discussed in detail in ref. [262]. transfer to (from) the adsorbate layer does not appreciably

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (20 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

change its electron affinity (ionization energy), this results in to that ­performed in Section 4. We start from the electronic
S  = 0, i.e., the system work function becomes independent of structure of the isolated metal substrate, described by its work
the substrate work function. This process is typically termed function, and the isolated molecular monolayer, character-
Fermi-level pinning (and will be described in more detail in the ized by the monolayer ionization energy and electron affinity
following section). (see Figure  14a). Upon contact formation, screening by the
Abrupt transitions between S = 0 → S = 1 → S = 0 are, indeed, metal substrate reduces the fundamental gap; moreover, the
commonly observed in ICT situations,[24,158,240,252–254,265,266] energy levels of substrate and adsorbate are shifted relative to
which is shown in Figure 13 for a series of solution-processed each other due to Pauli pushback. Provided that the resulting
organic semiconductors on metals, metal oxides, semicon- screened adiabatic ionization energy is smaller (or the adiabatic
ductors, and conducting polymers.[267] Also in the FCT case electron affinity larger) than the work function of the metal
S  = 0 is expected as soon as Fermi-level pinning sets in, as including pushback, charges are transferred. The resulting
exemplified by PTCDA on various substrates including coinage dipole layer triggers a shift of the energy levels such that, in a
metals,[227,266,268] and F4TCNQ on Ag and Au.[221,236] Neverthe- single-particle picture, the Fermi level of the substrate aligns
less, for a variety of material combinations, slope parameters with the partially filled band of the adsorbate layer (Figure 14b).
between 0 and 1 have been observed experimentally.[22] Also Then, the work function of the combined substrate/adsorbate
in self-consistent electrostatic models, S-values between 0 and system is essentially determined by the electronic structure of
1 have been calculated upon gradually changing the shape of the adsorbate and, thus, becomes independent of the substrate
the density of states of the adsorbate from Gaussian to Lorent- work function (S = 0, see above).
zian.[172] This suggests that the degree of hybridization as well When analyzing the properties of charge-transfer interfaces
as the presence of defect states at the interface (both impacting in the context of Fermi-level pinning it is worthwhile consid-
the interfacial DOS) can influence Fermi-level pinning, calling ering the following aspects:
for a more in-depth discussion of this effect.
i. The nature of the electronic state at which pinning occurs
has been somewhat controversially discussed in literature.
5.4. Fermi-Level Pinning As mentioned already above, it is undoubtedly not deter-
mined by the vertical ionization energy (electron affinity)
As mentioned above, Fermi-level pinning typically comes of the adsorbed monolayer that would, e.g., be measured by
into play when charges are transferred between substrate and (invers) photoelectron spectroscopy (UPS and IPES). Rather,
adsorbate in order to establish thermodynamic equilibrium. electronic and geometric relaxation effects have to be taken
To discuss its fundamental aspects and its implications, it is into account.[24,240] For ICT cases, this has been described
again useful to consider a gedankenexperiment analogous by pinning at so-called integer-charge-transfer or polaronic
states.[24] Beyond the relevance of relaxation effects, it has
been shown that in many cases pinning already occurs
at defect or metal-related intragap states,[162,168,170,266,269]
that are found in the tails of the respective densities of
states reaching deeply into the gap. Moreover, for specific
shapes of the density of states, band bending has been
suggested (see above),[162,172] which modifies the pinning
situation.
ii. The amount of transferred charge depends on the magni-
tude of the energy offset (prior to equilibration) between the
metal Fermi level and the OSC state at which pinning occurs.
The larger that offset, the larger the interface dipole has to be
to align the levels. For a given charge-transfer distance (i.e.,
adsorbate geometry), a larger offset, thus, means that more
charge needs to be transferred to establish thermodynamic
equilibrium. Additionally, the geometric structure of the
interface plays a role: The (average) amount of transferred
charge per molecule decreases with increasing molecular
coverage and with increasing adsorption distance. In both
instances less charge per molecule needs to be transferred to
induce the step in the electrostatic energy between substrate
and adsorbate that is necessary to establish thermodynamic
equilibrium.
Figure 13.  Measured dependence of the work function of substrates cov- iii. Even if the final work function in the case of Fermi-level pin-
ered by an organic semiconductor (here termed ΦORG/SUB) on the work
ning depends only on the electronic structure of the adsorbate
function of the pristine substrate (ΦSUB). OSC films have been deposited
from solution on metals, metal oxides, semiconductors, and conducting layer, one has to keep in mind that the IE and EA of that layer
polymers. Reproduced with permission.[267] Copyright 2014, Wiley-VCH will change with charge transfer. In a single-particle picture,
Verlag GmbH & Co. KGaA. this can be understood from a partial filling of the frontier

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (21 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

bands. Consequently, for interfacial den-


sities of states smeared over comparably
wide energy ranges, a situation with S ≠ 0
is to be expected. Such a situation could, for
example, be caused by a particularly strong
hybridization between substrate and ad-
sorbate states (leading to very wide bands)
or by the presence of defect states at the in-
terface.
iv.  In cases in which there are additional di-
poles located between the substrate and
the adsorbate orbital(s) involved in the
charge transfer, the potential steps caused
by these dipoles will (in a first approxima-
tion) not change the overall work function.
Due to the pinning situation, its impact is
compensated by additional charge trans-
fer.[153] This is schematically shown in
Figure 14c,d, where the additional dipole at
the left side of the SAM (green arrow) mas-
sively changes the energy landscape in the
vacuum-level alignment case (Figure 14c).
As soon as thermodynamic equilibrium
has been established, this extra dipole is
compensated by additionally transferred
charges (and a correspondingly larger bond
dipole; longer red arrow) (Figure 14d).
Such a behavior has been shown explicitly
in simulations of F4TCNQ on Ag(111),
where bending the polar CN groups
toward the substrate leaves the work
function of the system essentially un-
changed.[153] Conversely, when the polar
substituents are bent upward such that the
respective dipoles come to lie above the or-
bitals at which pinning occurs, a sizable
impact of the dipole on the work function
is observed.[153] Equivalently, when the
Fermi level of the substrate is pinned at
the π-system of an upright-standing SAM,
varying the docking group does not impact
the SAM-induced work-function change,
while adding polar tail-group substituents
allows “beating” the limit set by Fermi-
level pinning (see Figure 14e).[153] The in-
dependence of the system’s work function
on dipoles between the substrate and the
Figure 14.  a–d) Schematics of the interfacial level alignment for a substrate/adsorbate combi- adsorbate can make the determination of
nation for which in thermodynamic equilibrium Fermi-level pinning occurs. a) and c) display the adsorbate configuration based on the
the vacuum-level alignment case, where the lowest unoccupied level lies below the Fermi level
of the substrate. In (a), the surface dipoles of the adsorbed layer are symmetric and oriented
observed electronic properties highly am-
such that the electron affinity and ionization energy of the adsorbate are increased. In (c), a biguous: For example, when studying Ga-
much larger surface dipole is assumed at the side of the adsorbate facing the substrate (green Cl-Pc on Cu(111) we calculated essentially
arrow). b) and d) show the situation after establishing thermodynamic equilibrium, where identical work functions for a configura-
the asymmetric dipole in (d) causes a larger amount of charge transfer, which results in a tions with the polar GaCl group aligned
massively increased bond dipole (see red ellipse and red arrow), such that the final adsorbate- such that the Cl atoms points toward the
induced work-function change is the same as in (b). e) Calculated adsorption-induced work-
substrate and for dechlorinated molecules,
function change for a Fermi-level pinned, terpyrimidine-based self-assembled monolayer
(chemical structure shown on the left) upon changing the docking group (black squares) and in which this molecular dipole is absent
upon adding polar tail-group substituents (red circles). (e) Reproduced with permission.[153] (see Figure 8). Conversely, a significant
­Copyright 2010, American Chemical Society. impact of the GaCl dipole was obtained

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (22 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

for intact molecules with upward-pointing Cl atoms.[208] Con- Section 4.1.3).[130–132] This is schematically shown in the cen-
sequently, the measured work function allowed ruling out the tral panel of Figure 15a. In particular, electron withdrawing
latter configuration, but could not distinguish between the substituents in the periphery of the molecule (due to the as-
two former ones. This was possible only by an in-depth com- sociated local dipoles) will typically increase IE and EA, while
parison between calculated[208] and m ­ easured[270] ­adsorption electron donating substituents will decrease them. There-
distances (which supported the existence of dechlorinated fore, the work function of the combined metal/adsorbate
molecules). In view of the recent discussion about ad-atoms system in the case of Fermi-level pinning strongly depends
at the F4TCNQ/Ag(111) interface,[152] we note that due to the on molecular orientation. As depicted in the right panel of
compensation discussed in this paragraph, the presence of Figure 15a, one might even encounter the situation that for
surface adatoms will substantially affect the molecular geom- one of the molecular orientations Fermi-level pinning pre-
etry, but will likely have a surprisingly small and often almost vails (here, for the upright standing molecules), while in the
vanishing effect on the total work function of the system. other there is vacuum-level alignment (together with push-
v. Another practically highly relevant aspect to consider in the back; here, for the flat lying molecules). As a general conse-
context of Fermi-level pinning is that the ionization energies quence of this effect, the reorientation of molecules during
and electron affinities of adsorbate layers depend strongly film growth (e.g., as a consequence of increasing coverage)
on the molecular orientation due to surface dipoles (cf., can massively change the measured sample work function,

Figure 15.  a) Electronic structure of an interface between a metal substrate and an upright-standing (top panels) and a flat-lying (bottom panels)
adsorbate layer consisting of molecules bearing polar substituents, but no net dipole moment. The central panel shows the electronic structure of
the hypothetical free-standing monolayers highlighting the dependence of the monolayer IEs and EAs on the molecular orientation. Upon forming the
contact with the metal (right panels), for the displayed example, Fermi-level pinning is found for the standing monolayer, while the smaller electron
affinity of the lying monolayer results in vacuum-level alignment. For the sake of clarity, pushback, bandgap renormalization, screening, and relaxa-
tion effects (discussed in the main text) are not accounted for in the right graphs. b) Coverage-dependent work function of HATCN deposited onto a
Ag(111) substrate. c) Coverage-dependent work function of COHON on Ag, Cu and Au surfaces. (b) Reproduced with permission.[238] Copyright 2010,
American Physical Society. The data in panel (c) are the same as in ref. [271].

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (23 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

even when the molecules themselves do not possess a net Figure 3c). This is consistent with the common experimental
dipole moment. As, to the best of our knowledge, this effect observation that even at interfaces at which Fermi-level pin-
has not been comprehensively described in literature, it shall ning occurs, the adsorbate-induced work-function change dis-
be discussed in more detail in the following paragraph. plays a pronounced coverage dependence at very low nominal
As an example for a reorientation-triggered work-func- layer thicknesses.
tion modification, the coverage-dependent work function of Notably, also the aggregation of molecules into islands
HATCN on Ag(111) is shown in Figure 15b:[238] For nominal can result in an apparent coverage dependence of the
coverages up to 3 Å, a flat-lying (Fermi-level pinned) mono­ measured work-function change. This is in particular the
layer is formed. For larger coverages, a variety of experi- case, when sections of the substrate remain uncovered
ments suggest that the molecules rearrange.[238] This results (e.g., for Volmer–Weber type film growth) and when area
in an essentially upright-standing layer with a significantly averaging experimental techniques are employed in the
increased monolayer electron affinity due to the polar CN experiments (e.g., SECO-measurements in photoelectron
substituents. As Fermi-level pinning prevails, the resulting spectroscopy experiments or Kelvin probe; see discussion
sample work function is increased by nearly 1 eV. For mole- in Section 4).
cules with a particularly high dipole density in their periphery, vii. As discussed above, in the case of Fermi-level pinning, the
such effects can even be larger, as has been observed for final work function is only determined by the adsorbate prop-
doubly NO2-substituted pyrenetetraone on Ag(111). There, for erties. This implies that, for adsorption on (inert) semicon-
upright-standing molecular layers, work functions in excess ductors, the doping concentration (which is a substrate prop-
of 6.1 eV have been measured.[272] A particularly interesting erty) does not influence the final work function. Since the
situation is encountered for 1,2,5,6,9,10-coronenehexone initial work function of the substrate is typically only weakly
(COHON) on coinage metals (see Figure 15c):[271] While the dependent on the doping concentration (particularly for
work functions of the pristine substrates and at very low highly doped semiconductors, which are commonly used as
coverages of COHON on Au, Cu, and Ag are significantly electrodes), also ΔΦ- ‘hardly depends on it.[123,273] Neverthe-
different, at a nominal coverage of ≈6 Å (which is attributed less, the relative contributions of the BD and the adsorption-
to roughly a closed monolayer of lying molecules) the work induced band bending, ΔΒΒ, (cf. Equation (9)) do depend on
functions of COHON/Ag and COHON/Cu become virtu- the doping concentration: The smaller the amount of free
ally identical. This suggests Fermi-level pinning at those two charge-carriers, the larger the relative contribution of ΔΒΒ
interfaces, while on Au pinning apparently does not occur yet to ΔΦ’ (and the smaller that of BD).[273] In other words, in
and the electronic structure of the interface is dominated by the case of Fermi-level pinning the overall potential shift be-
Pauli pushback. The situation changes for even higher cover- tween substrate and adsorbate is a fixed quantity. However,
ages. For nominal thicknesses exceeding 20 Å, the work func- the larger the amount of free carriers (i.e., the larger the
tions of all three interfaces are essentially identical and, at the doping concentration) is, the smaller is the fraction of the po-
same time, larger than for the pinned case at 6 Å. This would tential drop that occurs within the substrate (i.e., the smaller
be fully consistent with Fermi-level pinning for all three sub- ΔBB and the larger BD). Since, as discussed in Section 4.2.2,
strates at an essentially upright standing COHON layer with only BD affects the relative offset of the substrate bands and
a significantly increased electron affinity due to the polar car- the adsorbate states directly at the interface, the doping con-
bonyl groups in the molecular periphery. Interestingly, at least centration indirectly affects the level alignment. This is in-
for the COHON/Cu interface there is indeed further spec- deed corroborated by quantum-mechanical simulations.[273]
troscopic evidence for such a coverage-dependent reorienta- For practical applications, however, two words of caveat
tion.[271] All these examples indicate that coverage-dependent are required: First, as discussed above, in realistic wide-
rearrangements of organic molecules on metal surfaces have bandgap semiconductors, the maximum band bending that
massive consequences for their electronic properties. At this can be obtained is often limited by surface states (acting
stage we speculate that such rearrangements might occur as deep donors). Due to the high concentration of deep
much more often than usually assumed. donor states, the corresponding charge-rearrangements are
vi. The above discussion deals with the impact of coverage be- usually very short ranged and, hence, are usually factored
yond a (flat-lying) monolayer. Coverage, however, impacts into BD (rather than ΔΒΒ). In practice, this means that the
Fermi-level pinning also in the sub-monolayer case. It has relative contributions of BD and ΔΒΒ, and hence the level
already been discussed in Section 3.2.3 (Figure 3c) that for alignment, is less doping-dependent than one might ini-
charge-transfer layers with comparably large dipole to dipole tially assume.
distances, the electrostatic energy at the position of the mol- Second, the strong doping dependence occurs only for
ecule undergoing charge transfer and the energy far away the adsorption on inert semiconductors, where the charge-
from the interface differ considerably. The former, however, transfer is driven by the difference of the electron affinity
determines the IE and EA of the adsorbate, while the latter (or ionization energy) of the organic adsorbate and the
is relevant for the work function. As pinning still aligns the Fermi level of the substrate. The situation changes funda-
Fermi level of the substrate with the states in the molecule, mentally for strong hybridization between adsorbate and
this results in an adsorbate-induced work-function change the substrate valence or conduction bands. There, the situ-
that is smaller than in the full coverage case. The difference in ation is more reminiscent of the FCT scenario on a metal
the two electrostatic energies diminishes with decreasing di- (including very short-ranged charge rearrangements, that
pole to dipole distance, i.e., for more tightly packed layers (see lead to an absence of band bending).[274]

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (24 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

6. Impact of Dipolar Layers on Physical


Observables-Contact Resistance, Core-Level
Excitation, and Monolayer Transport
To comprehensively discuss the impact of dipolar layers on
physical observables typically measured at hybrid interfaces, in
Figure  16 we show the (comparably complex) situation of an
upright-standing adsorbate layer containing i) an embedded
dipolar (ED) element as part of the molecular backbone and ii)
a polar tail-group dipole (TD). iii) Moreover, the molecules are
covalently bonded to the substrate, giving rise to yet another
dipole layer, the bond dipole (BD). Upright standing molecules
are chosen for the present discussion, as in them the BD-, ED-,
and TD-dipole layers can typically be identified unambiguously.
To understand how quantities like the work function, the ener-
getic positions of the frontier states, and core-level binding
energies in this system are influenced by the three dipole
layers, in the following a (hypothetical) photoelectron spectros-
copy experiment performed at this interface shall be discussed.
What will be crucial for each of the energies is, how many of
the dipole layers the photoelectron has to cross, before it can
leave the sample, respectively, before it reaches the detector:

i. The work function, Φ, is determined by measuring the


SECO, i.e., by the photoelectrons with vanishingly small en-
ergies that barely make it out of the sample (see green ar-
row to the left). Consequently, Φ is shifted by the steps in the
electrostatic energies due to all three dipole layers (BD, ED,
and TD). This has motivated the application of polar SAMs in
(optoelectronic)devices for tuning electrode work functions
and carrier-injection barriers (see Section 2).[2,35–39,42–45,49,51]
ii. The situation is different for the core-level binding energies.
They are measured relative to the Fermi level of the substrate.
Thus, they are affected by the dipole layers between the re- Figure 16.  a) Schematic illustration of the energy level alignment for a SAM
spective atom and the surface of the substrate. Consequently, consisting of molecules containing an embedded dipole (ED) and a tail-
for the atoms in region (1) of the adsorbate (see Figure 16), group dipole (TD). The resulting aligned dipole layers induce two steps in
the electrostatic energy landscape. A third step is caused by the bond dipole
only the bond dipole BD matters, while for those in section (2)
(BD) resulting from the docking group and bonding-induced charge rear-
also the embedded dipole layer ED is relevant. The tail-group rangements. The work function of the SAM-covered substrate, Φ, in the
dipole TD, due to the locality of collective electrostatic effects present example is reduced by all three dipole layers. The core and valence
(see Section 3), has no impact. Consequently, the inclusion levels of the bottom segment (1) and top segment (2) are shifted relative
of dipoles into adsorbate layers results in electrostatically in- to each other by the embedded dipole layer (ED). Relative to the Fermi level
duced shifts in the kinetic energies of the photoelectrons, as of the substrate, they are additionally shifted by the bond dipole (BD). The
green arrows within the monolayer symbolize the XPS measurement pro-
is shown in the right part of Figure 16a. These “electrostatic”
cess with the incident photon energy hν. EF denotes the Fermi level, which
XPS shifts are superimposed on the commonly observed in thermodynamic equilibrium is the same on the sample and detector sides
chemical shifts due to differences in the atomic charge densi- of the setup[275] (in contrast to the vacuum energy Evac). The shift in elec-
ties resulting from neighboring atoms. The electrostatic shifts trostatic energy due to the embedded dipoles results in two different meas-
can easily be in the range of one eV, as it is shown for ester- ured electron kinetic energies at the detector (Ekin1 and Ekin2). Valence levels
containing alkylthiolates in Figure 16b.[136,276,277] Both, in the plotted in red refer to states localized on the top- and bottom-segments
simulated and measured spectra contained in that plot, one of the SAM, while valence levels plotted in blue symbolize the situation of
delocalized states. For the sake of clarity, we assume an infinitely extended
observes a double-peak structure for the main C1s features sample and detector, i.e., no distinction is made between the vacuum level
between −284 and −286 eV binding energy. Based on the den- directly right of the sample and at a distance much larger than the sample
sity functional theory (DFT) simulations contained in the left dimensions, as this would not affect the difference in the kinetic energies of
panel of Figure 16b, these features can be unambiguously the photoelectrons. b) Left: Calculated C1s core-level energies in an ester-
assigned to the chemically identical CH2 groups left and right containing alkylthiolate SAM. The rigid shift between the core-levels of the
of the ester moiety. Consequently, the difference in the as- CH2 carbon atoms left and right of the ester layer is caused by the regular
arrangement of the densely packed ester dipoles. Right: Measured and cal-
sociated binding-energies is of purely electrostatic origin and
culated XP spectra for the SAM depicted in the left panel. The double-peak
is caused by the dipoles of the embedded ester groups. This structure associated with core-level excitations of the CH2 carbons is clearly
can be shown, for example, by diluting the dipole density in visible. (b) Reproduced under the terms of the Creative Commons CC-BY
a hypothetical low-coverage SAM, in which the molecules are license.[276] Copyright 2016, American Chemical Society.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (25 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

kept in their upright-standing orientation.[276] Similar obser-


vations have been made also for aromatic SAMs containing
embedded pyrimidine dipoles.[50]
In that spirit, XPS can be used as a probe for shifts in
the local electrostatic energy within an adsorbate layer. Due
to the sensitivity of core-level binding energies to the local
­electrostatic energy, XPS measurements can also be used
to probe inhomogeneities of the potential at the immediate
surface of the sample.[100] In a similar spirit, they can be
applied to probe the uniformity of mixed adsorbate layers,
provided that the different constituents contain chemically
identical atoms, but at the same time shift the electrostatic
energy differently (e.g., via embedded dipoles incorporated
in different orientations).[101] When associating core-level
binding energies with the local electrostatic energy, care
has to be taken for atoms close to the interface. There, the
electrostatic shifts can be masked by chemical shifts due to
interfacial charge rearrangements, by lateral potential inho-
mogenieties, or by shifts resulting from the adsorption-
height dependence of screening by the metal or semicon-
ductor substrate.
iii. As far as the frontier levels are concerned, their energetic po-
sitions relative to the Fermi level are shifted by the bond di-
poles, as discussed already in Section 4.2. Concomitantly, the
bond dipoles also modify electron- and hole-injection bar-
riers for carriers into the SAMs. Conversely, the tail-group
dipoles do not impact the level alignment, due to the locality
of the dipole-induced step in the energy.[70,90] The situation
for embedded dipoles is somewhat more involved: Provided
that the valence states are fully localized either left or right of
the embedded dipole layer, e.g., by breaking the conjugation
or by using extended semiconducting segments minimiz-
ing the localization energy (see ref. [278]), a situation analo-
gous to that discussed for the core-levels is encountered (see
“red valence states” in Figure 16a). In that case, all states are
shifted relative to the Fermi level by BD and the states right
of the embedded dipoles are additionally shifted by ED.[278]
Such a situation is shown in Figure  17. For the energetic
positions of the frontier states relative to the vacuum level,
Figure 17.  a) Contour map of the local density of states (LDOS) along the
the relevant dipole layers are TD and ED. Notably, an analo-
molecular backbone for a SAM consisting of two tolane units separated
gous rigid connection between core-level and frontier-level by a polar section consisting of two methyl-substituted pyrimidine rings.
energies has also been observed for the adsorption of mixed The plotted quantity has been obtained by integrating the LDOS in energy
donor/acceptor layers.[279] windows of 0.1 eV over the plane parallel to the Au surface within a unit
cell. b) Isodensity plots of the charge densities associated with the highest
When the electronic states are only partially localized by the occupied and lowest unoccupied states within the semiconducting units
of the model SAM on Au (the scale at the bottom of the plots gives
embedded dipoles, BD and TD still play the same role as above.
the energies of the states relative to the Fermi level). Reproduced under
Also ED still either stabilizes or destabilizes the states relative the terms of the Creative Commons CC-BY license.[278] Copyright 2015,
to EF (depending on the dipole orientation). The magnitude The Authors, published by WILEY-VCH Verlag GmbH & Co. KGaA.
of that shift is, however, smaller than ED. It is intermediate
between the shifts in electrostatic energy in the region left of monolayer transport measurements on molecules or mono­
the dipoles (unaffected by ED) and right of the dipoles (fully layers sandwiched between two electrodes. Here, fundamental
shift by ED) with the exact magnitude of the shift depending on differences between molecular and monolayer junctions are
the localization of the state in the respective regions (see blue to be expected,[281] as collective electrostatic shifts occur only
states in the valence region indicated in Figure 16a).[280] for densely packed monolayers (due to the BD and TD layers),
To probe the positions of the occupied (unoccupied) frontier while they are essentially absent for isolated molecules.[282–284]
levels one can, for example, perform UV-photoelectron spec- The situation becomes even more complex for charge ­transport
troscopy (and inverse photoelectron spectroscopy) experiments. through molecular clusters, for which also significant edge
Alternatively, the interfacial level alignment can be deter- effects occur:[64,285] The transport levels of the molecules
mined by scanning tunneling spectroscopy experiments and in the centers of such clusters are shifted most by the fields

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (26 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

originating from the surrounding dipoles, while the effect example, serve as interfacial quantum-wells.[278] Similarly, the
diminishes toward the periphery.[285] Transport measurements inclusion of polar layers between 2D semiconductors in van der
have also been used to probe the role of ED in embedded- Waals stacks can pave the way for novel quantum-cascade struc-
dipole SAMs, where distinct shifts of the so-called transition tures,[286] which, for example, facilitate exciton dissociation (see
voltage have been observed depending on the orientation of simulated data in Figure 18). Employing electrostatic design in
the embedded dipole,[280] fully consistent with the considera- covalent organic or metal–organic frameworks with properly
tions in the previous paragraph. arranged dipolar linkers can also be used to create quantum-
checkerboards with electrostatic potential distributions local-
izing electrons and holes in spatially separated regions
7. Electrostatic Design consisting of identical semiconducting entities.[287] Interest-
ingly, such electrostatically tuned covalent networks have in the
The omnipresence of collective electrostatic effects at interfaces meantime also been realized in practice, albeit not in the bulk,
raises the question, whether one could intentionally integrate but on surfaces.[288]
dipolar layers into materials to locally manipulate their elec-
tronic structure, thereby creating materials with unprecedented
properties. Building on density-functional theory calculations, 8. Conclusions
this led us to suggest the concept of electrostatic design, which
could, for example, be used to develop an electronic LEGO For interfaces, it is safe to state that “dipoles are always and
approach with monolayers consisting of semiconducting, polar, everywhere.” As periodic assemblies of dipoles cause a step in
or even radical building blocks.[278] The simulated electronic the electrostatic energy, they profoundly impact the electronic
structure of such a layer consisting of a docking functionality properties of interfaces. In the present paper, we provide an
(thiolate), one electrostatic shifting functionality (a methyl- in-depth discussion of the underlying effects and their conse-
substituted bipyrimidine segment), and two semiconducting quences starting with a description of collective electrostatic
regions (two tolane sections) is shown in Figure 17. More com- effects caused by periodic dipole assemblies. We also review
plex structures following the same design strategy could, for how the properties of hypothetical, infinitely extended interfaces

Figure 18.  Schematic representation and DFT-calculated local density of states (LDOS) of a) a quantum cascade and c) a quantum well realized by
inserting polar titanyl-phthalocyanine layers between MoS2 sheets. The corresponding LDOS’s shown in (b) and (d) display the spatial and energetic
localizations of the states in the semiconducting sheets and in the polar layers. They are resolved with respect to the stacking direction. The Fermi
level is set to the middle of the bandgap. Reproduced under the terms of a Creative Commons Attribution 3.0 International License.[286] Copyright
2018, IOP Publishing Ltd.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (27 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

differ from realistic samples with a finite spatial extent con- [1] N. Koch, ChemPhysChem 2007, 8, 1438.
taining defects and structural imperfections. This is followed by [2] C. Liu, Y. Xu, Y.-Y. Noh, Mater. Today 2015, 18, 79.
a discussion of how dipolar layers impact the electronic proper- [3] D. Käfer, L. Ruppel, G. Witte, Phys. Rev. B 2007, 75, 085309.
ties of the isolated subsystems eventually constituting an inter- [4] S. Liu, W. M. Wang, A. L. Briseno, S. C. B. Mannsfeld, Z. Bao, Adv.
Mater. 2009, 21, 1217.
face (e.g., metal or semiconductor substrates and thin organic
[5] A. A. Virkar, S. Mannsfeld, Z. Bao, N. Stingelin, Adv. Mater. 2010,
layers) and how new dipole layers originating from the inter- 22, 3857.
face formation further modify the alignment of electronic states [6] A. Pick, G. Witte, Langmuir 2016, 32, 8019.
and the sample work function. The multiple possible origins [7] C. Park, J. E. Park, H. C. Choi, Acc. Chem. Res. 2014, 47, 2353.
of interfacial dipoles are reviewed, discussing the properties [8] A. O. F. Jones, B. Chattopadhyay, Y. H. Geerts, R. Resel, Adv. Funct.
of aligned polar functional groups contained in the adsorbate, Mater. 2016, 26, 2233.
metal surface dipoles due to the tailing of the electron cloud [9] L.-L. Chua, J. Zaumseil, J.-F. Chang, E. C.-W. Ou, P. K.-H. Ho,
out of the surface, band bending in doped inorganic semicon- H. Sirringhaus, R. H. Friend, Nature 2005, 434, 194.
ductors, charge rearrangements due to the formation of strong [10] D. K. Hwang, M. S. Oh, J. M. Hwang, J. H. Kim, S. Im, Appl. Phys.
chemical bonds between the substrate and the adsorbate layer, Lett. 2008, 92, 013304.
[11] S. K. Possanner, K. Zojer, P. Pacher, E. Zojer, F. Schürrer, Adv.
and charge transfer to/from monolayers consisting of electron
Funct. Mater. 2009, 19, 958.
accepting or electron donating molecules. In this context, espe- [12] D. Braga, G. Horowitz, Adv. Mater. 2009, 21, 1473.
cially interfacial charge transfer processes display particularly [13] F. Gholamrezaie, A.-M. Andringa, W. S. C. Roelofs, A. Neuhold,
rich physics, as one has to distinguish between situations in M. Kemerink, P. W. M. Blom, D. M. de Leeuw, Small 2012, 8, 241.
which each of the adsorbed molecules is fractionally charged [14] M. Aghamohammadi, R. Rödel, U. Zschieschang, C. Ocal,
and cases in which a fraction of the molecules carries an integer H. Boschker, R. T. Weitz, E. Barrena, H. Klauk, ACS Appl. Mater.
charge. Moreover, Fermi-level pinning due to interfacial charge Interfaces 2015, 7, 22775.
transfer renders the system work function independent of that [15] T. M. Clarke, J. R. Durrant, Chem. Rev. 2010, 110, 6736.
of the substrate and negates the impact of ordered dipole layers [16] D. M. Stoltzfus, J. E. Donaghey, A. Armin, P. E. Shaw, P. L. Burn,
spatially arranged between substrate and adsorbate. It can P. Meredith, Chem. Rev. 2016, 116, 12920.
[17] J.-L. Brédas, J. E. Norton, J. Cornil, V. Coropceanu, Acc. Chem. Res.
also give rise to massive changes in the electronic structure of
2009, 42, 1691.
the interface upon molecular reorientation processes, even if [18] K. Walzer, B. Maennig, M. Pfeiffer, K. Leo, Chem. Rev. 2007, 107,
the adsorbed molecules do not possess a net dipole moment. 1233.
Notably, interfacial dipole layers not only change electrode [19] D. F. O’Brien, M. A. Baldo, M. E. Thompson, S. R. Forrest, Appl.
work functions and, concomitantly, carrier-injection barriers Phys. Lett. 1999, 74, 442.
(which makes them interesting for applications in (opto)elec- [20] V. I. Adamovich, S. R. Cordero, P. I. Djurovich, A. Tamayo,
tronic devices). They also modify core-level binding energies M. E. Thompson, B. W. D’Andrade, S. R. Forrest, Org. Electron.
and transport properties of monolayer junctions. All the aspects 2003, 4, 77.
mentioned above raise the question, whether polar layers could [21] H. Ishii, K. Sugiyama, E. Ito, K. Seki, Adv. Mater. 1999, 11, 605.
not be used for manipulating the electronic properties of mate- [22] A. Kahn, N. Koch, W. Gao, J. Polym. Sci., Part B: Polym. Phys. 2003,
41, 2529.
rials in order to realize systems with unprecedented properties
[23] J. C. Scott, J. Vac. Sci. Technol., A 2003, 21, 521.
in the framework of an “electrostatic” design approach. [24] S. Braun, W. R. Salaneck, M. Fahlman, Adv. Mater. 2009, 21, 1450.
[25] J. Hwang, A. Wan, A. Kahn, Mater. Sci. Eng., R 2009, 64, 1.
[26] Y. Gao, Mater. Sci. Eng., R 2010, 68, 39.
Acknowledgements [27] R. Otero, A. L. Vázquez de Parga, J. M. Gallego, Surf. Sci. Rep.
2017, 72, 105.
The authors acknowledge financial support by the Austrian Science [28] G. Horowitz, J. Mater. Res. 2004, 19, 1946.
Fund (FWF): grants P28051-N36, I2081-N20, and P28631-N36. The new [29] H. Sirringhaus, Adv. Mater. 2005, 17, 2411.
computational results presented in the paper have been achieved using [30] D. Natali, M. Caironi, Adv. Mater. 2012, 24, 1357.
the Vienna Scientific Cluster (VSC). This article is part of the Advanced [31] B. Yang, Y. Yuan, P. Sharma, S. Poddar, R. Korlacki, S. Ducharme,
Materials Interfaces Hall of Fame article series, which highlights the work A. Gruverman, R. Saraf, J. Huang, Adv. Mater. 2012, 24, 1455.
of top interface and surface scientists.
[32] H. Kim, K.-G. Lim, T.-W. Lee, Energy Environ. Sci. 2016, 9, 12.
[33] K.-G. Lim, S. Ahn, T.-W. Lee, J. Mater. Chem. C 2018, 6, 2915.
[34] I. H. Campbell, S. Rubin, T. A. Zawodzinski, J. D. Kress,
Conflict of Interest R. L. Martin, D. L. Smith, N. N. Barashkov, J. P. Ferraris, Phys. Rev.
B 1996, 54, R14321.
The authors declare no conflict of interest. [35] I. H. Campbell, J. D. Kress, R. L. Martin, D. L. Smith,
N. N. Barashkov, J. P. Ferraris, Appl. Phys. Lett. 1997, 71, 3528.
[36] S. Casalini, C. A. Bortolotti, F. Leonardi, F. Biscarini, Chem. Soc.
Rev. 2017, 46, 40.
Keywords [37] L. Zuppiroli, L. Si-Ahmed, K. Kamaras, F. Nüesch, M. N. Bussac,
collective electrostatic effects, energy-level alignment, Fermi-level D. Ades, A. Siove, E. Moons, M. Grätzel, Eur. Phys. J. B 1999, 11,
pinning, interfacial charge transfer, organic/inorganic interfaces 505.
[38] C. Ganzorig, K.-J. Kwak, K. Yagi, M. Fujihira, Appl. Phys. Lett. 2001,
Received: March 31, 2019 79, 272.
Revised: April 26, 2019 [39] H. Yan, Q. Huang, J. Cui, J. G. C. Veinot, M. M. Kern, T. J. Marks,
Published online: June 26, 2019 Adv. Mater. 2003, 15, 835.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (28 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

[40] B. de Boer, A. Hadipour, M. M. Mandoc, T. van Woudenbergh, [71] J. Goniakowski, C. Noguera, L. Giordano, Phys. Rev. Lett. 2007, 98,
P. W. M. Blom, Adv. Mater. 2005, 17, 621. 205701.
[41] H.-L. Yip, S. K. Hau, N. S. Baek, H. Ma, A. K.-Y. Jen, Adv. Mater. [72] F. Rissner, A. Natan, D. A. Egger, O. T. Hofmann, L. Kronik,
2008, 20, 2376. E. Zojer, Org. Electron. 2012, 13, 3165.
[42] X. Cheng, Y.-Y. Noh, J. Wang, M. Tello, J. Frisch, R.-P. Blum, [73] J. Topping, Proc. R. Soc. A 1927, 114, 67.
A. Vollmer, J. P. Rabe, N. Koch, H. Sirringhaus, Adv. Funct. Mater. [74] J. R. Macdonald, C. A. Barlow, J. Chem. Phys. 1963, 39, 412.
2009, 19, 2407. [75] B. L. Maschhoff, J. P. Cowin, J. Chem. Phys. 1994, 101,
[43] M. Kitamura, Y. Kuzumoto, W. Kang, S. Aomori, Y. Arakawa, Appl. 8138.
Phys. Lett. 2010, 97, 033306. [76] V. De Renzi, R. Rousseau, D. Marchetto, R. Biagi, S. Scandolo,
[44] O. Fenwick, C. Van Dyck, K. Murugavel, D. Cornil, F. Reinders, U. del Pennino, Phys. Rev. Lett. 2005, 95, 046804.
S. Haar, M. Mayor, J. Cornil, P. Samorì, J. Mater. Chem. C 2015, 3, [77] O. Gershevitz, C. N. Sukenik, J. Ghabboun, D. Cahen,
3007. J. Am. Chem. Soc. 2003, 125, 4730.
[45] T. Mosciatti, P. Greco, T. Leydecker, M. Eredia, F. Biscarini, [78] H. Fukagawa, H. Yamane, S. Kera, K. K. Okudaira, N. Ueno, Phys.
P. Samorì, ACS Omega 2017, 2, 3502. Rev. B 2006, 73, 041302(R).
[46] J. W. Borchert, B. Peng, F. Letzkus, J. N. Burghartz, P. K. L. Chan, [79] H. Fukagawa, S. Hosoumi, H. Yamane, S. Kera, N. Ueno, Phys.
K. Zojer, S. Ludwigs, H. Klauk, Nat. Commun. 2019, 10, 1119. Rev. B 2011, 83, 085304.
[47] P. Schulz, T. Schäfer, C. D. Zangmeister, C. Effertz, D. Meyer, [80] M. L. Blumenfeld, M. P. Steele, O. L. A. Monti, J. Phys. Chem. Lett.
D. Mokros, R. D. van Zee, R. Mazzarello, M. Wuttig, Adv. Mater. 2010, 1, 145.
Interfaces 2014, 1, 1300130. [81] D. Cornil, Y. Olivier, V. Geskin, J. Cornil, Adv. Funct. Mater. 2007,
[48] C. Bock, D. V. Pham, U. Kunze, D. Käfer, G. Witte, C. Wöll, J. Appl. 17, 1143.
Phys. 2006, 100, 114517. [82] M. L. Sushko, A. L. Shluger, Adv. Funct. Mater. 2008, 18, 2228.
[49] J. Kim, Y. S. Rim, Y. Liu, A. C. Serino, J. C. Thomas, H. Chen, [83] L. Romaner, G. Heimel, C. Ambrosch-Draxl, E. Zojer, Adv. Funct.
Y. Yang, P. S. Weiss, Nano Lett. 2014, 14, 2946. Mater. 2008, 18, 3999.
[50] T. Abu-Husein, S. Schuster, D. A. Egger, M. Kind, T. Santowski, [84] M. Piacenza, S. D’Agostino, E. Fabiano, F. Della Sala, Phys. Rev. B
A. Wiesner, R. Chiechi, E. Zojer, A. Terfort, M. Zharnikov, Adv. 2009, 80, 153101.
Funct. Mater. 2015, 25, 3943. [85] O. T. Hofmann, J.-C. Deinert, Y. Xu, P. Rinke, J. Stähler, M. Wolf,
[51] A. Petritz, M. Krammer, E. Sauter, M. Gärtner, G. Nascimbeni, M. Scheffler, J. Chem. Phys. 2013, 139, 174701.
B. Schrode, A. Fian, H. Gold, A. Cojocaru, E. Karner-Petritz, [86] B. J. Topham, M. Kumar, Z. G. Soos, Adv. Funct. Mater. 2011, 21,
R. Resel, A. Terfort, E. Zojer, M. Zharnikov, K. Zojer, B. Stadlober, 1931.
Adv. Funct. Mater. 2018, 28, 1804462. [87] A. Natan, N. Kuritz, L. Kronik, Adv. Funct. Mater. 2010, 20, 2077.
[52] K. P. Pernstich, S. Haas, D. Oberhoff, C. Goldmann, [88] S. J. Ausserlechner, M. Gruber, R. Hetzel, H.-G. Flesch,
D. J. Gundlach, B. Batlogg, A. N. Rashid, G. Schitter, J. Appl. Phys. L. Ladinig, L. Hauser, A. Haase, M. Buchner, R. Resel, F. Schürrer,
2004, 96, 6431. B. Stadlober, G. Trimmel, K. Zojer, E. Zojer, Phys. Status Solidi A
[53] S. Kobayashi, T. Nishikawa, T. Takenobu, S. Mori, T. Shimoda, 2012, 209, 181.
T. Mitani, H. Shimotani, N. Yoshimoto, S. Ogawa, Y. Iwasa, Nat. [89] N. K. Adam, J. F. Danielli, J. B. Harding, Proc. R. Soc. London, Ser.
Mater. 2004, 3, 317. A 1934, 147, 491.
[54] C. Huang, H. E. Katz, J. E. West, Langmuir 2007, 23, 13223. [90] G. Heimel, L. Romaner, E. Zojer, J.-L. Brédas, Nano Lett. 2007, 7,
[55] A. Janotti, C. G. Van de Walle, Phys. Rev. B 2007, 76, 165202. 932.
[56] Y. Jang, J. H. Cho, D. H. Kim, Y. D. Park, M. Hwang, K. Cho, Appl. [91] S. M. Sze, K. K. Ng, Physics of Semiconductor Devices, John Wiley &
Phys. Lett. 2007, 90, 132104. Sons, Hoboken, NJ, USA 2007.
[57] P. Pacher, A. Lex, V. Proschek, H. Etschmaier, E. Tchernychova, [92] H. Edlbauer, E. Zojer, O. T. Hofmann, J. Phys. Chem. C 2015, 119,
M. Sezen, U. Scherf, W. Grogger, G. Trimmel, C. Slugovc, E. Zojer, 27162.
Adv. Mater. 2008, 20, 3143. [93] E. Verwüster, E. Wruss, E. Zojer, O. T. Hofmann, J. Chem. Phys.
[58] J. M. Shannon, Solid-State Electron. 1977, 20, 869. 2017, 147, 024706.
[59] P. S. Davids, I. H. Campbell, D. L. Smith, J. Appl. Phys. 1997, 82, [94] I. Fernandez-Torrente, S. Monturet, K. J. Franke, J. Fraxedas,
6319. N. Lorente, J. I. Pascual, Phys. Rev. Lett. 2007, 99, 176103.
[60] N. Tessler, Y. Roichman, Appl. Phys. Lett. 2001, 79, 2987. [95] C. Wagner, D. Kasemann, C. Golnik, R. Forker, M. Esslinger,
[61] Rhoderick, F. H. , Metal-Semiconductor Contacts, Clarendon Press, K. Müllen, T. Fritz, Phys. Rev. B 2010, 81, 035423.
Oxford, UK 1978. [96] F. Krok, K. Sajewicz, J. Konior, M. Goryl, P. Piatkowski,
[62] J. J. Brondijk, F. Torricelli, E. C. P. Smits, P. W. M. Blom, M. Szymonski, Phys. Rev. B 2008, 77, 235427.
D. M. de Leeuw, Org. Electron. 2012, 13, 1526. [97] F. Mohn, L. Gross, N. Moll, G. Meyer, Nat. Nanotechnol. 2012, 7,
[63] M. Gruber, E. Zojer, F. Schürrer, K. Zojer, Adv. Funct. Mater. 2013, 227.
23, 2941. [98] L. Kou, Z. Ma, Y. J. Li, Y. Naitoh, M. Komiyama, Y. Sugawara,
[64] A. Natan, L. Kronik, H. Haick, R. T. Tung, Adv. Mater. 2007, 19, Nanotechnology 2015, 26, 195701.
4103. [99] S. Kawai, T. Glatzel, H.-J. Hug, E. Meyer, Nanotechnology 2010, 21,
[65] G. Heimel, L. Romaner, E. Zojer, J.-L. Bredas, Acc. Chem. Res. 245704.
2008, 41, 721. [100] K. Wandelt, Appl. Surf. Sci. 1997, 111, 1.
[66] D. Cahen, R. Naaman, Z. Vager, Adv. Funct. Mater. 2005, 15, [101] I. Hehn, S. Schuster, T. Wächter, T. Abu-Husein, A. Terfort,
1571. M. Zharnikov, E. Zojer, J. Phys. Chem. Lett. 2016, 7, 2994.
[67] A. Natan, Y. Zidon, Y. Shapira, L. Kronik, Phys. Rev. B 2006, 73, [102] D. Cahen, A. Kahn, Adv. Mater. 2003, 15, 271.
193310. [103] G. Koller, B. Winter, M. Oehzelt, J. Ivanco, F. P. Netzer,
[68] G. Heimel, F. Rissner, E. Zojer, Adv. Mater. 2010, 22, 2494. M. G. Ramsey, Org. Electron. 2007, 8, 63.
[69] O. L. A. Monti, J. Phys. Chem. Lett. 2012, 3, 2342. [104] H. Yamane, H. Honda, H. Fukagawa, M. Ohyama, Y. Hinuma,
[70] G. Heimel, L. Romaner, J.-L. Brédas, E. Zojer, Phys. Rev. Lett. 2006, S. Kera, K. K. Okudaira, N. Ueno, J. Electron Spectrosc. Relat.
96, 196806. Phenom. 2004, 137–140, 223.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (29 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

[105] T. Schultz, T. Lenz, N. Kotadiya, G. Heimel, G. Glasser, R. Berger, [138] H. Vázquez, R. Oszwaldowski, P. Pou, J. Ortega, R. Pérez,
P. W. M. Blom, P. Amsalem, D. M. de Leeuw, N. Koch, Adv. Mater. F. Flores, A. Kahn, Europhys. Lett. 2004, 65, 802.
Interfaces 2017, 4, 1700324. [139] H. Vázquez, Y. J. Dappe, J. Ortega, F. Flores, J. Chem. Phys. 2007,
[106] T. He, J. He, M. Lu, B. Chen, H. Pang, W. F. Reus, W. M. Nolte, 126, 144703.
D. P. Nackashi, P. D. Franzon, J. M. Tour, J. Am. Chem. Soc. 2006, [140] H. Vázquez, F. Flores, A. Kahn, Org. Electron. 2007, 8, 241.
128, 14537. [141] J. D. Sau, J. B. Neaton, H. J. Choi, S. G. Louie, M. L. Cohen, Phys.
[107] A. Vilan, D. Cahen, Chem. Rev. 2017, 117, 4624. Rev. Lett. 2008, 101, 026804.
[108] R. W. Strayer, W. Mackie, L. W. Swanson, Surf. Sci. 1973, 34, 225. [142] A. Biller, I. Tamblyn, J. B. Neaton, L. Kronik, J. Chem. Phys. 2011,
[109] N. D. Lang, W. Kohn, Phys. Rev. B 1973, 7, 3541. 135, 164706.
[110] C. Woll, Prog. Surf. Sci. 2007, 82, 55. [143] C. D. Spataru, Phys. Rev. B 2013, 88, 045101.
[111] E. Fatuzzo, G. Harbeke, W. J. Merz, R. Nitsche, H. Roetschi, [144] J. B. Neaton, M. S. Hybertsen, S. G. Louie, Phys. Rev. Lett. 2006,
W. Ruppel, Phys. Rev. 1962, 127, 2036. 97, 216405.
[112] A. Konishi, T. Ogawa, C. A. J. Fisher, A. Kuwabara, T. Shimizu, [145] Y. Li, D. Lu, G. Galli, J. Chem. Theory Comput. 2009, 5, 881.
S. Yasui, M. Itoh, H. Moriwake, Appl. Phys. Lett. 2016, 109, 102903. [146] C. Freysoldt, P. Rinke, M. Scheffler, Phys. Rev. Lett. 2009, 103,
[113] R. Joshi, P. Kumar, A. Gaur, K. Asokan, Appl. Nanosci. 2014, 4, 531. 056803.
[114] B. Bieniek, O. T. Hofmann, P. Rinke, Appl. Phys. Lett. 2015, 106, [147] M. Dell’Angela, G. Kladnik, A. Cossaro, A. Verdini, M. Kamenetska,
131602. I. Tamblyn, S. Y. Quek, J. B. Neaton, D. Cvetko, A. Morgante,
[115] J. Stähler, P. Rinke, Chem. Phys. 2017, 485–486, 149. L. Venkataraman, Nano Lett. 2010, 10, 2470.
[116] C. G. Van de Walle, Phys. Rev. Lett. 2000, 85, 1012. [148] D. A. Egger, Z.-F. Liu, J. B. Neaton, L. Kronik, Nano Lett. 2015, 15,
[117] F. Oba, M. Choi, A. Togo, I. Tanaka, Sci. Technol. Adv. Mater. 2011, 2448.
12, 034302. [149] D. Otálvaro, T. Veening, G. Brocks, J. Phys. Chem. C 2012, 116,
[118] C. G. Van de Walle, J. Neugebauer, J. Appl. Phys. 2004, 95, 7826.
3851. [150] M. N. Faraggi, N. Jiang, N. Gonzalez-Lakunza, A. Langner,
[119] P. D. C. King, T. D. Veal, J. Phys.: Condens. Matter 2011, 23, 334214. S. Stepanow, K. Kern, A. Arnau, J. Phys. Chem. C 2012, 116, 24558.
[120] D.-J. Lee, H.-M. Kim, J.-Y. Kwon, H. Choi, S.-H. Kim, K.-B. Kim, [151] D. Wegner, R. Yamachika, Y. Wang, V. W. Brar, B. M. Bartlett,
Adv. Funct. Mater. 2011, 21, 448. J. R. Long, M. F. Crommie, Nano Lett. 2008, 8, 131.
[121] T. Schultz, J. Niederhausen, R. Schlesinger, S. Sadofev, N. Koch, J. [152] P. J. Blowey, S. Velari, L. A. Rochford, D. A. Duncan, D. A. Warr,
Appl. Phys. 2018, 123, 245501. T.-L. Lee, A. De Vita, G. Costantini, D. P. Woodruff, Nanoscale
[122] H. L. Mosbacker, Y. M. Strzhemechny, B. D. White, P. E. Smith, 2018, 10, 14984.
D. C. Look, D. C. Reynolds, C. W. Litton, L. J. Brillson, Appl. Phys. [153] O. T. Hofmann, D. A. Egger, E. Zojer, Nano Lett. 2010, 10, 4369.
Lett. 2005, 87, 012102. [154] I. Lange, S. Reiter, M. Pätzel, A. Zykov, A. Nefedov, J. Hildebrandt,
[123] R. Schlesinger, Y. Xu, O. T. Hofmann, S. Winkler, J. Frisch, S. Hecht, S. Kowarik, C. Wöll, G. Heimel, D. Neher, Adv. Funct.
J. Niederhausen, A. Vollmer, S. Blumstengel, F. Henneberger, Mater. 2014, 24, 7014.
P. Rinke, M. Scheffler, N. Koch, Phys. Rev. B 2013, 87, 155311. [155] A. R. McNeill, A. R. Hyndman, R. J. Reeves, A. J. Downard,
[124] M. Knupfer, Appl. Phys. A 2003, 77, 623. M. W. Allen, ACS Appl. Mater. Interfaces 2016, 8, 31392.
[125] S. Sharifzadeh, A. Biller, L. Kronik, J. B. Neaton, Phys. Rev. B 2012, [156] C. K. Chan, E.-G. Kim, J.-L. Brédas, A. Kahn, Adv. Funct. Mater.
85, 125307. 2006, 16, 831.
[126] S. Refaely-Abramson, S. Sharifzadeh, M. Jain, R. Baer, J. B. Neaton, [157] T. Nishi, K. Kanai, Y. Ouchi, M. R. Willis, K. Seki, Chem. Phys. Lett.
L. Kronik, Phys. Rev. B 2013, 88, 081204(R). 2005, 414, 479.
[127] N. Dori, M. Menon, L. Kilian, M. Sokolowski, L. Kronik, [158] Q. Bao, S. Braun, C. Wang, X. Liu, M. Fahlman, Adv. Mater. Inter-
E. Umbach, Phys. Rev. B 2006, 73, 195208. faces 2019, 6, 1800897.
[128] S. Refaely-Abramson, S. Sharifzadeh, N. Govind, J. Autschbach, [159] H. Ishii, N. Hayashi, E. Ito, Y. Washizu, K. Sugi, Y. Kimura,
J. B. Neaton, R. Baer, L. Kronik, Phys. Rev. Lett. 2012, 109, 226405. M. Niwano, Y. Ouchi, K. Seki, Phys. Status Solidi A 2004, 201,
[129] N. Sato, K. Seki, H. Inokuchi, J. Chem. Soc., Faraday Trans. 2 1981, 1075.
77, 1621. [160] J. Hwang, E.-G. Kim, J. Liu, J.-L. Brédas, A. Duggal, A. Kahn,
[130] S. Duhm, G. Heimel, I. Salzmann, H. Glowatzki, R. L. Johnson, J. Phys. Chem. C 2007, 111, 1378.
A. Vollmer, J. P. Rabe, N. Koch, Nat. Mater. 2008, 7, 326. [161] J. C. Blakesley, N. C. Greenham, J. Appl. Phys. 2009, 106, 034507.
[131] G. Heimel, I. Salzmann, S. Duhm, J. P. Rabe, N. Koch, Adv. Funct. [162] I. Lange, J. C. Blakesley, J. Frisch, A. Vollmer, N. Koch, D. Neher,
Mater. 2009, 19, 3874. Phys. Rev. Lett. 2011, 106, 216402.
[132] I. Salzmann, S. Duhm, G. Heimel, M. Oehzelt, R. Kniprath, [163] J. K. Wenderott, B. X. Dong, P. F. Green, J. Mater. Chem. C 2017,
R. L. Johnson, J. P. Rabe, N. Koch, J. Am. Chem. Soc. 2008, 130, 5, 7446.
12870. [164] Q. Bao, S. Fabiano, M. Andersson, S. Braun, Z. Sun, X. Crispin,
[133] B. H. Hamadani, D. A. Corley, J. W. Ciszek, J. M. Tour, D. Natelson, M. Berggren, X. Liu, M. Fahlman, Adv. Funct. Mater. 2016, 26,
Nano Lett. 2006, 6, 1303. 1077.
[134] A. W. Hains, J. Liu, A. B. F. Martinson, M. D. Irwin, T. J. Marks, [165] M. Schneider, A. Wagenpfahl, C. Deibel, V. Dyakonov, A. Schöll,
Adv. Funct. Mater. 2010, 20, 595. F. Reinert, Org. Electron. 2014, 15, 1552.
[135] N. Crivillers, S. Osella, C. Van Dyck, G. M. Lazzerini, D. Cornil, [166] H. Wang, P. Amsalem, G. Heimel, I. Salzmann, N. Koch,
A. Liscio, F. Di Stasio, S. Mian, O. Fenwick, F. Reinders, M. Oehzelt, Adv. Mater. 2014, 26, 925.
M. Neuburger, E. Treossi, M. Mayor, V. Palermo, F. Cacialli, [167] G. Paasch, H. Peisert, M. Knupfer, J. Fink, S. Scheinert, J. Appl.
J. Cornil, P. Samorì, Adv. Mater. 2013, 25, 432. Phys. 2003, 93, 6084.
[136] O. M. Cabarcos, S. Schuster, I. Hehn, P. P. Zhang, M. M. Maitani, [168] T. Sueyoshi, H. Fukagawa, M. Ono, S. Kera, N. Ueno, Appl. Phys.
N. Sullivan, J.-B. Giguère, J.-F. Morin, P. S. Weiss, E. Zojer, Lett. 2009, 95, 183303.
M. Zharnikov, D. L. Allara, J. Phys. Chem. C 2017, 121, 15815. [169] W. L. Kalb, B. Batlogg, Phys. Rev. B 2010, 81, 155315.
[137] M. Oehzelt, K. Akaike, N. Koch, G. Heimel, Sci. Adv. 2015, 1, [170] F. Bussolotti, S. Kera, K. Kudo, A. Kahn, N. Ueno, Phys. Rev. Lett.
e1501127. 2013, 110, 267602.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (30 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

[171] S. Yogev, R. Matsubara, M. Nakamura, U. Zschieschang, H. Klauk, [206] M. L. Blumenfeld, M. P. Steele, N. Ilyas, O. L. A. Monti, Surf. Sci.
Y. Rosenwaks, Phys. Rev. Lett. 2013, 110, 036803. 2010, 604, 1649.
[172] M. Oehzelt, N. Koch, G. Heimel, Nat. Commun. 2014, 5, 4174. [207] Y. L. Huang, W. Chen, F. Bussolotti, T. C. Niu, A. T. S. Wee,
[173] L.-H. Zhao, R.-Q. Png, C. C. H. Chiam, H. Guo, J.-M. Zhuo, N. Ueno, S. Kera, Phys. Rev. B 2013, 87, 085205.
L.-L. Chua, A. T. S. Wee, P. K. H. Ho, Appl. Phys. Lett. 2012, 101, [208] E. Wruss, O. T. Hofmann, D. A. Egger, E. Verwüster, A. Gerlach,
053304. F. Schreiber, E. Zojer, J. Phys. Chem. C 2016, 120, 6869.
[174] W. Zhao, E. Salomon, Q. Zhang, S. Barlow, S. R. Marder, A. Kahn, [209] S. S. Harivyasi, O. T. Hofmann, N. Ilyas, O. L. A. Monti, E. Zojer,
Phys. Rev. B 2008, 77, 165336. J. Phys. Chem. C 2018, 122, 14621.
[175] J. Q. Zhong, H. Y. Mao, R. Wang, D. C. Qi, L. Cao, Y. Z. Wang, [210] Y. Zhou, C. Fuentes-Hernandez, J. Shim, J. Meyer, A. J. Giordano,
W. Chen, J. Phys. Chem. C 2011, 115, 23922. H. Li, P. Winget, T. Papadopoulos, H. Cheun, J. Kim, M. Fenoll,
[176] Y. Gao, H. Ding, H. Wang, D. Yan, Appl. Phys. Lett. 2007, 91, A. Dindar, W. Haske, E. Najafabadi, T. M. Khan, H. Sojoudi,
142112. S. Barlow, S. Graham, J.-L. Bredas, S. R. Marder, A. Kahn,
[177] S. Braun, M. P. de Jong, W. Osikowicz, W. R. Salaneck, Appl. Phys. B. Kippelen, Science 2012, 336, 327.
Lett. 2007, 91, 202108. [211] V.-E. Choong, M. G. Mason, C. W. Tang, Y. Gao, Appl. Phys. Lett.
[178] I. Salzmann, G. Heimel, M. Oehzelt, S. Winkler, N. Koch, Acc. 1998, 72, 2689.
Chem. Res. 2016, 49, 370. [212] C. Urban, Y. Wang, J. Rodríguez-Fernández, R. García,
[179] S. Olthof, R. Meerheim, M. Schober, K. Leo, Phys. Rev. B 2009, 79, M. Á. Herranz, M. Alcamí, N. Martín, F. Martín, J. M. Gallego,
245308. R. Miranda, R. Otero, Chem. Commun. 2014, 50, 833.
[180] S. Olthof, W. Tress, R. Meerheim, B. Lüssem, K. Leo, J. Appl. Phys. [213] N. Ilyas, S. S. Harivyasi, P. Zahl, R. Cortes, O. T. Hofmann,
2009, 106, 103711. P. Sutter, E. Zojer, O. L. A. Monti, J. Phys. Chem. C 2016, 120,
[181] Y. Shinmura, T. Yoshioka, T. Kaji, M. Hiramoto, Appl. Phys. Express 7113.
2014, 7, 071601. [214] P. S. Bagus, V. Staemmler, C. Wöll, Phys. Rev. Lett. 2002, 89,
[182] M. Hiramoto, M. Kikuchi, S. Izawa, Adv. Mater. 2019, 31, 1801236. 096104.
[183] K. Akaike, Jpn. J. Appl. Phys. 2018, 57, 03EA03. [215] H. Mizushima, H. Koike, K. Kuroda, Y. Ishida, M. Nakayama,
[184] M. L. Tietze, J. Benduhn, P. Pahner, B. Nell, M. Schwarze, K. Mase, T. Kondo, S. Shin, K. Kanai, Phys. Chem. Chem. Phys.
H. Kleemann, M. Krammer, K. Zojer, K. Vandewal, K. Leo, Nat. 2017, 19, 18646.
Commun. 2018, 9, 1182. [216] H. Öström, L. Triguero, M. Nyberg, H. Ogasawara,
[185] I. E. Jacobs, A. J. Moulé, Adv. Mater. 2017, 29, 1703063. L. G. M. Pettersson, A. Nilsson, Phys. Rev. Lett. 2003, 91, 046102.
[186] S. Lacher, Y. Matsuo, E. Nakamura, J. Am. Chem. Soc. 2011, 133, [217] Y. Morikawa, H. Ishii, K. Seki, Phys. Rev. B 2004, 69, 041403(R).
16997. [218] S. Rentenberger, A. Vollmer, E. Zojer, R. Schennach, N. Koch,
[187] M. Gärtner, E. Sauter, G. Nascimbeni, A. Petritz, A. Wiesner, J. Appl. Phys. 2006, 100, 053701.
M. Kind, T. Abu-Husein, M. Bolte, B. Stadlober, E. Zojer, A. Terfort, [219] C. D. Bain, E. B. Troughton, Y. T. Tao, J. Evall, G. M. Whitesides,
M. Zharnikov, J. Phys. Chem. C 2018, 122, 28757. R. G. Nuzzo, J. Am. Chem. Soc. 1989, 111, 321.
[188] A. Ulman, Chem. Rev. 1996, 96, 1533. [220] L. Romaner, G. Heimel, J.-L. Brédas, A. Gerlach, F. Schreiber,
[189] F. Schreiber, Prog. Surf. Sci. 2000, 65, 151. R. L. Johnson, J. Zegenhagen, S. Duhm, N. Koch, E. Zojer, Phys.
[190] K. Heister, M. Zharnikov, M. Grunze, L. S. O. Johansson, J. Phys. Rev. Lett. 2007, 99, 256801.
Chem. B 2001, 105, 4058. [221] G. M. Rangger, O. T. Hofmann, L. Romaner, G. Heimel, B. Bröker,
[191] W. Azzam, C. Fuxen, A. Birkner, H.-T. Rong, M. Buck, C. Wöll, R.-P. Blum, R. L. Johnson, N. Koch, E. Zojer, Phys. Rev. B 2009, 79,
Langmuir 2003, 19, 4958. 165306.
[192] F. Schreiber, J. Phys.: Condens. Matter 2004, 16, R881. [222] T.-C. Tseng, C. Urban, Y. Wang, R. Otero, S. L. Tait, M. Alcamí,
[193] J. C. Love, L. A. Estroff, J. K. Kriebel, R. G. Nuzzo, D. Écija, M. Trelka, J. M. Gallego, N. Lin, M. Konuma, U. Starke,
G. M. Whitesides, Chem. Rev. 2005, 105, 1103. A. Nefedov, A. Langner, C. Wöll, M. Á. Herranz, F. Martín,
[194] C. Vericat, M. E. Vela, G. Benitez, P. Carro, R. C. Salvarezza, Chem. N. Martín, K. Kern, R. Miranda, Nat. Chem. 2010, 2, 374.
Soc. Rev. 2010, 39, 1805. [223] P. Borghetti, A. Sarasola, N. Merino-Díez, G. Vasseur, L. Floreano,
[195] D. Käfer, A. Bashir, G. Witte, J. Phys. Chem. C 2007, 111, 10546. J. Lobo-Checa, A. Arnau, D. G. de Oteyza, J. E. Ortega, J. Phys.
[196] A. Shaporenko, P. Cyganik, M. Buck, A. Terfort, M. Zharnikov, Chem. C 2017, 121, 28412.
J. Phys. Chem. B 2005, 109, 13630. [224] A. Kumar, K. Banerjee, M. Dvorak, F. Schulz, A. Harju, P. Rinke,
[197] A. Bashir, D. Käfer, J. Müller, C. Wöll, A. Terfort, G. Witte, Angew. P. Liljeroth, ACS Nano 2017, 11, 4960.
Chem., Int. Ed. 2008, 47, 5250. [225] G. Heimel, S. Duhm, I. Salzmann, A. Gerlach, A. Strozecka,
[198] J. Ossowski, G. Nascimbeni, T. Żaba, E. Verwüster, J. Rysz, J. Niederhausen, C. Bürker, T. Hosokai, I. Fernandez-Torrente,
A. Terfort, M. Zharnikov, E. Zojer, P. Cyganik, J. Phys. Chem. C G. Schulze, S. Winkler, A. Wilke, R. Schlesinger, J. Frisch, B. Bröker,
2017, 121, 28031. A. Vollmer, B. Detlefs, J. Pflaum, S. Kera, K. J. Franke, N. Ueno,
[199] C. Queffélec, M. Petit, P. Janvier, D. A. Knight, B. Bujoli, Chem. J. I. Pascual, F. Schreiber, N. Koch, Nat. Chem. 2013, 5, 187.
Rev. 2012, 112, 3777. [226] F. S. Tautz, Prog. Surf. Sci. 2007, 82, 479.
[200] M. Timpel, M. V. Nardi, S. Krause, G. Ligorio, C. Christodoulou, [227] S. Duhm, A. Gerlach, I. Salzmann, B. Bröker, R. L. Johnson,
L. Pasquali, A. Giglia, J. Frisch, B. Wegner, P. Moras, N. Koch, F. Schreiber, N. Koch, Org. Electron. 2008, 9, 111.
Chem. Mater. 2014, 26, 5042. [228] L. Romaner, D. Nabok, P. Puschnig, E. Zojer, C. Ambrosch-Draxl,
[201] S. P. Pujari, L. Scheres, A. T. M. Marcelis, H. Zuilhof, Angew. New J. Phys. 2009, 11, 053010.
Chem., Int. Ed. 2014, 53, 6322. [229] P. C. Rusu, G. Giovannetti, C. Weijtens, R. Coehoorn, G. Brocks,
[202] S. A. DiBenedetto, A. Facchetti, M. A. Ratner, T. J. Marks, Adv. Phys. Rev. B 2010, 81, 121101(R).
Mater. 2009, 21, 1407. [230] F. Jensen, Introduction to Computational Chemistry, John Wiley &
[203] H. Ma, H.-L. Yip, F. Huang, A. K.-Y. Jen, Adv. Funct. Mater. 2010, Sons, Hoboken, NJ, USA 1999.
20, 1371. [231] R. J. Murdey, W. R. Salaneck, Proc. SPIE 2004, 5519, 125.
[204] H. J. Lee, A. C. Jamison, T. R. Lee, Acc. Chem. Res. 2015, 48, 3007. [232] O. T. Hofmann, G. M. Rangger, E. Zojer, J. Phys. Chem. C 2008,
[205] D. E. Barlow, K. W. Hipps, J. Phys. Chem. B 2000, 104, 5993. 112, 20357.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (31 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

[233] J. I. Martínez, E. Abad, C. González, J. Ortega, F. Flores, Org. Elec- [260] W. Steurer, S. Fatayer, L. Gross, G. Meyer, Nat. Commun. 2015, 6,
tron. 2012, 13, 399. 8353.
[234] B. Bröker, R.-P. Blum, J. Frisch, A. Vollmer, O. T. Hofmann, [261] M. Hollerer, Doctoral Thesis, University of Graz, 2019.
R. Rieger, K. Müllen, J. P. Rabe, E. Zojer, N. Koch, Appl. Phys. Lett. [262] E. Wruss, E. Zojer, O. T. Hofmann, J. Phys. Chem. C 2018, 122,
2008, 93, 243303. 14640.
[235] A. Hauschild, K. Karki, B. C. C. Cowie, M. Rohlfing, F. S. Tautz, [263] M. Fahlman, A. Crispin, X. Crispin, S. K. M. Henze, M. P. de Jong,
M. Sokolowski, Phys. Rev. Lett. 2005, 94, 036106. W. Osikowicz, C. Tengstedt, W. R. Salaneck, J. Phys.: Condens.
[236] N. Koch, S. Duhm, J. P. Rabe, A. Vollmer, R. L. Johnson, Phys. Rev. Matter 2007, 19, 183202.
Lett. 2005, 95, 237601. [264] H. Vázquez, W. Gao, F. Flores, A. Kahn, Phys. Rev. B 2005, 71,
[237] B. Stadtmüller, T. Sueyoshi, G. Kichin, I. Kröger, S. Soubatch, 041306(R).
R. Temirov, F. S. Tautz, C. Kumpf, Phys. Rev. Lett. 2012, 108, [265] M. T. Greiner, M. G. Helander, W.-M. Tang, Z.-B. Wang, J. Qiu,
106103. Z.-H. Lu, Nat. Mater. 2012, 11, 76.
[238] B. Bröker, O. T. Hofmann, G. M. Rangger, P. Frank, R.-P. Blum, [266] M. Samadi Khoshkhoo, H. Peisert, T. Chassé, M. Scheele, Org.
R. Rieger, L. Venema, A. Vollmer, K. Müllen, J. P. Rabe, Electron. 2017, 49, 249.
A. Winkler, P. Rudolf, E. Zojer, N. Koch, Phys. Rev. Lett. 2010, 104, [267] Q. Bao, O. Sandberg, D. Dagnelund, S. Sandén, S. Braun,
246805. H. Aarnio, X. Liu, W. M. Chen, R. Österbacka, M. Fahlman, Adv.
[239] H. Glowatzki, B. Bröker, R.-P. Blum, O. T. Hofmann, A. Vollmer, Funct. Mater. 2014, 24, 6309.
R. Rieger, K. Müllen, E. Zojer, J. P. Rabe, N. Koch, Nano Lett. 2008, [268] P. C. Rusu, G. Giovannetti, C. Weijtens, R. Coehoorn, G. Brocks,
8, 3825. J. Phys. Chem. C 2009, 113, 9974.
[240] C. Tengstedt, W. Osikowicz, W. R. Salaneck, I. D. Parker, C.-H. Hsu, [269] J.-P. Yang, F. Bussolotti, S. Kera, N. Ueno, J. Phys. D: Appl. Phys.
M. Fahlman, Appl. Phys. Lett. 2006, 88, 053502. 2017, 50, 423002.
[241] S. R. Marder, B. Kippelen, A. K.-Y. Jen, N. Peyghambarian, Nature [270] A. Gerlach, T. Hosokai, S. Duhm, S. Kera, O. T. Hofmann, E. Zojer,
1997, 388, 845. J. Zegenhagen, F. Schreiber, Phys. Rev. Lett. 2011, 106, 156102.
[242] T. Hughbanks, R. Hoffmann, J. Am. Chem. Soc. 1983, 105, 1150. [271] M.-T. Chen, O. T. Hofmann, A. Gerlach, B. Bröker, C. Bürker,
[243] R. Hoffmann, Rev. Mod. Phys. 1988, 60, 601. J. Niederhausen, T. Hosokai, J. Zegenhagen, A. Vollmer, R. Rieger,
[244] B. Stadtmüller, M. Gruenewald, J. Peuker, R. Forker, T. Fritz, K. Müllen, F. Schreiber, I. Salzmann, N. Koch, E. Zojer, S. Duhm,
C. Kumpf, J. Phys. Chem. C 2014, 118, 28592. J. Phys.: Condens. Matter 2019, 31, 194002.
[245] M. Gruenewald, C. Sauer, J. Peuker, M. Meissner, F. Sojka, [272] O. T. Hofmann, H. Glowatzki, C. Bürker, G. M. Rangger, B. Bröker,
A. Schöll, F. Reinert, R. Forker, T. Fritz, Phys. Rev. B 2015, 91, J. Niederhausen, T. Hosokai, I. Salzmann, R.-P. Blum, R. Rieger,
144202. A. Vollmer, P. Rajput, A. Gerlach, K. Müllen, F. Schreiber, E. Zojer,
[246] R. Otero, D. Écija, G. Fernández, J. M. Gallego, L. Sánchez, N. Koch, S. Duhm, J. Phys. Chem. C 2017, 121, 24657.
N. Martín, R. Miranda, Nano Lett. 2007, 7, 2602. [273] Y. Xu, O. T. Hofmann, R. Schlesinger, S. Winkler, J. Frisch,
[247] B. Stadtmüller, D. Lüftner, M. Willenbockel, E. M. Reinisch, J. Niederhausen, A. Vollmer, S. Blumstengel, F. Henneberger,
T. Sueyoshi, G. Koller, S. Soubatch, M. G. Ramsey, P. Puschnig, N. Koch, P. Rinke, M. Scheffler, Phys. Rev. Lett. 2013, 111, 226802.
F. S. Tautz, C. Kumpf, Nat. Commun. 2014, 5, 3685. [274] E. Wruss, L. Hörmann, O. T. Hofmann, J. Phys. Chem. C 2019, 123,
[248] N. Sai, P. F. Barbara, K. Leung, Phys. Rev. Lett. 2011, 106, 7118.
220401. [275] Ratner, B. D., Castner, D. G., Surface Analysis: The Principal Tech-
[249] S. Fatayer, B. Schuler, W. Steurer, I. Scivetti, J. Repp, L. Gross, niques, Wiley-VCH, New York, USA 1997.
M. Persson, G. Meyer, Nat. Nanotechnol. 2018, 13, 376. [276] T. C. Taucher, I. Hehn, O. T. Hofmann, M. Zharnikov, E. Zojer,
[250] O. T. Hofmann, P. Rinke, M. Scheffler, G. Heimel, ACS Nano 2015, J. Phys. Chem. C 2016, 120, 3428.
9, 5391. [277] O. M. Cabarcos, A. Shaporenko, T. Weidner, S. Uppili, L. S. Dake,
[251] S. Erker, O. T. Hofmann, J. Phys. Chem. Lett. 2019, 10, 848. M. Zharnikov, D. L. Allara, J. Phys. Chem. C 2008, 112, 10842.
[252] A. Crispin, X. Crispin, M. Fahlman, M. Berggren, W. R. Salaneck, [278] B. Kretz, D. A. Egger, E. Zojer, Adv. Sci. 2015, 2, 1400016.
Appl. Phys. Lett. 2006, 89, 213503. [279] A. El-Sayed, P. Borghetti, E. Goiri, C. Rogero, L. Floreano, G. Lovat,
[253] S. Braun, W. Osikowicz, Y. Wang, W. R. Salaneck, Org. Electron. D. J. Mowbray, J. L. Cabellos, Y. Wakayama, A. Rubio, J. E. Ortega,
2007, 8, 14. D. G. de Oteyza, ACS Nano 2013, 7, 6914.
[254] S. Braun, W. R. Salaneck, Chem. Phys. Lett. 2007, 438, 259. [280] A. Kovalchuk, T. Abu-Husein, D. Fracasso, D. A. Egger, E. Zojer,
[255] P. Amsalem, J. Niederhausen, A. Wilke, G. Heimel, R. Schlesinger, M. Zharnikov, A. Terfort, R. C. Chiechi, Chem. Sci. 2016, 7, 781.
S. Winkler, A. Vollmer, J. P. Rabe, N. Koch, Phys. Rev. B 2013, 87, [281] A. Vilan, D. Aswal, D. Cahen, Chem. Rev. 2017, 117, 4248.
035440. [282] D. A. Egger, F. Rissner, E. Zojer, G. Heimel, Adv. Mater. 2012, 24,
[256] M. Gruenewald, L. K. Schirra, P. Winget, M. Kozlik, 4403.
P. F. Ndione, A. K. Sigdel, J. J. Berry, R. Forker, J.-L. Brédas, T. Fritz, [283] V. Obersteiner, D. A. Egger, G. Heimel, E. Zojer, J. Phys. Chem. C
O. L. A. Monti, J. Phys. Chem. C 2015, 119, 4865. 2014, 118, 22395.
[257] M. Hollerer, D. Lüftner, P. Hurdax, T. Ules, S. Soubatch, [284] V. Obersteiner, D. A. Egger, E. Zojer, J. Phys. Chem. C 2015, 119,
F. S. Tautz, G. Koller, P. Puschnig, M. Sterrer, M. G. Ramsey, ACS 21198.
Nano 2017, 11, 6252. [285] V. Obersteiner, G. Huhs, N. Papior, E. Zojer, Nano Lett. 2017, 17, 7350.
[258] X. Yang, I. Krieger, D. Lüftner, S. Weiß, T. Heepenstrick, [286] C. Winkler, S. S. Harivyasi, E. Zojer, 2D Mater. 2018, 5, 035019.
M. Hollerer, P. Hurdax, G. Koller, M. Sokolowski, P. Puschnig, [287] V. Obersteiner, A. Jeindl, J. Götz, A. Perveaux, O. T. Hofmann,
M. G. Ramsey, F. S. Tautz, S. Soubatch, Chem. Commun. 2018, 54, E. Zojer, Adv. Mater. 2017, 29, 1700888.
9039. [288] T. Joshi, C. Chen, H. Li, C. S. Diercks, G. Wang, P. J. Waller, H. Li,
[259] J. Repp, W. Steurer, I. Scivetti, M. Persson, L. Gross, G. Meyer, J.-L. Bredas, O. M. Yaghi, M. F. Crommie, Adv. Mater. 2019, 31,
Phys. Rev. Lett. 2016, 117, 146102. 1805941.

Adv. Mater. Interfaces 2019, 6, 1900581 1900581  (32 of 32) © 2019 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like