You are on page 1of 27

Molecular Self-Assembled Monolayers and Multilayers

for Organic and Unconventional Inorganic Thin-Film


Transistor Applications
By Sara A. DiBenedetto, Antonio Facchetti,* Mark A. Ratner,* and Tobin J. Marks*
1. Introduction
Organic thin lm transistors (OTFTs) are desired for the
manufacture of low-cost electronic devices such as electronic
pricetags, postage stamps, RFID tags, smart cards, exible
electronic paper, and backplane circuitry for active matrix
displays.
[13]
While much of the attention of the organic TFT
community has been focused on the search for high-mobility,
[46]
ambient stable,
[79]
and solution-processable small molecule
[1017]
and polymeric
[1820]
semiconductor materials,
[2130]
it is nowclear
that substantial improvements in TFT performance can also be
achieved by replacing and/or modifying the gate dielectric.
[1,3136]
As will be discussed here, self-assembled monolayers (SAMs) and
multilayers (SAMTs), such as self-assembled nanodielectrics
(SANDs),
[37]
are gaining signicant attention as gate dielectrics
due to their robust insulating properties, tunable thicknesses, and
efcient solution processability.
[38,39]
The electrical properties of SAMs com-
posed of Cd
2
salts of fatty acids having
different chain lengths were rst character-
ized by Mann and Kuhn in 1971.
[40]
This
work led to the development of standard
deposition procedures for ordered mono-
layers via formation of chemical bonds
between the precursor molecules and the
substrate/conductive electrodes. Note that
although organic monolayers fabricated
with the LangmuirBlodgett technique
were rst reported in 1920,
[41]
these lms
are very delicate because of the weak
physisorption to the substrate. In this
Review, we classify various SAM types by
the material they are typically deposited on,
either metals or oxides. While results for
both types of substrates will be discussed,
slightly more emphasis will be given to
SAMs on Si/SiO
2
since this is the standard
substrate/conductive electrode in the OTFT
community. The article is organized as follows: after a brief
overview of OTFTs in Section 1.1, we review the different
methods used to deposit SAMs on metal and oxide surfaces, and
introduce the fabrication of SAMTs (Section 2). A theoretical and
experimental overview of SAM electronic transport is given in
Section 3, and applications of SAMs as surface treatments, gate
dielectrics, and active semiconducting channels in OTFTs are
presented in Section 4.
1.1. Dielectrics for TFTs. Operation and Requirements
There are a number of excellent recent reviews describing the
details of OTFT operation,
[23,27,29,36,4244]
and here we only
introduce the relevant OTFT material components and basic
properties. The three fundamental OTFT materials components
are the contacts (source, drain, and gate), the semiconductor, and
the dielectric.
The arrangement of these components can be classied as
either bottom-gate or top-gate, and there are four common device
geometries (Fig. 1). Here, we focus mainly on the bottom gate
device geometries because they are the most commonly
employed with SAM and SAMTdielectrics. In top-contact devices
the source and drain electrodes are dened on top of the
semiconductor, whereas in bottom-contact devices the semi-
conductor is deposited on top of the source/drain and dielectric
R
E
V
I
E
W
www.advmat.de
[*] Dr. A. Facchetti, Prof. M. A. Ratner, Prof. T. J. Marks,
Dr. S. A. DiBenedetto
Department of Chemistry and the Materials Research Center
Northwestern University
2145 Sheridan Road, Evanston, IL 60208 (USA)
E-mail: a-facchetti@northwestern.edu; ratner@northwestern.edu;
t-marks@northwestern.edu
DOI: 10.1002/adma.200803267
Principal goals in organic thin-lm transistor (OTFT) gate dielectric research
include achieving: (i) low gate leakage currents and good chemical/thermal
stability, (ii) minimized interface trap state densities to maximize charge
transport efciency, (iii) compatibility with both p- and n- channel organic
semiconductors, (iv) enhanced capacitance to lower OTFT operating vol-
tages, and (v) efcient fabrication via solution-phase processing methods. In
this Review, we focus on a prominent class of alternative gate dielectric
materials: self-assembled monolayers (SAMs) and multilayers (SAMTs) of
organic molecules having good insulating properties and large capacitance
values, requisite properties for addressing these challenges. We rst describe
the formation and properties of SAMs on various surfaces (metals and
oxides), followed by a discussion of fundamental factors governing charge
transport through SAMs. The last section focuses on the roles that SAMs and
SAMTs play in OTFTs, such as surface treatments, gate dielectrics, and nally
as the semiconductor layer in ultra-thin OTFTs.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1407
R
E
V
I
E
W
www.advmat.de
Antonio Facchetti obtained his Ph.D in Chemical Sciences (University of Milan, Italy) under Giorgio
Pagani. He carried out postdoctoral research at the University of California (Berkeley, USA) with Prof.
Andrew Streitwieser and then at Northwestern University with Prof. Tobin J. Marks. In 2002 he joined
Northwestern University where he is currently an Associate Professor Adjunct. Dr. Facchettis
research interests include organic semiconductors and dielectrics for thin-lm transistors, molecular
electronics, organic nonlinear optical materials, and organic photovoltaics.
Mark A. Ratner is Morrison Professor of Chemistry and Professor of Materials Science and
Engineering at Northwestern University. He received his B.S. degree from Harvard University (1964)
and his Ph.D. from Northwestern University (1969) under G. Ludwig Hofacker. He is interested in
structure and function at the nanoscale, and theory of chemical processes, and tries to unite structure
and function in molecular nanostructures, based on theoretical notions, exemplary calculations, and
(importantly) collaborations. Interest areas are molecular electronics, self-assembly, nonlinear
response, and exact and approximate theories of quantum dynamics and using nanoscience to attack
the energy problems facing this world.
Tobin J. Marks is Ipatieff Professor of Chemistry and Professor of Materials Science and Engineering
at Northwestern University. He received his B.S. degree from the University of Maryland (1966) and
his Ph.D. degree from MIT (1971). His research interests include mechanistic organometallic
chemistry and catalysis, optoelectronics, chemical vapor deposition, and molecular electronics.
Sara A. DiBenedetto is currently pursuing her Ph.D. degree under the supervision of Professors Tobin
J. Marks and Mark A. Ratner at Northwestern University. She obtained her undergraduate degree in
chemistry at Agnes Scott College, Decatur, Georgia (2004). Her research focuses on computational
and physical chemistry for the development of hybrid organic-inorganic dielectric materials for thin
lm transistors.
1408 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
components. In top-gate arrangements, the gate lines are simply
deposited on top of the semiconductor layer. The basic equations
describing the OTFT drain current are
[45]
I
SD

lin

W
L
_ _
mC V
G
V
T

V
SD
2
_ _
V
SD
(1)
I
SD

sat

W
2L
_ _
mC V
G
V
T

2
(2)
where m is the eld-effect carrier (electron or hole) mobility of the
semiconductor, W the channel width, L the channel length, C the
capacitance per unit area of the dielectric layer, V
T
the threshold
voltage, V
SD
the drain voltage, and V
G
the gate voltage. OTFT
currentvoltage (IV) characteristics are typically investigated in
two regimes, the linear current regime (Eq. 1) initially observed at
low drain voltages (V
SD
<V
G
V
T
), and the saturation regime
(Eq. 2) when the drain voltage exceeds the gate voltage. In contrast
to conventional Si transistors, organic TFTs normally operate in
accumulation mode, where applying a gate voltage enhances
channel conductivity. The semiconductor eld-effect mobility is
calculated from the IV data according to either Equation 1 or 2.
The device current I
on
:I
off
ratio, and the subthreshold slope
(related to how efciently the gate eld modulates the off to on
current and how abruptly the device turns on) are also important
TFT characteristics. These gures of merit are used to compare
the electrical properties of devices fabricated with different
materials components.
Traditionally, OTFTs have been fabricated
using a thick SiO
2
layer (usually >100 nm
thick) as the gate dielectric. This material
prevents gate leakage currents, acts as an
effective capacitor, and allows for the accurate
measurement of the electrical performance of
the organic semiconducting layer. The major
motivation to search for alternative gate dielec-
trics is to enable inexpensive fabrication (e.g., by
printing near room temperature) and to
signicantly reduce the OTFT operating vol-
tages. According to Equation 1 and 2, a viable
approach to substantially increase the drain
current, while operating at low biases, is to
increase the capacitance of the dielectric;
[46]
for
a planar structure C"
0
(kA/d), where k is the
dielectric constant, A the area of the electrodes,
and d the dielectric thickness. However,
alternative OTFT gate dielectrics should also
have low gate leakage currents, and be able to
sustain the maximum possible electric dis-
placement D
max
"
0
"E
2
B
, where E
B
is the
dielectric breakdown eld.
[4749]
The observable
dielectric parameters, including E
B
, J (leakage
current density), and e (dielectric permittivity)
are easily measured in two terminal, planar
metal-insulator-metal(semiconductor) MIM(S)
devices.
[1]
One of the primary effects of SAMs on the
response of MIM(S) and OTFT devices the inuence of a
proximate molecular dipole moment on the semiconductor
electron afnity or on the metal work function. The effect of a
dipole layer on the surface potential can be estimated from the
Helmholtz equation (Eq. 3).
[50,51]
DV
Nmcos u
""
0
(3)
Here N is the dipole density (cm
2
), m the dipole moment
(debyes, D), u the average angle the dipole makes with the surface
normal, and e is the dielectric permittivity. Numerous studies
have documented how SAM dipole moment ultimately mediates
the charge conduction/injection from the substrate (bottom
electrode)
[50,52,53]
and the conduction/injection through the SAM
and into the semiconductor channel of the OTFT.
[5456]
While, the
main focus of this Review is on SAMs and SAMTs for use as gate
dielectrics in OTFTs, it is clear from the Helmholtz relation (Eq.
3) that the interpretation of the electrical properties is complicated
by the intrinsic interplay between the electrostatics of the SAM
itself
[50]
and the electrical interactions between the SAM and the
underlying electrodes/semiconductors,
[5761]
which will differ
depending on the exact TFTcontact geometry.
[6264]
Despite some
of the unresolved aspects of the SAM dielectric-semiconductor
interactions,
[65]
this Review demonstrates how SAMs and SAMTs
are rapidly advancing the eld of plastic electronics. Therefore, we
survey many aspects of SAMs that should be considered in TFT
applications, starting with a description of SAMs and SAMTs
having various chemical structures on metal and oxide surfaces.
Figure 1. A) Common TFT device geometries. B) Examples of typical p-type (group on left), and
n-type (group on right) small molecule organic semiconductors.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1409
R
E
V
I
E
W
www.advmat.de
2. Fabrication of SAMs and SAMTs
SAMs are ordered molecular assemblies formed by the
spontaneous adsorption of an active molecular precursor onto
a solid surface. Usually the precursor molecular species are
dissolved in common solvents, however SAMs can be deposited
by other techniques, such as vapor deposition, as well. Since there
are detailed descriptions of the chemistries, structures, and
characterization of various SAMs available in the literature,
[6672]
here we only briey summarize the techniques and molecular
structures of common SAMs and SAMTs.
2.1. SAMs on Metals
The most extensively investigated types of SAMs are alkanethiols
on gold,
[7380]
silver,
[8184]
copper,
[8587]
palladium,
[88,89]
and plati-
num
[90,91]
substrates. However, Au is most commonly used because
it is easy to deposit as planar thin lms and to pattern with
conventional lithographic tools or chemical etchants. Further-
more, Au is reasonably inert to oxidation and readily binds
organothiols.
[92]
The most common procedure for SAM deposi-
tion is substrate immersion in a dilute solution of the target thiol
at room temperature for 1218 h (due to slow reorganization
processes during lm growth).
[93,94]
The choice of solvent,
solution temperature, concentration, and immersion time are
among the various factors affecting the structure of the resulting
SAM.
[92,95,96]
Examples of thiol (RSH) and dithiol (HSRSH)
molecular structures (Fig. 2A) typically deposited on Au for
electronic applications are alkane(di) thiols,
[9799]
and those based
on oligophenylenes (OPs),
[100103]
oligo(phenyleneethynylenes)
(OPEs), and oligo(phenylenevinylenes) (OPVs),
[104109]
where the
labile thioacetyl (RSAc) or disulde (RSSR) functional
groups are sometimes used for substrate
chemisorption instead of the RSH group.
Thiol-derived SAMs with interesting function-
alities (Fig. 2B), such as terthiophenes,
[110112]
azo-groups,
[113117]
and tetracyanoquinodi-
methane
[118,119]
have also been investigated.
Alkane(di)thiols form densely packed and well-
ordered domains of up to several hundreds of
square nanometers on Au,
[120]
however the
nature of the metalsulfur bond and the spatial
arrangement of the sulfur groups are still
controversial topics. Nevertheless, alkanethiols
have been termed the benchmark for any
new technology related to molecular electronics
since the tunneling electrical properties are
relatively well documented/characterized.
[120,121]
2.2. SAMs on Oxides
The use of organosilane precursors (RSiX
3
, with
XCl, OMe, OEt) to form monolayers requires
hydroxylated substrate surfaces, including (but
not limited to) the technologically relevant
surfaces of SiO
2
, Al
2
O
3
, and tin-doped indium
oxide (ITO).
[122126]
In the case of SiO
2
surfaces,
the driving force for self-assembly is the in situ
formation of siloxanes, which connect the precursor silane to the
surface silanol (SiOH) groups via very strong SiOSi bonds
(Fig. 3).
[127]
Since the substrate surfaces are amorphous, the
packing and ordering of the chemisorbed organosilanes are
determined by the underlying siloxane network, by interchain
interactions, and by the reaction temperature.
[128]
Also, silanes
with particularly short chain lengths and high vapor pressures
[such as hexamethyldisilazane (HMDS)] can be deposited on
hydroxylated surfaces fromthe vapor phase by simple exposure to
the molecular vapor at room temperature, or by heating, or by
exposure under vacuum.
[68,129]
Figure 3A illustrates the struc-
tures of commonly used SAM silane precursors, such as simple
alkane chains octyltrichlorosilane and octadecyltrichlorosilane
(OTS and ODTS, respectively), 3-mercaptopropyltrimethoxysilane
(MPTMS), hexamethyldisilizane (HMDS), and various types of
functionalized sp molecules.
Other classes of materials deposited on oxide surfaces include
n-alkanoic acids (carboxylic end groups, Fig. 3B) and phosphonic
acids (Fig. 3C). These classes of molecules have gained attention
due to their ability to bind to a wide range of metal oxide surfaces
(AgO, Al
2
O
3
, ITO) and to form robust SAMs of similar quality to
that of thiols on Au (which may be useful for various technologies
involving bottom contact electrodes other than Au).
[130,131]
In the
case of organophosphonate SAMs, both vapor and solution
deposition have been demonstrated. Using XPS it was shown that
vapor-phase deposited phenylphosphonic acid (PPOA) reacts
with the alumina surface to form POAl bonds.
[132]
For a series
of solution deposited phosphonate SAMs, the inuence of alkyl
chain length on deposition was investigated.
[129]
It was found that
there is a distinct packing difference between short (C
10
C
13
)
alkyl and long (C
16
C
18
) alkyl phosphate SAMs. For SAMs on
TiO
2
surfaces, shorter molecules assemble into a less dense,
Figure 2. Some examples of the types of molecular structures used to make SAMs on Au for
electronic applications. A) Mono(di)thiols, and molecular wires (OPE, OPVs, and OPs). B)
Molecular wire examples of azo (top), thiophene (middle), cyanoquino alkanedisulde (bottom)
functional groups.
1410 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
liquid-like structure as compared to longer
chains.
[129]
The difference is most likely due to
strong bridging bidentate bonding with the
surface for short chains, as opposed to long
chains, which are stabilized via the enthalpy
gain due to van der Waal interactions between
the tightly packed alkyl chains.
SAMTs are becoming increasingly impor-
tant as gate dielectrics in TFTs (as will be
described in Section 4). There are various
methods for depositing multilayer lms
described in the literature.
[133135]
Our group
at Northwestern University has been investi-
gating a special type of SAMTs, SANDs
composed of alternating layers of s and p
constituent molecules. In addition, multilayers
of diphosphonic acids alternating with (and
held together by) Zr
4
ions have also been
demonstrated.
[68,136]
See Figure 3D for exam-
ples of general molecular precursor structures
used in SAMTs.
The present two-step method of fabricating
SANDs (Fig. 4) involves an iterative combination
of: (i) self-limiting chemisorption of siloxane
building blocks, such as a,v-difunctionalized
hydrocarbon chains (Alk), or highly polarizable,
siloxy-protected stilbazolium layers (Stb), and
(ii) in situ siloxy group removal concurrent with
capping using an octachlorotrisiloxane-
derived layer (Cap).
[37]
This second step
deposits a robust polysiloxane layer ($0.8 nm
thick), which is essential for stabilizing/
planarizing the molecular layer and regenerat-
ing a reactive hydroxyl surface for subsequent
monolayer deposition.
[137]
The different types
of multilayers are identied by the combination
of different layers according to the following
nomenclature: Alk Cap (type I), Stb Cap
(type II), and Alk Cap Stb Cap (type III).
The microstructures and electrical properties
of IIII have been characterized by specular
X-ray reectivity, standing wave X-ray reectiv-
ity, optical absorption spectroscopy, optical
second-harmonic generation measurements,
atomic force microscopy (AFM), MIM(S)
devices, cyclic voltammetry, and scanning
electron microscopy (SEM).
[1]
2.3. SAMs on H-Passivated Si
For certain organic electronic applications
(such as high capacitance gate dielectrics) it
is desirable to study the direct interface
between the organic SAM and Si without the
inuence of the native oxide layer.
[70]
However,
depositing SAMs directly on the Si surface
requires removing the native oxide, forming a
reactive surface, and changing the anchoring
Figure 4. Depiction of the molecular reagents used in the growth of SANDs, and fabrication
scheme of these components to make the multilayers: types IIII. Reproduced with permission
from [1]. Copyright 2005 Wiley-VCH Verlag GmbH.
Figure 3. Examples of molecular structures used for self-assembly on oxide surfaces: A) silanes,
B) carboxylic acids, C) phosphonic acids, and D) various precursor structures for SAMTs.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1411
R
E
V
I
E
W
www.advmat.de
group of the molecule (since trichlorosilanes do not react cleanly
with the H-passivated Si surface). The reactive Si surface can be
generated either by hydrogen or halogen termination or by
surface reconstruction. The subsequent deposition of the organic
SAM is effected by either SiC or SiO bond formation.
[70,138]
There are several types of surface reactions to create these types of
SAMs, including: hydrosilylation, thermally induced hydrosilyla-
tion, UV-induced hydrosilylation, and chemical grafting using
alcohols or organolithium/organomagnesium reactions.
[139,140]
One of the more extensively studied methods uses the
functionalized reconstructed surface under ultrahigh vacuum
(UHV), where molecular precursors with an unsaturated head
group attach to the HSi surface via a cycloaddition reaction.
[141]
Many studies have used the SiH surface to investigate the
electrical properties of SAMs, and a recent review describes these
types of systems in more detail.
[70]
3. Electrical Characterization of SAMs
SAMs sandwiched between conductive electrodes (see Fig. 5 for
various test-beds) can be represented by a MIM(S) structure.
SAMs of aliphatic chains are expected to be dielectric in
nature, due to the very large gap between the highest occupied
molecular orbital (HOMO) and the lowest occupied molecular
orbital (LUMO). Utilizing this wide gap, well-ordered SAMs
of alkyl chains exhibiting leakage current densities (J)
$10
8
10
6
Acm
2
have been successfully integrated in OTFTs
as surface treatments on thick oxides and as gate dielectrics
(Section 4.3). SAMs of conjugated molecular structures have also
been extensively studied as molecular rectiers.
[142148]
As
conjugated SAMs become more widely used in (and on) the
gate dielectric of TFTs, it is important to consider the
consequences that their smaller HOMOLUMO gaps may have
on the conduction mechanism of the SAM.
[149]
There is abundant
literature on conduction mechanisms
[84,150153]
and on charge
transport in SAMs;
[70,118,120,144,145,154158]
therefore we present a
summary of some of the relevant conduction mechanisms and
the most recent results relating to various SAMs.
3.1. Tunneling
Nonresonant tunneling (through bonds) is the most common
transport mechanism observed in molecular SAMs,
[159163]
however for p-conjugated molecular SAMs, resonant tunneling
(through the molecular orbitals) may also occur due to the smaller
HOMOLUMO gap.
[164170]
The simplest tunneling model
assumes a nite potential barrier at the metalinsulator interface
and describes the nite probability for electrons to travel a short
distance into the SAM (or insulator) despite the lack of available
energy levels. This processes is given by the Simmons relation
(Eq. 4), which is expressed here in the simplest form for a
rectangular barrier to demonstrate the exponential dependence of
the current density (J
DT
) on the thickness (d) and barrier
height (f)
[70,144]
J
DT

q
2
V
h
2
d
2mf
1=2
exp
4pd
h
2mf
1=2
_ _
(4)
where q electron charge, Vapplied voltage, h Plancks
constant, and melectron mass. At low voltages Equation 4
can be simplied to J
DT
/ (1/d)exp(bd), where the tunneling
decay parameter, b (Eq. 5)
[171]
b
4p2mf
h
a (5)
is accepted to be $0.61 A

1
for saturated alkanes, and 0.20.6
A

1
for conjugated molecules, and smaller b indicates more
efcient tunneling.
[145]
Here a is a unitless parameter describing
asymmetry of the potential prole (a1 for a rectangular
barrier). For very large applied voltages (V>f) the barrier
changes to a triangular shape, and the tunneling current is given
by the FowlerNordheim equation (Eq. 6).
[70,172175]
J
FN

q
3
E
2
8phf
FN
exp
4

2m

p
3qhE
qf
FN

3=2
_ _
(6)
Here f
FN
is the tunneling barrier height, E the electric eld
(V/d), and m
*
is the effective electron mass. FowlerNordheim
emission has the strongest dependence on the applied voltage,
but (like Eq. 4) is essentially independent of the temperature
(since it is a pure electronic tunneling).
Figure 5. Representative testbeds used to measure the electrical proper-
ties of SAMs. A) Scanning tunneling microscopy. Reprinted with permission
from[155]. Copyright 2004, American Chemical Society. B) Hg-SAM-SAM-Metal
(Au) device. C) Crossed wires (Au wires). Figure 1b and 1c reproduced with
permission from [120]. Copyright 2008, Institute of Physics. D) Metal-
insulator-semiconductor. Reprinted with permission from [174]. Copyright
2006, American Physical Society. E) Nanopore. F) Nanoparticle-SAM-metal
(Au) device structure. Figure 1e and 1f reprinted with permission from [155].
Copyright 2004, American Chemical Society.
1412 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
As an example, conduction through SAMs derived from
v-substituted aromatic or heteroaromatic alkanethiols (Fig. 6A)
was investigated in AuSAMAu junctions, where the top Au
electrode was deposited by cathodic deposition directly on top of
the monolayer.
[176]
High breakdown eld strengths (E
B
)>50 MV
cm
1
were measured, and it was concluded that the terminal
aromatic rings have strong pp interactions which densify the
surface, hindering the penetration of metal. Furthermore, the
authors modied the typical square-well tunneling potential to t
the JV data within the direct tunneling model, but did not
investigate the J(V, T) dependence. In a recent study, edge
molecular junctions were used to investigate the electrical
properties of conjugated 4,4
0
-biphenyldithiol (BPDT) SAMs
compared to an aliphatic C
9
dithiol alkyl SAMs (Fig. 6B).
[177]
The top electrode (either Ag or Au) was vapor-deposited at
reduced temperatures and at a 608 angle to create the edge
structure. Conduction through the conjugated BPDT SAM was
determined to be via tunneling, despite the larger currents and
smaller HOMOLUMO gap than in the C
9
SAM.
Conduction in SAMs also depends on the
nature of the chemical linker that binds the
molecule to the bottom metal electrode.
[55,178]
For example, the conductance differences
within a series of 1,4-butylene alkanes termi-
nated with dimethyl phosphine, methyl sulde,
or amine (Fig. 6C) was compared.
[179]
The
authors assumed nonresonant tunneling as the
dominant conduction mechanism, and found
that phosphine termination provides the lowest
contact resistance (highest conductance) of the
series. Similarly, a correlation between mole-
cular structure and electrical resistance (R) in
the nonresonant tunneling regime was demon-
strated in another study by a systematic
comparison of alkyl versus conjugated and
mono- versus di-thiol substituted bridging
molecules (Fig. 6D).
[180]
Based on a multi-
barrier tunneling model,
[181]
for a given chain
length R, the resistance of the monothiol
junction was roughly 2 greater than the
dithiol junction, most likely due to the proper-
ties of the chemisorbed versus physisorbed
nature of the top contact. Interestingly, the
contact resistance of the alkane dithiols and
their conjugated analogs were found
[180]
to be
similar and independent of the differing
HOMOLUMOgaps of the different molecular
structures.
The conduction mechanism of SAMs on
oxide surfaces is far less developed than for
SAMs on Au. For example, in 1996 Vuillaume
et al. reported a suppression of tunneling in
metal/alkylsilane/SiO
2
/Si structures, resulting
in low conductivities (which is advantageous
for TFT gate dielectric applications as will be
discussed in Section 4). The conduction was
shown to be independent of monolayer thick-
ness (Fig. 7A), indicating that tunneling is not
the dominant conduction mechanism, how-
ever no alternative mechanism was suggested.
[182]
In 1999,
Waldeck and co-workers performed photocurrent measurements
on Si/SiO
2
coated with different alkylsilanes and observed a much
weaker dependence of the conduction on monolayer thickness
than expected (Fig. 7B), and the authors suggested hopping
through traps in the lm (see below) as a possible mechanism to
explain the weak distance dependence.
[183]
In 2002, Cahen and
co-workers measured the conductance through alkanes in Hg/
alkylsilane/SiO
2
/Si MIS devices.
[165]
The results are striking since
at rst glance it would appear that there is a clear dependence of
the current density on the chain length (Fig. 7C). However, the
authors explained that the curves are essentially identical within
the experimental uncertainty. The results are then in agreement
with the results from the earlier studies, where the conduction in
alkylsilanes on native SiO
2
does not depend on chain length.
Furthermore, the J values of 10
6
Acm
2
at 1 Vare lower than the
theoretical model for tunneling, even with an added tunneling
barrier (where the added barrier was included to account for the
possibility that not all of the silane chains are bonded to the SiO
2
).
Figure 6. Examples of molecular structures used in Au-SAM-Au measurements.
A) v-Substituted aromatic or heteroaromatic alkanethiol used in [176]. B) Illustration of an
edge Au/SAM/Ag device, and corresponding IV curves for AuC9Ag and AuBPDTAg (ex),
and theoretical estimation of the C9 response (Th). Adpated with permission from [177].
Copyright 2007, American Institute of Physics. C) Molecular structures of n-butylene diamine,
diphosphine, and disulde molecules used in breakjunctions of [179]. D) Structures of alkyl
versus conjugated polymeric structures used in [180].
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1413
R
E
V
I
E
W
www.advmat.de
Again, the authors suggest hopping as the transport mechanism
since the barrier heights are similar to those for hopping in other
bilayer structures.
[121,184]
3.2. Thermally Activated Processes
Hopping refers to Ohmic transport, which is typically dominant
at low elds and moderate temperatures. For simple hopping,
J follows a classical Arrhenius relation (Eq. 7),
[152,185]
s s
0
exp
E
a
kT
_ _
(7)
where s is the conductivity, s J/E (EV/d), and E
a
is
the activation energy. McCreery
[144]
associated hopping with
nuclear motion (or molecular reorganization), and Bassler and
co-workers
[186188]
proposed that hopping occurs among large
defects or impurities. Whatever the origin of the hopping, it is
observed much less frequently in SAMs than is tunneling
because the typical lengths of the molecules investigated are
seldom greater than $2 nm.
[162,189,190]
Extensive theoretical work has been reported concerning the
tunneling/hopping transition,
[154,191197]
and earlier data on some
organics
[198201]
and DNA
[202,203]
have clearly revealed a length
dependent transition between tunneling and hopping conduction.
Recently, the transition was demonstrated in Au-molecule-Au
junctions, where the electrical resistance of oligophenyleneimine
(OPI) molecules of various lengths were measured with a
conductive AFM tip (Fig. 8).
[204]
Hopping transport was found in
OPI molecules longer than 4 nm, while molecules <4 nm in
length exhibited nonresonant tunneling. Furthermore, the
authors intentionally disrupted the molecular conjugation with
cyclohexyl groups and found that conduction in molecules with
hopping transport (>4 nm length) was affected far more than
molecules with tunneling transport (<4 nm in length).
In one of the earliest conduction studies, thermally activated
transport through SAMs of thioacetylbiphenyl
[205]
(molecular
length $1.2 nm) in Au/SAM/Au nanopore devices (Fig. 5E) was
attributed to hopping. Using a physisorbed aryl-Ti top contact, the
SAM/Au barrier height was estimated to be 0.22 eV according to
the thermal emission (Schottky emission, see below) model. For
the direction of transport through the Au-thiol bond (negative
applied voltages), the authors proposed hopping transport
through the SAM with a barrier of 0.19 eV (Fig. 9A). However,
these authors pointed out that the nature of the hopping barrier
(defects in the SAM, hopping along the wire, or between
molecules) is unclear. In 1999, this work was extended to
Figure 8. Change in resistance of OPI molecules based on molecular
length, and OPI molecular structure (OPI 4). Reprinted with permission
from [204]. Copyright 2008, AAAS.
Figure 7. Conduction of alkylsilane SAMs on SiO
2
. A) conductivity versus
thickness of alkylsilane SAMs. Reprinted with permission from [182]. Copy-
right 1996, American Physical Society. B) Photocurrent versus chain alkyl
chain length. Reprinted with permission from [183]. Copyright 2004, Amer-
ican Chemical Society. C) JV for different length SAMalkyl chains. Reprinted
with permission from [165]. Copyright 2004, American Chemical Society.
1414 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
elucidate the nature of hopping in conjugated SAMs
[206]
by
investigating phenylene diisocyanide in similar nanopore devices
(Fig. 9B). It was observed that the dominant conduction
mechanism depends on the defect level introduced during the
fabrication process. For very large defect densities (typically in
devices with evaporated metal contacts), only hopping was
observed with a barrier of $0.3 eV. However both hopping and
thermionic emission (see below) were observed when the defect
level was reduced (with corresponding barrier heights similar to
the previous study of thioacetylbiphenyl SAMs).
The energy distribution of electrons in metals is given by the
FermiDirac distribution function. This implies that at elevated
temperatures (and at larger applied biases than for hopping) a
larger fraction of the electrons will have sufcient energy to
surmount the energetic barrier presented by the SAM.
Thermionic (Schottky) emission (Eq. 8) assumes that an electron
from the contact can be injected into the dielectric once it has
acquired sufcient thermal energy to cross the potential
maximum resulting from the superposition of the external and
the imagecharge potential.
[186,207]
However, if the SAM has
structural imperfections (such as defects or impurity ions), these
defect states can act as electron traps. Thermally excited trapped
electrons will contribute to the current density according to
PooleFrenkel emission (Eq. 9) at high temperatures and
intermediate elds.
[70,172]
J
S
A

T
2
exp
q f
s

qE=4p""
0
_
_ _
kT
_
_
_
_
(8)
J
PF
/ E exp
q f
PF

qE=4p""
0
_
_ _
kT
_
_
_
_
(9)
Since both mechanisms result fromCoulombic lowering of the
potential barrier under an applied electric eld, the two processes
have similar J(V,T) dependencies. The square root term for
PooleFrenkel emission is greater than Schottky emission by a
factor of 2, and often the exponent of Eq. 8 and 9 are simpli-
ed by introducing a b parameter [b
S
(q
3
/4pee)
1/2
and b
PF
(q
3
/
pee)
1/2
].
[208]
Here, A* is the modied Richardsons constant
(A* 120 Acm
2
K
2
),
[186]
E is the electric eld, f
S
and f
PF
are the
Shottky and PooleFrenkel barrier heights, e is the dielectric
permittivity, e the permittivity of vacuum, and k the Boltzmann
constant.
[152]
As an example, thermally activated defect conduction was also
observed in cross-linkable organo-siloxane hybrid dielectrics,
synthesized by a solgel process.
[209]
The leakage current
characteristics were described by a PooleFrenkel emission
mechanism, related to traps in the bulk of the lm. In this case,
the lms are much thicker (260 nm) and the material is not
self-assembled. However, since the Cap layer (Fig. 4) of SANDs is
also a crosslinked siloxane, it is worthwhile to mention briey the
electronic properties of these alkoxysilane hybrid dielectrics.
From ATR-FTIR spectra (Fig. 10A), the authors observed a
reduction of the silanol n(SiOH) peak intensity and enhance-
ment of the transition corresponding to the siloxane bond
n(SiOSi) with increasing lm annealing temperatures (150,
170, and 190 8C). Furthermore, the leakage current density
(10
6
10
5
A cm
2
at 2 V) is lower for the lms annealed at
190 8C versus those annealed at 150 8C. The authors t the JV
data according to the PooleFrenkel mechanism and found that
the b parameter value extracted from the slope of the ts matched
the b parameter estimated from b
PF
(q
3
/pee)
1/2
. Since
PooleFrenkel emission is related to thermal excitation of
trapped electrons into the insulator conduction band, and since
it has been reported that electrons can be deeply trapped in
hydrated thermal silicon oxides (forming SiO

),
[210]
the authors
proposed that the silanol groups act as trap sites (Fig. 10B) in the
siloxane-based dielectric. Recently, inkjet printing of these
materials for gate dielectrics in TFTs was demonstrated,
[211]
which is important for the low cost development of low leakage
current gate dielectrics. Deposition of a high capacitance
self-assembled SAM has yet to be demonstrated, although very
recently inkjet printed source/drain electrodes were utilized in
top contact TFTs with a SAM-based gate dielectric,
[212]
and
organic single-crystal TFT arrays were demonstrated by relief
printing of thick ($13 nm) OTS lms.
[213]
Recently, the conduction mechanism of 3-aminopropyltri-
methoxysilane (APTMS) SAMTs on Si/SiO
2
was investigated, and
it was found that as thickness of the SAMTs is increased, the
Figure 9. Molecular structures of SAM precursors and corresponding
J(V,T) characteristics in Arrhenius plots for A) thioacetylbiphenyl. Rep-
rinted with permission from [205]. Copyright 1997, American Institute of
Physics. B) Phenylene diisocyanide. Reprinted with permission from [206].
Copyright 1996, Elsevier.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1415
R
E
V
I
E
W
www.advmat.de
current densities do not follow the typical dependence for
tunneling (Fig. 10B).
[214]
Also, hysteresis and negative differential
resistance
[215,216]
were observed from the SAMT devices, but not
from the monolayer devices. The authors suggested that the
transport in APTMS SAMTs is hopping. The Northwestern group
has also investigated annealing conditions and conduction
mechanisms in SAMTs (Fig. 4). For example, it was found that
SAND III leakage current densities do not change signicantly
with increasing annealing temperatures (see
Section 4.3.1.1 for results with multiple layers
of III),
[217]
suggesting similar transport in
SAND Cap layers to the crosslinked alkoxysi-
lane network described above. To further probe
the conduction of SANDs, the J(V,T) transport
characteristics of for II and III were studied
over the temperature range of 65100 8C. We
nd (in unpublished data) slightly different
conduction in II versus III as might be expected
from the different interfaces (p vs. s, respec-
tively) and thicknesses (3.2 and $6.0 nm,
respectively). Analysis of the J(V,T) character-
istics according to tunneling and injection
models will be reported shortly.
4. TFT Device Applications
In the previous sections we introduced the
importance of molecular orbital alignment,
since SAMs with different HOMOLUMO
gaps and/or dipoles can shift the metal work
function/semiconductor electron afnity and
change the barrier to conduction.
[54,218,219]
In
this section, we highlight some experimental
results where SAMs are used as surface
treatments in devices to modify electron
injection.
[220]
4.1. Metal Surfaces/Bottom Contact TFTs
Practical device technologies will most likely
depend on prepatterning of the metal electro-
des for bottom contact TFTs and integrated
circuits. In this conguration, the major
challenge is increased contact resistance arising
from several effects such as interfacial charge
migration, surface dipoles, the insulating nature
of semiconductor side chains, or physical
delamination/dewetting.
[221]
Therefore, a major
potential enhancement of SAMs in bottom
contact TFTs is the possibility of reducing
contact resistances by enhancing the semicon-
ductor adhesion and growth orientation relative
to the metal source/drain electrodes.
[64,222]
For
the commonly used organic semiconductor,
pentacene (P5, Fig. 1), surface modication is
essential to enhancing bottom contact TFT
performance since optimal wetting of the
substrate is essential for favorable large-grained rst layer P5
growth.
[221,223]
As an example, surface treatment of Au bottomcontact source/
drain electrodes with simple alkanethiols (such as
1-hexadecanethiol) have been shown to modify the surface
energy such that large P5 crystal grains form over large areas
both on top of and between the source/drain electrodes. These
TFTs exhibit a signicant increase in P5 mobility (Fig. 11A) from
0.16 cm
2
V
1
s
1
(without a SAM) to 0.48 cm
2
V
1
s
1
(with a
Figure 11. A) Pentacene TFT output characteristics for untreated (top) and SAM treated
(bottom) Au bottom contact source/drain electrodes. Reprinted with permission from [223].
Copyright 2001, IEEE. B) AFM images of pentacene grown on an untreated (top), aliphatic SAM
treated (middle), and aromatic SAM treated (bottom) bottom-contact Au source/drain electrodes.
Reprinted with permission from [225]. Copyright 2006, American Institute of Physics.
Figure 10. A) ATR-FTIR spectra of the solgel alkylsiloxane dielectrics, and cartoon of the
proposed SiO

electron traps. Adapted with permission from [209]. Copyright 2006, American
Institute of Physics. B) Semilog plot of measured J as a function of the multilayer thickness (squares),
and calculated theoretical J assuming tunneling (circles), and illustration of the APTMS multilayer
structure. Reprinted with permission from [214]. Copyright 2008, Springer.
1416 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
SAM).
[223]
However, alkanethiols in general have a large energy
gap between the HOMO and the LUMO, typically E
g
$6 eV,
[224]
and hence are good insulators.
[225]
Due to the insulating behavior,
reduced charge carrier injection is expected, which makes
alkanethiols not suitable as contact primers for OTFTs. In
contrast, aromatic SAMs have smaller MO gaps and hence are of
interest for functionalization of bottom contact electrodes in
TFTs.
To this end, several groups focused on comparing the effects of
aliphatic versus aromatic SAMs on the device performance of
bottomcontact TFTs (see Fig. 2, e.g., of molecular structures). The
two major advantages of using aromatic SAMs instead of aliphatic
SAMs are increased I
on
:I
off
ratios due to favorable electron
injection,
[226228]
and increased wettability of the SAM surface by
the semiconductor due to SAM-semiconductor attractive inter-
actions (Fig. 11B),
[225]
both of which lower the contact resistance
and improve TFTperformance.
[229,230]
In this example, generally
improved bottom contact TFT device characteristics were
achieved with an aromatic anthracene-2-thiol SAM (on/off ratio
of 10
6
and strongly reduced trap density as indicated by the low
subthreshold swing of 0.55 V decade
1
). AFM and SEM images
revealed signicant dewetting of P5 lms on untreated Au
electrodes, but no dewetting on both aliphatic and aromatic thiol
SAM treated Au electrodes (Fig. 11B). The authors concluded that
the contact resistance observed in TFTs without SAMs is related
to morphological irregularities. However, since the P5 lm
morphology is similar on both aromatic and
aliphatic SAMs, the contact resistance in TFTs
with aliphatic SAMs must arise from poor
electron injection.
[225]
In an effort to correlate molecular structure
with TFT performance, Katz et al. screened a
series of thiols and other commercially avail-
able sulfur reagents with a variety of terminal
functional groups as surface treatments for
TFTs using a uorinated naphthalenediimide
derivative as the semiconductor layer.
[221]
Several of the thiol treatments provided a
benecial effect compared to devices with
untreated electrodes. The highest TFT I
SD
currents were obtained with 2-chlorobenzyl
mercaptan, and the current enhancements
were attributed to increased wetting or sticking
of the semiconductor to the thiol-treated gold
surface versus bare gold, due either to favorable
arylaryl interactions or to COOH-carbonyl
hydrogen bonding between the SAMs and the
semiconductor (rather than due to different
dipole moment strengths).
In a different approach for SAM-Au treat-
ment, lower contact resistance was demon-
strated by directly coupling a molecular
monolayer similar in structure to that of P5
(the thioketone in Fig. 12A) to the source/drain
electrode surface prior to P5 deposition.
[229]
In
this conguration, the semiconductor SAM
surface treatment acts as a template for
enhanced P5 growth, and signicant enhance-
ment of TFT response characteristics over the
untreated and small molecule-thiol treated bottom contact TFTs
was observed. Recently, injection barrier effects of various SAMs
[thiophenol (TP), 4-uorothiophenol (4-FTP), or pentauorothio-
phenol (PFTP)] on bottom contact TFT performance were
investigated using TIPS-pentacene as the semiconductor (see
Fig. 12B for molecular structures, TFT electrical characteristics,
and proposed energy level diagram).
[227]
No TIPS-pentacene lm
morphology change was observed for lms deposited on both the
SAM-treated and bare Ag electrodes. However, a signicant work
function shift was measured (from 4.14 to 5.35 eV) by electrode
treatment with the SAMs, and as a result enhanced TFT
performance with SAM treated electrodes was attributed to better
alignment of the modied metal work function with the
semiconductor HOMO level. In untreated devices the Ag work
function (4.7 eV) and semiconductor HOMO energy level (5.3 eV)
mismatch creates a hole injection barrier of 0.6 eV. However, Ag
surface modication with 4-FTP or PFTP dipolar SAMs modies
the work function to 5.21 and 5.35 eV, respectively, signicantly
reducing the injection barrier and creating an ohmic contact with
reduced contact resistance.
4.2. Oxide Surfaces/Top Contact TFTs
One overall goal in the organic electronics eld has been to
enhance OTFT performance to reach metrics comparable to
Figure 12. A) Thioketone used for SAM, and cartoon of electrode/SAM/pentacene interface
(top). TFT transfer characteristics (bottom) for devices with SAM (red) and without SAM (blue).
Reprinted with permission from.[229] Copyright 2006, American Chemical Society. B) Bottom
contact TIPS-pentacene TFT output with SAMs of TP (top) and PFTP (middle) on the bottom Ag
source/drain electrodes, and corresponding energy level diagram (bottom). Reprinted with
permission from[228]. Copyright 2008, American Institute of Physics.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1417
R
E
V
I
E
W
www.advmat.de
amorphous hydrogenated silicon (a-Si:H), the current material
used in fabricating thin lm transistors (TFTs) for LC displays,
which exhibits an electron carrier mobility of $1.0 cm
2
V
1
s
1
and an I
on
:I
off
ratio >10
6
. P5 is one of the most promising organic
semiconductors for use as the active material in OTFTs because of
its relatively high mobility and large TFT I
on
:I
off
ratios, which are
comparable to or surpass those of amorphous silicon.
[231]
Thus,
many studies have employed SAM-modied SiO
2
as the gate
dielectric and P5 as the semiconductor (Table 1) to elucidate
fundamental correlations between dielectricsemiconductor
interfacial effects and TFT performance (see ref.
[42]
for a recent
review).
[232]
Typically it is observed that SAMs (as gate dielectric
surface treatments) in OTFTs achieve one or more of the
following: reduce the device subthreshold slope, increase the
semiconductor mobility, increase the I
on
: I
off
ratio, alter the
polarity of the majority charge carrier type, modulate the carrier
density in the channel, and shift the threshold voltage (compared
to control devices with bare gate dielectric oxides).
[65,232]
Explanations for the overall performance enhancements are
usually described in the context of the semiconductor lm growth
morphology
[35,231]
(crystallinity and grain size), and/or the
dielectricsemiconductor interfacial properties
[233]
(such as sur-
face energy, trap sites, molecular chemical functionality, and the
structure of the SAM itself).
The rst indications that the dielectricsemiconductor inter-
face has a large inuence on the overall OTFTperformance were
the observations of: (i) different measured mobility values for the
same organic semiconductor, and (ii) different lm growth
characteristics on different types of gate dielectrics. For example,
in 1997 Gundlach and co-workers deposited P5 on LiF and SiO
2
surfaces, and reported distinct differences in the AFM images of
P5 lms of varying thicknesses.
[234]
In 1992 Horowitz et al.
measured different mobility values for lms of the 6T
(sexithiophene, Fig. 1) semiconductor on different oxide and
polymer gate dielectrics,
[235]
and in 2005 Muccini and co-workers
studied the supermolecular organization of thin lms of 6T on
SiO
2
for a correlation between intermolecular interactions and
transport properties as compared to 6Tsingle crystals, which have
mobilities up to 20 cm
2
V
1
s
1
.
[236]
In a later study Horowitz
et al. used AFM images of very thin (1 nm) to thicker (7.5 nm) P5
lms deposited on top of 1-phosphonooctane SAMs (using Al/
Al
2
O
3
substrates) to conrm that the P5 growth mode is different
when the growth is near the dielectric surface compared to
growth in the bulk (Fig. 13A).
[237]
In this case, the average P5
mobility was 2 cm
2
V
1
s
1
(largest value $3cm
2
V
1
s
1
on the
1-phosphonohexadecane SAM), which was a signicant improve-
ment over the bare alumina control TFT of 1.1 cm
2
V
1
s
1
. The
authors claimed near single-crystal quality P5 growth during the
early stages of deposition, which explains the large observed
mobilities of the TFTdevices since the rst few monolayers of the
semiconductor lm are the active charge transporting region.
[237]
In 2006, Mottaghi and Horowitz observed similar differences in
P5 growth modes on bare versus eicosanoic acid
(CH
3
(CH
2
)
18
COOH) SAM modied Al/Al
2
O
3
gate dielec-
trics.
[238]
The P5 lms were modeled as multi-layer dielectrics,
and it was found that the bulk P5 mobility increases linearly with
the gate voltage up to about 5 V on the SAM-coated Al/Al
2
O
3
and
10 V on bare Al/Al
2
O
3
, then saturates at large values of $5cm
2
V
1
s
1
and $3 cm
2
V
1
s
1
, respectively (Fig. 13B), conrming
that the semiconductor mobility and growth mode are sensitive to
the dielectricsemiconductor interface.
Since, these pioneering studies, many research efforts have
been devoted to the investigation of the dielectricsemiconductor
interface, and major improvements in the OTFT electrical
characteristics have been documented.
[239248]
For example,
Forrest and co-workers observed major TFT mobility and I
on
:I
off
enhancement when using OTS-treated gate dielectrics in P5
TFTs, from 0.06 cm
2
V
1
s
1
and I
on
:I
off
$10
5
on bare SiO
2
, to as
large as 1.2 cm
2
V
1
s
1
and I
on
:I
off
$10
8
on OTS treated
SiO
2
.
[249]
Schwartz and co-workers found dramatic improve-
ments (I
on
:I
off
ratios of 10
8
and subthreshold slopes of 0.2 V
decade
1
) over other devices using octadecylsilane and other
phosphonates.
[250]
In this study, a new (anthracene)phosphonate
SAM was reported (Fig. 14C), and compared to the other
Table 1. Comparison of pentacene OTFT performance on various oxide gate dielectric surface treatment SAMs (m [cm
2
V
1
s
1
], current on/off ratio I
on
/
I
off
, threshold voltage V
T
[V], and subthreshold slope SS [V decade
1
]).
Reference SAMOxide m (I
on
/I
off
) V
T
(SS) Year
[233]
OTS SiO
2
0.4 (10
7
) (1.6) 1997
[248]
OTS iO
2
0.6 (10
7
) (0.6) 2002
[236]
1-Phosphonooctane Al
2
O
3
2.1 (1.7 10
6
) 2.1(2.5) 2003
[257]
Alkylsilane SiO
2
0.13 (10
7
) 5.0 2004
Peruoroalkylsilane SiO
2
0.20(10
8
) 17
Aminoalkylsilane SiO
2
2.4 10
3
(10
8
) 11
[251]
HMDS SiO
2
0.2 (10
5
) 0.8 2005
[242]
poly(imide-Piloxane) SiO
2
2005
n 0 0.11(10
4
) 8(3.1)
n 52 0.25(10
5
) 7.2(2.1)
n 130 0.56(5 10
5
) 5.4(1.7)
[237]
Eicosanoic acid Al
2
O
3
5 1.2 2006
[239]
HMDS SiO
2
0.5(10
6
) 1.1(2.9) 2006
HMDS Al
2
O
3
0.01(10
5
) 1.2(4.1)
[249]
(9-Anthracene)phosphonate SiO
2
(10
8
) 4.5(0.2) 2007
[254]
b-Phenethyltrichlorosilane 1.5(10
6
) 18 2007
[258]
ODTS disordered SiO
2
0.3(10
6
) 2008
ODTS ordered SiO
2
0.6(10
6
)
1418 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
organophosphonate SAMs tested. The anthracene-based SAM
exhibited the best TFT performance. The authors ascribed the
improvements to structural similarities between P5 and the SAM
substituent, which resembles a continuous anthracene lm, thus
providing nucleation sites for favorable growth of P5 islands.
Similarly, Mottaghi and Horowitz
[238]
explained OTFT perfor-
mance enhancements related to P5 grain size using the multiple
trapping and release model. If the initial increase in bulk P5
mobility with gate voltage can be explained by lling of traps
located in grain boundaries, then the smaller, more regular P5
grains observed on SAM treated Al/Al
2
O
3
should (and do) perform better than the larger,
more irregular grains observed on bare Al/
Al
2
O
3
.
Correlations between P5 lm grain size/
crystallinity and TFTmobility have been studied
in great detail,
[239,249,251255]
however exact
trends remain unclear,
[1,65,256]
and the behavior
is entangled with other effects (surface
energy,
[36]
surface roughness,
[257]
and molecular
structure
[258]
). For example, high aqueous
contact angles (8081158), or more hydrophobic
substrate surfaces, typically result in the largest
OTFT performance increase. However, it is not
always simple to isolate the unambiguous origin
of the improved performance. Cho and
co-workers, deposited P5 on ordered (alkyl
chains aligned and packed) and disordered
(loosely oriented alkyl chains) ODTS SAMs
(Fig. 15A), and a barrage of structural techni-
ques were used to analyze the structural and
morphological characteristics of both the ODTS
and the P5 lms.
[259]
While the surface energy
and surface roughness of the two ODTS SAMs
were measured to be the same, the OTFT
performance was signicantly different, with
enhanced performance observed in the more
ordered SAM. X-ray diffraction spectra (Fig. 15B) of the P5 lms
revealed more crystalline P5 lms when deposited on ordered
versus disordered SAMs. Furthermore, a larger diffraction peak
intensity ratio of the thin-lm phase to the bulk phase was
observed for the more ordered SAMs, indicating that P5
deposited on the more ordered ODTS monolayer is more
sensitive to interactions with the surface since the thin-lm phase
is the surface-induced (strained) phase.
[259]
In another example,
the grain size of P5 lms was reported to be unchanged when
deposited on HMDS-treated SiO
2
and bare SiO
2
; nevertheless,
differences in the OTFT performance charac-
teristics were again observed.
[252]
In this case, it
was proposed that the suppressed off-currents
observed for HMDS-treated devices were due
to a reduction in the density of interfacial
trapping states.
[260,261]
Strong evidence for this interfacial trapping
affect was presented by Friend and co-workers
in 2005.
[210]
In this study, the authors used
multiple-reection attenuated-total-reection
FTIR spectrometry (ATR-FTIR) to track the
changes of the SiOH stretching and bending
combination band near 3 8004 700 cm
1
with
time and applied voltage. Based on the long
times associated with peak intensity changes
and wavelength shifts compared to the time for
loss of TFT activity, the authors proposed that
the spectral shifts were due to the generation of
SiO

induced by deeply trapped electrons,


indicating electrochemical trapping of elec-
trons by SiOH surface groups. The interfacial
trapping effect was observed in TFTs using the
Figure 14. OTFT characteristics for pentacene deposited on SiO
2
/SAM dielectrics of A)
n-octadecylphosphonate (ODPA), B) (quarterthiophene)phosphonate, and C) (anthracene)pho-
sphonate. Adapted with permission from [250]. Copyright 2007, American Chemical Society.
Figure 13. A) AFM images of pentacene lms of thickness 1nm (top), 3.5 nm (middle), and
7.5 nm (bottom) grown on 1-phosphonohexadecane SAM. Reprinted with permission from
[237]. Copyright 2003, American Chemical Society. B) Cartoon of pentacene growth mode
differences and mobility saturation versus gate voltage for SAM and bare Al
2
O
3
gate dielectrics.
Reprinted with permission from [238]. Copyright 2006, Elsevier.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1419
R
E
V
I
E
W
www.advmat.de
polymer F8BT [poly(9,9
0
-dioctyluorene- co-benzothiadiazole)] as
the semiconductor and alkyl-SAMs of various lengths: hexam-
ethyldisilazane (C1), decyltrichlorosilane (C10) and ODTS (C18)
as Si/SiO
2
gate dielectric surface treatments. Initially, enhance-
ment of the n-channel TFT performance was observed with
increasing alkyl-SAM chain length (Fig. 16) and surface OH
passivation. However, TFT currents eventually degraded over
time, and since siloxane-terminated SAMs cannot completely
eliminate surface SiOH groups, the degradation was attributed to
those trap sites at or near the SiO
2
interface. This study not only
demonstrates the importance of the dielectricsemiconductor
interface on TFT performance, but also demonstrates that SAM
modication benets not only p-type (P5) TFTcharacteristics but
also a wide range of n-type semiconductor materials.
[7,8,244,262]
Lastly, the chemical composition of the semiconductor
dielectric interface can directly tune the semiconductor char-
acteristics in a different way. In addition to charge carrier
inversion (from p- to n-), it has been shown that the chemical
structure of the SAM molecules can modify the carrier density in
the semiconductor channel and induce appreciable threshold
voltage (V
T
) shifts. For example, P5 and C
60
I
SD
currents were
observed to depend strongly on the SAM molecular structure,
where the currents at V
G
0 V were enhanced by six orders of
magnitude in devices with SAMs composed of peruoroalk-
ylsilane molecules compared to devices fabricated with aminoalk-
ylsilane SAMs (Fig. 17A).
[258]
The V
T
values were shifted for the
amino- and uoro-functionalized SAMs, however the V
T
values
for the unsubstituted alkylsilane SAMand untreated devices were
essentially the same, suggesting that the additional carriers in the
semiconductor channel are generated by the built-in potentials
created by the different SAM dipole moments.
In 2007, Katz and co-workers capitalized on this effect to
fabricate complimentary circuits.
[263]
Using a nonpolar silane,
PTS (phenyltrimethoxysilane), and a series of uorinated dipolar
silanes trichloro(3,3,3-triuoropropyl)silane (FPTS), trichloro1H,
1H,2H,2H-peruorooctyl)-silane (FOTS), and trichloro(1H,1H,
2H,2H-peruorodecyl)silane (FDTS), they demonstrated control
over the doping level of the p-type semiconductor 5,5
0
-bis
(4-hexylphenyl)-2,2
0
-bithiophene (6PTTP6) and fabricated uni-
polar inverters (see Fig. 17B for structures). Where the
enhancement-mode TFTuses the p-type semiconductor 6PTTP6
with the nonpolar PTS SAM (since this SAM induces a slight
negative V
T
shift the device is normally-off), and the
depletion-mode TFT uses 6PTTP6 with the polar uorosilanes
(since these SAMs exhibit a positive V
T
shift they are normally
on). The conclusion is that SAMs can induce a controllable V
T
shift and modulate the current in logic circuits without the need
for external electric eld poling. This represents one of the best
examples of how fundamental insights concerning the SAM-
semiconductor interaction enable technological advances.
An alternative surface functionalization approach has recently
been demonstrated by Podzorov and co-workers,
[264]
where
organosilane SAMs deposited on the surface of single crystal
rubrene (Fig. 1) induce a pronounced increase of the surface
conductivity s 10
5
S square
1
, compared to 10
8
10
7
S
square
1
typically observed in OTFTs.
[258]
In this work, top gate
OTFTs were fabricated on the SAM-functionalized single crystals
Figure 16. F8BT polymer semiconductor TFTs with various organosilane
SAMs on 200 nm SiO
2
as the gate dielectric, or with polyethylene on SiO
2
as a buffer. Reprinted with permission from [210]. Copyright 2005,
Macmillan Publishers Ltd.
Figure 15. A) Cartoon depicting the difference between ordered (left) and
disordered (right) ODTS monolayers. B) X-ray diffraction patterns of the 50
nm-thick P5 lms deposited on ordered (top) and disordered (bottom)
ODTS monolayers at 30, 60, and 90 8C. The inset shows the enlarged thin
lm (002) and bulk (002)
0
phase reections of the XRD pattern. Reprinted
with permission from [259]. Copyright 2008, American Chemical Society.
1420 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
using parylene as the gate insulator and
observed ms of 12 cm
2
V
1
s
1
. Although
these mobilities are lower than the highest
reported,
[265]
and the details about SAM growth
and surface binding remain unclear, the
enhanced conductivity observed with SAM
functionalization is promising for top gate
OTFT applications utilizing the SAM as the
gate insulator (instead of as a surface treatment)
between the single crystal channel beneath and
the deposited gate electrode on top.
4.9. SAMs and SAMTs As Gate Insulators
The notion of high capacitance SAMs and their
use as the gate dielectric in OTFTs (Table 2) was
pioneered by Vuillaume et al. in 1996 where it
was established that SAMs of n-ODTS grafted
on Si native oxide were good insulators
(J $10
8
A cm
2
at 5.8 MV cm
1
and break-
down at 912 MVcm
1
) despite a thicknesses of
only 2.8nm.
[38]
The authors demonstrated that
the interface state density (D) of the OTS SAMs
can be reduced by 1 order of magnitude (to
2 10
11
cm
2
eV
1
) by annealing the lm at
350 8C.
[266]
They also investigated the electrical
properties of alkyl monolayers versus varying
chain lengths (1.92.6 nm) and found sup-
pressed leakage current densities and low
conductivities ($4.610
15
S cm
1
) for well-
ordered SAMs, and much larger conductivities
in deliberately disordered SAMs.
[182]
One of the
rst attempts utilizing SAMs of alkylsilanes
with a COOH end group as the gate insulator
in a p-type (sexithiophene, 6T) OTFT was
demonstrated with respectable performance
metrics (m$3.6 10
4
cm
2
V
1
s
1
, I
on
:I
off
ratio $10
4
, and V
T
1.3 V).
[267]
However, TFTs
with an aromatic terminated alkylsilane,
18-phenoxyoctadecyltrichlorosilane (PhO-OTS),
Table 2. Summary of the capacitance C [nF cm
2
], dielectric constant e
eff
, and OTFT characteristics (gate voltage V
G
[V], m [cm
2
V
1
s
1
], current on/off
ratio I
on
/I
off
, threshold voltage V
T
[V], and subthreshold slope SS [mV decade
1
]) for various SAM and SAMT gate dielectrics.
Reference SAM C Semicond V
G
m (I
on
/I
off
) V
T
(SS)
[265]
SiCl
3
(CH
2
)
12
COOH 6T 2 3.6 10
4
(10
4
) 1.3 (350)
[39]
PhO-OTS 500 P5 2.5 1.0 (10
6
) 1.3 (100)
[37]
SAND I 400 6T 1 0.04 (8 10
2
) 0.03
SAND II 710 6T 1 0.02 (7 10
2
) 0.08
SAND III 390 6T 1 0.06 (10
3
) 0.06
[268]
v-SAND 1 400 P5 2 1.9 ($10
5
) 1.0
v-SAND 2 400 P5 2 3.4 ($10
5
) 1.0
[271]
Phenylundecanoic acid Al
2
O
3
(6 nm) P5 60 0.35 9
[273]
7-OTS/Ti SAMT P5 1 1.3 (500) 0.48
[285]
ODPAAl
2
O
3
(3.6 nm) 700 P5 3 0.6 (10
7
) (100)
[272]
ODPAHfO
2
(3.1 nm) 580 P5 1.5 0.15 (10
5
) 0.53(130)
sp-PA1 HfO
2
(3.1 nm) 690 P5 1.5 0.22 (10
6
) 0.41(110)
sp-PA2 HfO
2
(3.1 nm) 640 P5 1.5 0.15 (10
6
) 0.41(100)
Figure 17. A) Chemical structures of SAM precursor molecules as the SiO
2
gate dielectric
coating in OTFTs (top). Pentacene (middle) and C
60
(bottom) TFT transfer characteristics.
Dielectrics are: bare SiO
2
(untreated), CH
3
terminated SAM/SiO
2
, NH
2
terminated SAM/SiO
2
,
and uorinated SAM/SiO
2
. Adapated with permission from [258]. Copyright 2004, Macmillan
Publishers Ltd. B) Chemical structures of the SAM molecular precursors and the semiconductor
(top). Illustration of a unipolar inverter (middle) and gains of the PMOS-like inverter circuit
(bottom, f 10 mHz). The arrows indicate the hysteresis direction of the inverter output; the
supply voltage V
DD
is 50 V. Reprinted with permission from [263]. Copyright 2007, American
Chemical Society.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1421
R
E
V
I
E
W
www.advmat.de
SAM gate dielectric were investigated by a different group, and
drastically enhanced P5 TFT performance was reported.
[39]
Operating at 2 V, the devices exhibited m1 cm
2
V
1
s
1
, a very
high on/off ratio of $10
6
, V
T
$1.3 V, and a subthreshold slope of
100 mV decade
1
. The OTS-OPh SAM is tightly-packed due to
intra-SAMpp interactions and exhibits very low leakage currents
($8 10
7
A cm
2
at 2.5 V), large breakdown elds and
capacitances of $14 MV cm
1
and 900 nF cm
2
, respectively.
The Northwestern group investigated SAMTs (i.e., SANDtypes
I, II, and III described in Section 2.4.1, Fig. 4) as gate dielectric
materials. SANDs were established as excellent insulators via
solution-phase cyclic voltammetry and MIS leakage current
measurements (current densities in the range of 10
8
10
5
Acm
2
),
where measured breakdown elds for IIII were $57 MVcm
1
.
[37]
Capacitancevoltage (CV) measurements on MIS structures
reveal maximum capacitances C
i
400 (I); 710 (II); 390 (III) nF
cm
2
at 10
2
Hz. It was found that annealing at 120180 8C
reduces CV hysteresis width to 0.1 V and reduces frequency-
dependent CV dispersion, suggesting that pristine IIII
contain quantities of xed positive charge densities (Q
f
),
2 10
12
5 10
12
cm
2
.
[266,268]
Interface state densities (D)
calculated from capacitancevoltage (CV) plots were 3 10
12
eV
1
cm
2
, and it was found that annealing (120180 8C) reduces
Q
f
and D to 10
11
cm
2
and 10
11
eV
1
cm
2
, respectively. Various
oligothiophene semiconductors and n-type CuPc TFTs were
fabricated with SANDs as the gate dielectrics, and comparable
mobilities were measured to those obtained on 300 nm SiO
2
, but
at much lower operating voltages for the devices with SAND gate
dielectrics. It was noted that TFTperformance could be improved
by patterning the gate electrode, or by incorporating higher-k
molecules in the SAND gate dielectric.
To enhance the TFT performance of SAND gated TFTs, a new
self-assembly procedure for the fabrication of multilayer SAND-
like gate dielectrics via room temperature vapor phase deposition
(v-SAND, Fig. 18) was demonstrated.
[269]
In addition, the
structures of the new molecular constituents (1 and 2) are
improved over the original molecular component (Stb) of SANDs
by (i) their ability to self-assemble via head-to-tail intra-molecular
hydrogen-bonds, and (ii) anticipated larger molecular polariz-
abilities than Stb. Here, the trends in dielectric permittivities are
evaluated qualitatively using the ClausiusMossotti relation
e /(3 2aN)/(3 aN) (where a is the polarizability along the
conjugated long axis of the molecule and N is the molecular
density in cm
3
) in conjunction with sum-over-states
[270]
calculated molecular polarizabilities. Uniform lms (rms rough-
ness<1 nm) on native SiO
2
were measured to be 3.5 nm thick by
X-ray reectivity.
[271]
Additionally, a process for the room
temperature, vapor-phase deposition of an SiO
x
Cap layer (similar
to step (ii) Fig. 4) was developed. Here, the capping layer thickness
on the SAMs is 5.5 nmfor a 20 minexposure to Si
2
OCl
6
vapor. Low
leakage current densities $10
6
A cm
2
and large capacitances
(400 nF cm
2
) at 2 V were measured in MIS devices. Back
calculation of the molecular k using the parallel plate model for
three capacitors in series reveals high-ks of 11 and 9 for 1 and 2,
respectively. Interestingly, P5 TFTs fabricated with v-SANDs have
very large mobilities (23 cm
2
V
1
s
1
) and I
on
:I
off
ratios ($10
5
),
achieving the goal of optimizing TFT performance. Since the
dielectric-semiconductor interface is essentially the same for
solution processed SAND and v-SAND based P5 TFTs, the
performance enhancement can be rationalized by: (i) increased
polarizability of the molecular layers 1 and 2 versus the Stb
molecular layer used in SAND, and (ii) favorable P5 growth
morphology on top of v-SANDs, which exhibits large grains of the
P5 in the thin lm phase
[35]
by XRD characterization.
Zuppiroli et. al. claimed that vapor deposition leads to better
electrical performance than solution-based methods.
[272]
In this
example of vapor-phase gate dielectric OTFTs, fabrication begins
with an Al
2
O
3
adhesion layer (6 nm) deposited on top of the SiO
2
substrate, followed by deposition of either anthracene- 9-carboxy
acid (anthracene-COOH), or phenylundecanoic acid (see Fig. 19A
for molecular structures), and only 10 nm of P5 for the active
channel. Enhanced P5 mobilities were observed in devices with
phenylundecanoic acid gate dielectrics compared to bare SiO
2
gate dielectrics, although both P5 lms had similar grain sizes.
Since it is known that silanol groups give rise to surface traps in
SiO
2
dielectrics, the authors claimed that charge transfer fromthe
P5 to trap sites of the bare SiO
2
surface generates residual charge
carriers at the semiconductor-dielectric interface. Four-probe
conductivity measurements were used to quantify the density of
these residual charge carriers, r s
2D
/(me), where e is the charge
of the electron. For SiO
2
(with the most surface defects),
r 1.0 10
13
cm
2
, for the anthracene-COOH SAM
r 1.3 10
12
cm
2
, and for the phenylundecanoic acid SAM,
the charge density is the smallest r 1.4 10
11
cm
2
.
Figure 18. A) Molecular structures of highly polarizable components 1 and
2 and SiO
x
Cap materials used in v-SAND fabrication. B) OTFTs device
characteristics for 1 (top) and 2 (bottom). Reprinted with permission
from [269]. Copyright 2008, American Chemical Society.
1422 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
Recently, hybrid inorganic/phosphoric acid SAMs were used
as gate dielectrics in low-voltage OTFTs.
[273]
In this contribution,
phosphonic acid (PA) SAMs deposited on SiO
2
(1.7 nm)/HfO
2
(3.1 nm thick) substrates were used as the gate dielectric in P5
TFTs. Phosphoric acid SAMs were used because of their better
stability to moisture and less homocondenstation during device
fabrication compared to their silane-based counterparts. The
SAM molecular structures investigated in this study (Fig. 19B)
combine the advantages of phosphoric acid SAMs discussed
previously (such as PhO-OTS, and anthracene-COOH) by
incorporating a long insulating carbon chain with a polarizable
anthracene terminal group in both (2-anthryl)undecoxycarbonyl-
decylphosphonic acid (ps-PA1) and (2-anthryl)undecoxycarbonyl-
undecylphosphonic acid (ps-PA2). The hybrid HfO
2
/PA-SAMs
exhibit large capacitances of 690 nF cm
2
(ps-PA1) and 640 nF
cm
2
(ps-PA2) compared to SAMs of n-octadecylphosphonic acid
(ODPA, 580 nF cm
2
), and low leakage current densities of
10
9
10
8
A cm
2
at 2 V, allowing for a P5 TFToperating voltage
of only 1.5 V. Improvements in the P5 TFTperformance such as
larger mobilities, enhanced I
on
:I
off
ratios, and lower subthreshold
slopes ($100 mV decade
1
) were observed for the ps-PA-based
OTFTs compared to the bare HfO
2
control devices, and were
explained by a combination of surface energy and chemical
functionality at the P5/dielectric interface.
In another example of hybrid organi-
c-inorganic gate dielectrics, rapid vapor-phase
fabrication techniques were used to build
superlattices of alkene-terminated SAMs with
TiOH interlayers.
[274]
The hybrid SAMs were
formed by exposing TiO
2
-coated Si substrates
to 7-octenyltrichlorosilane (7- OTS) in the
presence of H
2
O vapor at temperatures of
100 and 20 8C. The terminal SAM vinyl groups
were converted to carboxylic groups with ozone
treatment in the ALD (atomic layer deposition)
growth chamber. SAMTs were then built up by
generating an active titanium hydroxide layer
on the COOH-terminated SAM by vapor phase
titanium isopropoxide adsorption, followed by
an exchange reaction with water. SAMT lms
were fabricated by repetition of these three
steps (Fig. 19C). Leakage current density versus
voltage characteristics of the Pt/multilayer
SAM/Pt capacitors were measured for SAM
thicknesses from 1.199 nm (Fig. 19C). P5
TFTs using the 100 nm-thick multilayer SAMs
(further treated with a hydrophobic SAM as a
surface treatment), exhibit m1.3cm
2
V
1
s
1
,
at operating voltage of only 1 V (due to high
dielectric constant of the SAM17), I
on
:I
off
current ratio >500, and V
T
0.48 V. Further-
more, the authors demonstrated two-terminal
electrical bistable devices, which can be
switched on and off reproducibly more than
10
6
times due to charge lling and delling of
the defects in the TiO
2
layer or interfacial
layers.
Similarly, the effects of multilayers of SAND
type III (named III-n, where n indicates the
number of repeating SAND III units, typically n 3, and
d $16 nm) as the gate dielectric in TFTs have been intensely
investigated. SAND robustness studies were begun by rst
demonstrating the compatibility of III-3 with carbon nanotube
semiconductors for low voltage TFTs.
[275]
Here the SAND
multilayers ($16 nm thick) were deposited conventionally from
the solution phase, and have a measured capacitance of 170 nF
cm
2
. Single-wall carbon nanotubes (SWCNTs) were grown by
CVD onto SiO
2
/Si wafers and then transfer printed directly onto
the SAND dielectric. The resulting lm has $10 tubes mm
2
which act as the semiconducting lm in the TFTs. Good
SWNT-SAND adhesion allows direct photolithographic pattern-
ing of the source and drain electrodes by liftoff. TFTperformance
is greatly improved over control devices using 100 nm SiO
2
gate
dielectrics as indicated by substantially lowered hysteresis and V
T
shifts. Since the SiO
2
control and SAND have very similar surface
properties (such as the number of xed charges, interface state
densities, and surface chemistry), the enhanced properties of the
TFTs were attributed principally to reduced operating voltage,
which avoids charge injection traps arising from adsorbed water
near/on the SWCNTs at high voltages. TFT performance is
excellent with m$5.6 cm
2
V
1
s
1
(in the linear regime),
V
T
0.2 V, and a low gate leakage current of $10 nA at
V
G
1 V. In addition, compatibility with n-type SWCNTs (by
Figure 19. A) Molecular structures used in SAMs for gate dielectrics: COOH-anthracene (left)
phenylundecanoic acid (right) used in [272]. B) Molecular structures used in SAMs for gate
dielectrics: ps-PA1 (right) ps-PA2 (left) used in [273]. C) Schematic of the molecular layer
deposition procedure to build SAMTs (top) and electrical characteristics of SAMTs (bottom). JV
plots of different thickness SAMTs in MIM devices (left) and pentacene TFTs (right). Reprinted
with permission from [274]. Copyright 2007, American Chemical Society.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1423
R
E
V
I
E
W
www.advmat.de
PEI coating) was demonstrated by the small observed hysteresis
and measured TFT properties: m4.1 cm
2
V
1
s
1
and
V
T
0.2 V for applications in low voltage complementary logic
gates. Signicant gains ($8) were measured, which were the
highest at the time compared to 100 nm SiO
2
gate dielectrics.
[275]
High-performance single carbon nanotube eld effect tran-
sistors with a thin gate dielectric based on a SAMgate dielectric of
PhO-OTS
[39]
(2 nm thick) were demonstrated by Klauk et al.
[276]
This is the rst systematic study of the stability of a SAM to
various e-beam exposures. The authors nd negligible change in
leakage current density compared to evaporated Au metal
contacts when using an electron dose of 300 mC cm
2
. However,
leakage current increases to >10
6
A cm
2
at doses larger than
1800 mC cm
2
. Operating under ambient conditions, the TFTs
exhibited large transconductance (20 mS), small hysteresis, and a
low subthreshold slope (60 mVdecade
1
) from which the authors
estimate a low interface state density (D6 10
11
cm
2
V
1
)
compared to other SWCNT transistors. These results demon-
strate the important point that organic SAMs can withstand
processing conditions typically used in the fabrication of
inorganic electronics, which suggests that integration with
inorganic semiconductor technologies is an
attractive option to overcome challenges of
printing electronics.
As further examples, the compatibility of the
SAND gate dielectrics with an inorganic
semiconductor (In
2
O
3
) was demonstrated in
fully transparent, high-performance TFTs.
[277]
In
2
O
3
is a wide-bandgap (3.63.75 eV), n-type
semiconductor (m
single crystal
160 cm
2
V
1
s
1
), with high transparency in the visible
region (>90%). Thin lms of In
2
O
3
(60 nm on
III-3, and 120 nm on 300 nm SiO
2
) were
deposited at room temperature by ion-assisted
deposition (IAD) directly on top of the dielectric
(note that SAND is stable to the in situ ion/
plasma exposure during In
2
O
3
growth). The
conductivities of the thin lms were measured
to be 10
4
10
5
S cm
2
, and XRD revealed
substantial crystallinity on both SiO
2
and
SAND growth surfaces. Signicant In
2
O
3
TFT performance enhancement is observed
with SAND-gated devices, where m140 cm
2
V
1
s
1
, interfacial trap density D10
11
cm
2
,
V
T
0.0 V with nearly hysteretic free response,
on/off 10
5
, and the subthreshold slope
150 mV decade
1
, compared to the perfor-
mance on SiO
2
-gated devices, where m
10 cm
2
V
1
s
1
, on/off 10
5
. To correct for
the performance differences arising from the
different dielectric layers, the TFT transfer
currents versus charge carrier density are
evaluated. It is found that SAND-based devices
turn on at much lower accumulated charge
carrier densities than SiO
2
, indicating greater
charge injection efciency in the former than
the latter. To realize fully transparent TFTs, the
same fabrication procedures were followed
except utilizing glass/ITO as the bottom gate
electrode and In
2
O
3
source and drain electrodes. The perfor-
mance of SAND-based transparent TFTs is the same as on n

-Si
substrates with an improved subthreshold slope 90 mV
decade
1
. These materials combinations advance the eld toward
the realization of plastic electronic displays featuring optical
transparency, mechanical ruggedness, environmental stability,
and inexpensive/large area fabrication.
In further work, it was demonstrated that the power
consumption efciency (which is the main technical challenge
in the development of truly viable exible displays) of single ZnO
NW (nanowire)-based TFTs is enhanced by the use of SAND type
III-3 (16 nm) as the gate dielectric (Fig. 20A).
[278]
ZnO nanowires
(80 nm average diameter, and 5 mm average length) were
purchased from Nanolab Inc. The NWs were dispersed in
2-propanol and the dispersions were transferred to the SAND-
coated Si substrates. Source and drain Al electrodes were
deposited by electron beam evaporation and patterned by
photolithography and lift-off. The SAND dielectrics were rst
electrically characterized in MIS devices (Al/SAND/Si), and a
leakage current density $10
8
A cm
2
and a capacitance of
180 nF cm
2
at 1 V were measured, verifying the compatibility
Figure 20. A) Field-emission SEM image of a 130 nm diameter ZnO-NW TFT (the scale bar is
2 mm) (top), and TFT output characteristics for a SAND-based ZnO-NW TFT (middle), and TFT
using 70 nm SiO
2
as the gate dielectric (bottom). Reprinted with permission from [278].
Copyright 2005, American Chemical Society. B) TFT transfer characteristics for varying proton
radiation doses of ZnO-NW TFTs with SiO
2
(top) and SAND gate dielectrics (bottom). Reprinted
with permission from [279]. Copyright 2006, American Institute of Physics.
1424 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
of SAND with photolithography and e-beam evaporation
methodologies. SAND-gated ZnO
2
NW-TFTs showed reduced
operating voltages from 2.5 V to <1.5 V, while maintaining the
device on/off ratio and increasing the on-current to 2 mA (from
0.3 mA on SiO
2
). The V
T
of SAND-gated ZnO NW-FETs was rst
observed to be 0.4 V (possibly due to the native SiO
2
and
inherent xed charge in the SAND), and the subthreshold slope
was 400 mVdecade
1
, suggesting some surface traps. Therefore,
SAND-based ZnO NW-TFTs were treated with shielded ozone
exposures, and the subthreshold slope was reduced to 150 mV
decade
1
. In addition, reduced V
T
(0.2 V) and improved I
on
:I
off
ratios (10
8
) were observed.
[278]
This performance enhancement is
attributed to the NWdoping density, and an improved NW-SAND
interface. The calculated mobility (taking into account the
cylindrical geometry of the channel), is 196 cm
2
V
1
s
1
, which
is much greater than 818 cm
2
V
1
s
1
measured for ZnO NWon
thick SiO
2
dielectrics, and 54cm
2
V
1
s
1
on the control SiO
2
(70 nm thick). The mobility varies from 164181 cm
2
V
1
s
1
with varying NW diameter and length.
To examine the radiation stability of SANDs, proton radiation
(10 MeV H

) hardness testing of these single ZnO-NW SAND


gated TFTs was conducted.
[279]
The radiation tolerance was
investigated for space-based applications, and after various dosing
and exposure conditions, the V
T
of the TFTs did not shift
signicantly (Fig. 20B). SAND gated TFTs recently went to the
International Space Station on the Endeavor space shuttle for
further radiation testing.
[280]
These results suggest that the bulk
oxide trap density and interface trap density formed in SAND (or at
the SAND-ZnO NW interface) during H

irradiation are sig-


nicantly lower than traditional SiO
2
gate dielectrics, thus pro-
mpting the interface studies of ZnO-NWTFTs using low-frequency
noise and temperature-dependent IV measurements.
[281]
The technique of low-frequency noise measurements was used
to quantify the surface/interface states in the ZnO-NW TFTs. For
SAND gated ZnO-NWTs, the current noise amplitude increases
with increasing V
SD
bias (Fig. 21A).
[281]
The authors quantify the
interface trap densities by applying Hooges empirical model of 1/
f noise behavior to the SAND-based devices and to the SiO
2
-based
control devices. Lower 1/f noise constants are found for
SAND-based devices, and the authors conclude that the interface
trap densities are comparable to those for the aforementioned
SWCNT devices (D$10
12
cm
2
V
1
) by comparison of Hooges
constants. Larger temperature variation of the transfer curves,
and larger threshold voltage shifts (DV
T
) versus temperature are
observed for the SiO
2
/ZnO-NW TFTs compared to the SAND
gated devices, and provide further evidence that the SAND/
ZnO-NW TFTs have low interface trap and defect densities. Also,
temperature-dependent leakage current densities were collected
over 25100 8C (Fig. 20B), and the data analyzed in three regions:
(i) thermal generation, (ii) subthreshold region, and (iii) on-state
region. For SAND/ZnO NWTs, the maximum activation energy
of 0.36 eVis measured at V
G
<0.65 V, which is just below the V
T
.
It is found that increasing the positive bias above the V
T
decreases
the barrier to electron injection, which corresponds to the smaller
activation energies measured above V
T
(0.050.1 eV) and the role
of the contact resistance on the TFT performance.
From the foregoing results it is clear that SANDs are robust
and compatible with a wide range of inorganic semiconductors
(Table 3). However, the successful implementation of any organic
dielectric material with a solution-processed inorganic semi-
conductor layer in a TFT had not been achieved until recently,
when it was demonstrated that SAND-based TFTs could be
fabricated with solution-processed cadmium selenide (CdSe)
Figure 21. A) Current noise spectrumof a SAND-based ZnO-NW TFT, and
B) temperature-dependent transfer plots for SiO
2
/ZnO-NW TFT (left) and
SAND/ZnO-NW TFT (right). Reprinted with permission from [281].
Copyright 2008, American Institute of Physics.
Table 3. Summary of the capacitance C [nF cm
2
] and TFT characteristics (gate voltage V
G
[V] and source/drain voltage V
SD
[V], m [cm
2
V
1
s
1
], current on/
off ratio I
on
/I
off
, threshold voltage V
T
[V], and subthreshold slope SS [mVdecade
1
]) for various SAMT gate dielectrics with inorganic semiconductors (ISC).
Reference SAM C (e
eff
) ISC V
G
(V
SD
) m (I
on
/I
off
) V
T
(SS)
[275]
PhO-OTS SiO
2
(4 nm) 500 (3.3) SWCNT 1 (1) 5.6 (10
5
) 0.2 (60)
[274]
SAND III-3 170 p-SWCNT 1 (1.5) 5.6 0.2
SAND III-3 170 n-SWCNT (PEI coating) 1 (1) 4.1 0.2
[276]
SAND III-3 180 In
2
O
3
1 (1) 140 (10
5
) 0.33 (150)
[277]
SAND III-3 180 (3) ZnO NW 1.5 (0.1) $196 ($10
8
) 0.2 (150)
[216]
SAND III-3 (400 8C annealed) 274 CdSe 4 (2) 41 (10
5
) 2.5 (260)
[287]
SAND III-4 In
2
O
3
NW 3 (0.1) 258 (10
5
) 0.1 (250)
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1425
R
E
V
I
E
W
www.advmat.de
as the semiconducting layer.
[217]
The chemical
bath deposition method (employing Cd
2
and
sodium selenosulfate, SeSO
2
3
solutions) was
used to deposit the CdSe lms on SAND III-3
as the gate dielectric. MIS and TFT character-
istics were compared as a function of anneal-
ing. It was found that for unannealed, and for
300 8C and 400 8C annealed SAND samples,
the leakage current density remains low
$1.1 mA cm
2
at 4 V, but the capacitance
increases with increasing annealing tempera-
tures [160 nF cm
2
(unannealed) <225 nF
cm
2
(300 8C) <274 nF cm
2
(400 8C)], which
is most likely due to sharper sp organic/Cap
interfaces and lm reconstructive morphology
changes (e.g., thermal condensation of residual
SiOH groups). The maximum performance
(I
on
:I
off
ratio 10
6
, V
T
3.0 V, and subthres-
hold slope 0.26 V decade
1
) was found for
CdSe SAND-based TFTs annealed at 400 8C.
Considering the results of alkoxysilane anneal-
ing
[209]
and SiO

trapping
[210]
mentioned
previously, this result may not be surprising
given that there are multiple siloxane Cap
layers in the total 16 nm thickness of SAND
III-3. In general, it was found that SAND-based
CdSe TFTs have larger mobilities (m15 and
m41 cm
2
V
1
s
1
for 300 and 400 8C
annealing, respectively) than SiO
2
-based
devices (m1.3 and m4.0 cm
2
V
1
s
1
for
300 and 400 8C annealing, respectively). The thermal robustness
and solution-phase semiconductor deposition compatibility of
SANDs make the scope of applications in a wide variety of organic
and inorganic electronics much broader.
Thus far the focus has been on the compatibility of SANDs
with unconventional semiconductors. However, SANDs are
functional on a wide range of substrates. To this end, junction
eld effect performance enhancement and surface passivation
using SANDs (I and III) was compared to that achieved by
1-octadecanethiol (ODT) on GaAs (Fig. 22).
[282,283]
This problem
is signicant because it is known
[284]
that the GaAs surface states
pin the Fermi level near mid-gap for the unpassivated surface,
and Schottky barrier heights of 0.8 eVare commonly observed for
metal contacts to the GaAs surface, regardless of metal work
function.
[285]
Thus, successful control of semi-
conductor interface state density represents a
major advance in the understanding of the less
well understood GaAs surface. Although the
GaAs/I and GaAs/III interfaces are identical
(due to the Alk layer), SAND type III shows
better passivation (lower subthreshold slope).
The positive V
T
shifts observed for ODT and
SAND I are attributed to the substrate cleaning
process. Even though the same cleaning
procedure is used for SAND III, the deposition
of additional layers (especially the highly
polarizable Stb layer) reduces the surface state
density and induces a negative V
T
shift (Fig.
22B). This effect is modeled in 2-D device
simulations by incorporating negative xed charges
(Q
F
1.85 10
12
cm
2
) to the t data. These charges are
attributed to the negative iodide counter ions in the p-conjugated
Stb layer, which create a strong local electrical eld oriented
toward the underlying GaAs surface, indicating that the ionic
charges present in SAND III are essential for the enhanced GaAs
junction eld effect transistor performance.
In related work, Klauk et al. demonstrated low operating
voltage complementary organic circuits with SAM gate dielectric
OTFTs.
[286]
The ODPA SAM was rst characterized in MIS
devices to have capacitances of 0.7 mF cm
2
and leakage current
densities of J <5 10
6
A cm
2
(at 3 V). ODPA-based P5 TFTs
exhibited m0.4 cm
2
V
1
s
1
, I
on
:I
off
current ratio 10
7
, and a
subthreshold swing of 100 mV decade
1
. Furthermore, the
Figure 22. A) GaAs/SAND JFET device structure, and B) transfer characteristics utilizing the
indicated gate dielectrics. Reprinted with permission from [282]. Copyright 2008, American
Institute of Physics.
Figure 23. A) SAM molecular structure and device fabrication schematic. B) Complementary
circuits with ODPA based TFTs (for the lowest leakage current deposition procedure). Reprinted
with permission from [286]. Copyright 2008, American Chemical Society.
1426 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
authors demonstrated compatibility of the SAM with an n-type
organic semiconductor hexadecauorocopperphthalocyanine
(F
16
CuPc), and measured performance metrics similar to those
reported in the literature. Using these SAMs with P5 and F
16
CuPc
organic semiconductors, low voltage (1.53 V) complementary
circuits and ring oscillators were demonstrated with a static
power consumption of less than 1 nW per logic gate. Expanding
on this work, Halik and co-workers next demonstrated micro-
contact-printed ODPA SAMs as the gate dielectric for TFTs and
complementary circuits on AlO
x
/Al substrates (Fig. 23).
[287]
They
found that TFTs and complementary circuits fabricated by
printing and subsequent wet etching perform identically to the
TFTs and ring oscillators reported previously where the gate
electrodes were patterned by shadow-masking. Nonetheless, this is
the rst demonstration of a printing process that uses the direct
printing of a SAM layer, and subsequently uses that layer as an etch
resist to pattern the bottom gate electrodes.
The insight from the above examples of SAM- and
SAMT-based TFTs has enabled one of the most unconventional
hybrid organic inorganic electrical technologies reported to date:
transparent active matrix organic light-emitting diode (AMOLED)
displays powered by nanowire electronics.
[288]
This is the rst
example of transparent AMOLED display elements composed of
54 176 mm OLED pixels, in which the switching and driving
circuits are comprised exclusively of nanowire transistor (NWT)
electronics fabricated at room temperature. Proof of concept
green-emitting polymer LEDs with interfacial charge-blocking
materials were integrated with a transparent bottom contact
electrode. The circuit for a unit pixel consists of one switching
NWT, two parallel driving NWTs, and one storage capacitor. The
device is composed of multiple layers making up the transistor
and the OLED (Fig. 24). Briey, the fabrication begins with
200 nm thick SiO
2
layer deposited by e-beam evaporation on a
glass substrate. Next, 100 nm of ITO is deposited by IAD at room
temperature, and patterned by photolithography. Next, a 22 nm
SANDIII multilayer (III-4) is deposited on the patterned ITOgate
electrodes via solution self-assembly. Contact holes are patterned
as anode openings for the OLED units and as bottom gate electrode
contacts for the pixel. Next, a suspension of single crystal In
2
O
3
nanowires is dispersed on the substrate. ITOsource/drain electrodes
are deposited by IADand patterned by lift-off. Next, 200nmof SiO
2
is
deposited for PLED fabrication and characterization. The mobilities
of the present SAND-based In
2
O
3
TFTs (m$258cm
2
V
1
s
1
) are
similar to In
2
O
3
NWs on oxide dielectrics (m$6.9279.1cm
2
V
1
s
1
) andbulk single crystals (m$160cm
2
V
1
s
1
). Minor hysteresis,
large I
on
currents (610
6
A at V
G
2.0 V), and low subthreshold
slope (0.35 V decade
1
) indicate negligible charge trapping and
detrapping in/on the SAND and at the NW-SAND interface. Fully
transparent, proof-of-concept 2 mm2 mmNW-AMOLED arrays
(300 pixels 900 NWTs) were fabricated using a very thin Al
cathode on glass substrates. The optical transmission values are
$72% (before OLED deposition- limited by the thin Al cathode)
and $35% (after OLED deposition) in the 3501350 nm
wavelength range, which corresponds to a green peak lumines-
cence of >300 cd m
2
. Note that transmission coefcients up to
70% have been reported for OLED structures on plastic
substrates.
[289]
This fabrication process involves fewer mask
steps than conventional Si approaches and is very promising for
future AMOLED displays.
4.4. SAMs as the TFT Semiconductor Channel
As described in Section 4.2, several groups have studied the
thickness dependence of mobility for various semiconductors,
and for P5 and CuPc the mobilities begin to saturate after 6
monolayers of semiconductor deposition.
[290,291]
This phenom-
Figure 24. A) Illustration of SAND AMOLED with SAND based In
2
O
3
-NW
TFTs. B) Typical transfer characteristics at different V
SD
(0.1, 0.2, 0.5, and
1.0 V), and the inset shows the device hysteresis at V
SD
0.1 V. C)
Measured luminance versus supply voltage curves for a 2 mm 2 mm
AMOLED array with all scan and data lines enabled, demonstrating light
emission through the Al cathode (circles) and through the glass substrate
(squares). The bottom inset is an optical image of NW-AMOLED consist-
ing of three 2 mm 2 mm AMOLED pixel arrays, 340 unit pixels, 80 tran-
sistor/circuit test devices, 6 alignment marks, 20 test patterns, and contact
pads. Reprinted with permission from [288]. Copyright 2008, American
Chemical Society.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1427
R
E
V
I
E
W
www.advmat.de
enon opens the possibility of using SAMs and SAMTs as the
active channel in TFTs.
[292294]
In 2005, Pilkuhn and co-workers
studied ODTS SAMs with different end group functionalizations,
such as methyl, thiol, thiophene, phenoxy, and biphenyl.
[292]
Of
particular interest were the good insulating properties of the alkyl
chain (large breakdown eld E
B
16 MV cm
1
), and the
simultaneous lateral (in-plane) conductance of the I
2
-doped
biphenyl end group, suggesting these SAMs could be used as
SAM-TFTs. In the same year, Malliaras and co-workers studied
the dependence of P5 m with lm thickness, and found that the
mobility saturates at $0.45 cm
2
V
1
s
1
after 6 monolayers.
[295]
In
2007, Horowitz and co-workers demonstrated with rather com-
plicated device fabrication (e-beam lithography of short channel
lengths<1 mm), that SAMs of bifunctional silanes on SiO
2
(Fig. 25A) could be used as the semiconductor layer in TFTs.
[296]
However, for sensing applications mobility saturation starting
at only 12 monolayers is desirable.
[297]
In 2005 Dinelli et al.
studied the hole mobility in ultra thin lms of 6Tas a function of
n

-Si/SiO
2
/HMDS surface coverage.
[298]
It was observed that the
charge carrier mobility saturates after 2 monolayers at 0.015 or
0.043 cm
2
V
1
s
1
(depending on 6T deposition rate), indicating
that the rst two molecular layers next to the dielectric interface
dominate the charge transport, whereas the upper semiconductor
layers do not actively contribute. Capitalizing on the rapid
mobility saturation of the 6PTTP6 semiconductor, Katz et al.
demonstrated a TFT with only 1.4 6PTTP6 monolayers, which
exhibited a mobility of 0.024 cm
2
V
1
s
1
(within a factor of 3 of
the bulk mobility). The authors investigated the sensing ability of
their thin lm semiconductor TFTs by fabricating a co-evaporated
blended lm of 6PTTP6 with a functionalized
hydroxy bithiophene (HO6OPT), where the
surface hydroxyl groups of HO6OPT serve as
receptor sites for the device. TFTs were tested
upon exposure to 5 ppm of dimethyl-
methylphosphonate (DMMP) vapors, and it
was found that the mobility of the device
decreased signicantly when the DMMP vapor
was injected into the chamber. However, the
TFTmobility could be recovered by owing N
2
or by moderate heating of the device (Fig.
25B).
[297]
Very recently de Leeuw and
co-workers reported on SAM-based TFTs using
a-substituted quinquethiophene as the semi-
conducting core (for long range pp coupling in
the monolayer) spaced from the monochlor-
osilane surface anchoring group by an unde-
cane aliphatic chain (Fig. 25C).
[299]
Thermally
grown SiO
2
was used as the gate dielectric, and
the top source/drain electrodes were fabricated
by conventional photolithographic methods.
The SAM-TFTs exhibited highly reproducible
bulk-like hole mobilities (0.04 cm
2
V
1
s
1
,
40 mm channel length), and large current
modulation. Furthermore, the authors fabri-
cated SAM-based TFT inverters (Fig. 25C) and
integrated circuits with performance similar to
that of state-of-the-art organic integrated cir-
cuits for radio-frequency identication trans-
ponders.
[299]
While a true SAM-TFT has yet to
be realized, these examples demonstrate signicant advances in
the area of ultra-thin OTFTs.
5. Conclusions
It is well established that OTFT semiconductor properties and
consequently OTFTdevice performance are strongly linked to the
electrical properties of the gate dielectric. However, the gate
dielectric does more than just affect the semiconductor growth
morphology. SAMs with different molecular dipole moments on
conductive electrodes induce surface potential changes that shift
the work function or the electron afnity of the bottom metal or
semiconductor electrode, respectively. These energy level shifts
inuence the transport/injection characteristics of the SAM,
affecting both the charge accumulation and conduction in the
semiconducting channel of the OTFT.
[59]
While a denite
correlation between SAM surface phenomena and TFT perfor-
mance is not yet established, it is clear that hybrid and SAND gate
dielectrics represent an important advance in both fundamental
science, and in device applications for the fabrication of low
operating voltage unconventional electronic circuity.
[286]
This
constitutes a major step towards exible electronics.
Acknowledgements
We thank ONR (N00014-02-1-0909), AFOSR(FA9550-08-1-0331), and the
NSF-MRSEC program through the Northwestern Materials Research
Figure 25. A) Output characteristic of a typical Si/Al
2
O
3
/Pt/SAM transistor with 20 nm channel
length and 5 mm channel width; and chemical structures used in SAMs. (note carboxylic acids
were not used in the TFTs due to their instability to the e-beam fabrication procedure.). Reprinted
with permission from[295]. Copyright 2005, Wiley-VCHVerlag GmbH. B) The eld-effect mobility
of 6PTTP6 TFTs at various lm thicknesses (left), the inset is the same data on a linear scale. The
change of the mobility of the ultra-thin TFT with receptor layer upon exposure to 5 ppm DMMP
vapor (right), and chemical structures of semiconductors. Reprinted with permission from [296].
Copyright 2007, Wiley-VCH Verlag GmbH. C) Output characteristics of a SAM-TFT having a
channel length and width of 40 and 1000 mm, respectively (left). Static inputoutput charac-
teristics of SAM-TFT based unipolar inverter (right). Where V
in
is the input voltage and V
out
is the
output voltage, V
g
0. The inset shows a diagram of the logic gate and a plot of the measured
gain, and the chemical structure self-assembled semiconductor is shown to the right. Reprinted
with permission from [299]. Copyright 2008, Macmillan Publishers Ltd.
1428 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
Center (DMR-0520513) for support of this research. This article is part of a
Special Issue on Interfaces in Organic Electronics.
Received: November 6, 2008
[1] A. Facchetti, M.-H. Yoon, T. J. Marks, Adv. Mater. 2005, 17, 1705.
[2] T. W. Kelley, P. F. Baude, C. Gerlach, D. E. Ender, D. Muyres, M. A. Haase,
D. E. Vogel, S. D. Theiss, Chem. Mater. 2004, 16, 4413.
[3] S. R. Forrest, Nature 2004, 428, 911.
[4] C. Rolin, K. Vasseur, S. Schols, M. Jouk, G. Duhoux, R. Muller, J. Genoe, P.
Heremans, Appl. Phys. Lett. 2008, 93, 033305.
[5] L. Burgi, M. Turbiez, R. Pfeiffer, F. Bienewald, H.-J. Kirner, C. Winnewisser,
Adv. Mater. 2008, 20, 2217.
[6] A. Babel, Y. Zhu, K.-F. Cheng, W.-C. Chen, S. A. Jenekhe, Adv. Funct. Mater.
2007, 17, 2542.
[7] H. Usta, A. Facchetti, T. J. Marks, J. Am. Chem. Soc. 2008, 130, 8580.
[8] B. Jones, A. Facchetti, M. R. Wasielewski, T. J. Marks, Adv. Funct. Mater.
2008, 18, 1329.
[9] H. Tian, J. Shi, D. Yan, L. Wang, Y. Geng, F. Wang, Adv. Mater. 2006, 18,
2149.
[10] T. Kojima, J.-I. Nishida, S. Tokito, Y. Yamashita, Chem. Lett. 2007, 36,
1198.
[11] M. L. Tang, M. E. Roberts, J. J. Locklin, M. M. Ling, Z. Bao, Chem. Mater.
2006, 18, 6250.
[12] R. Schmidt, S. Goettling, D. Leusser, D. Stalke, A.-M. Krause, F. Wurthner,
J. Mater. Chem. 2006, 16, 3708.
[13] M. Mas-Torrent, C. Rovira, Chem. Soc. Rev. 2008, 37, 827.
[14] H. Yan, Y. Zheng, R. Blache, C. Newman, S. Lu, J. Woerle, A. Facchetti,
Adv. Mater. 2008, 9999, 1.
[15] Z. Wang, C. Kim, A. Facchetti, T. J. Marks, J. Am. Chem. Soc. 2007, 129,
13362.
[16] S. E. Koh, J. E. Mendvedeva, A. Facchetti, B. Delley, A. J. Freeman, T. J.
Marks, M. A. Ratner, J. Phys. Chem. B. 2006, 110, 24361.
[17] B. Yoo, D. Basu, T. Jung, D. Fine, B. A. Jones, A. Facchetti, M. R.
Wasielewski, T. J. Marks, K. Dimmier, A. Dodabalapur, Adv. Mater.
2007, 19, 4028.
[18] J. A. Letizia, M. R. Salata, C. M. Tribout, A. Facchetti, M. A. Ratner, T. J.
Marks, J. Am. Chem. Soc. 2008, 130, 9679.
[19] H. Kong, D. H. Lee, I.-N. Kang, E. Lim, Y. K. Jung, J.-H. Park, T. Ahn, M. H.
yi, C. E. Park, H.-K. Shim, J. Mater. Chem. 2008, 2008. 1895.
[20] a) Lu, H. Usta, C. Risko, L. Wang, A. Facchetti, M. A. Ratner, T. J. Marks,
J. Am. Chem. Soc. 2008, 130, 7670. b) H. Yan, Z. Chen, Y. Zheng, C. E.
Newman, J. Quin, F. Dolz, M. Kastler, A. Facchetti, Nature 2009, 457, 679.
[21] Organic Semiconductors in Sensor Applications, Springer Series in Materials
Science, Vol. 107, (Eds. D. A. Bernards, R. M. Owens, G. G. Malliaras),
Springer, Berlin/Heidelberg 2008.
[22] J. Weidner, in Proc. ECS Trans, Curran Assoc. Inc., Washington DC 2008,
Vol. 11, pp. 89.
[23] S. Allard, M. Forster, B. Souharce, H. Thiem, U. Scherf, Angew. Chem. Int.
Ed. 2008, 47, 4070.
[24] Organic Electronic Materials: Conjugated Polymers and Low Molecular
Weight Organic Solids, (Eds. R. Farchioni, G. Grosso), Springer, Berlin
2001.
[25] G. Malliaras, Organic Semiconductors and Devices, Wiley, Hoboken
2003.
[26] M. Berggren, D. Nilsson, N. D. Robinson, Nat. Mater. 2007, 6, 3.
[27] A. Facchetti, Mater. Today 2007, 10, 28.
[28] A. Facchetti, T. J. Marks, H. E. Katz, J. Veinot, in Printed Organic and
Molecular Electronics, (Eds.: D. R. Gamota, P. Brazis, K. Kalyanasundaram,
J. Zhang), Kluwer Academic, Boston 2004, pp. 83159.
[29] C. R. Newman, C. D. Frisbie, D. A. d. S. Filho, J.-L. Bredas, P. C. Ewbank, K.
R. Mann, Chem. Mater. 2004, 16, 4436.
[30] Printed Organic and Molecular Electronics (Eds.: D. R, Gamota, P. Brazis, K.
Kalyanasundaram, J. Zhang), Kluwer Academic, Boston 2004.
[31] J. Hicks, D. Bergstrom, M. Hattendorf, J. Jopling, J. Maiz, S. Pae, C.
Prasad, J. Wiedemer, Intel Technol. J. 2008, 12, 131.
[32] F.-C. Chen, C.-H. Liao, Appl. Phys. Lett. 2008, 93, 103310.
[33] J. X. Tang, C. S. Lee, M. Y. Chan, S. T. Lee, Appl. Surf. Sci. 2008, 254, 7688.
[34] K. D. Kim, D. S. Kim, C. K. Kim, C. K. Song, Jpn. J. Appl. Phys. 2008, 47,
6496.
[35] C. Kim, A. Facchetti, T. J. Marks, Adv. Mater. 2007, 19, 2561.
[36] C. D. Dimitrakopoulos, P. R. L. Malenfant, Adv. Mater. 2002, 14, 99.
[37] M.-H. Yoon, A. Facchetti, T. J. Marks, Proc. Natl. Acad. Sci. 2005, 102, 4678.
[38] D. Vuillaume, C. Boulas, J. Collet, J. V. Davidovits, F. Rondelez, Appl. Phys.
Lett. 1996, 69, 1646.
[39] M. Halik, H. Klauk, U. Zschieschang, G. Schmid, C. Dehm, M. Schutz, S.
Maisch, F. Effenberger, M. Brunnbauser, F. Stellacci, Nature 2004, 431, 963.
[40] B. Mann, H. Kuhn, J. Appl. Phys. 1971, 42, 4398.
[41] I. Langmuir, Trans. Faraday Soc. 1920, 15, 62.
[42] M. Kitamura, Y. Arakawa, J. Phys.: Condens. Matter 2008, 20, 184011.
[43] C.-A. Di, G. Yu, Y. Liu, D. Zhu, J. Phys. Chem. B 2007, 111, 14083.
[44] J. Zaumseil, H. Sirringhaus, Chem. Rev. 2007, 107, 1296.
[45] G. Horowitz, J. Mater. Res. 2004, 19, 1946.
[46] J. C. Cho, J. L. Lee, Y. He, B.-S. Kim, T. Lodge, C. D. Frisbie, Adv. Mater.
2008, 20, 686.
[47] Dielectric Films for Advanced Microelectronics (Eds: M. Baklanov, M. Green,
K. Maex ), John Wiley & Sons Ltd., New York 2007.
[48] J. Zhao, K. Uosaki, Appl. Phys. Lett. 2003, 83, 2034.
[49] M. Porti, M. Nafria, X. Aymerich, A. Olbrich, B. Ebersberger, J. Appl. Phys.
2002, 91, 2071.
[50] O. Gershewitz, M. Grinstein, C. N. Sukenik, K. Regev, J. Ghabboun, D.
Cahen, J. Phys. Chem. B 2004, 108, 664.
[51] S. Kera, Y. Yabuuchi, H. Yamane, H. Setoyama, K. K. Okudaira, A. Kahn,
N. Ueno, Phys. Rev. B 2004, 70, 085304.
[52] Y.-J. Lui, H.-Z. Yu, Chem. Phys. Chem. 2003, 4, 335.
[53] A. Scott, D. B. Janes, C. Risko, M. A. Ratner, Appl. Phys. Lett. 2007, 91,
033508.
[54] S. Lenfant, C. Krzeminski, C. Delerue, G. Allan, D. Vuillaume, Nano Lett.
2003, 3, 741.
[55] B. Kim, J. M. Beebe, Y. Jun, X.-Y. Zhu, C. D. Frisbie, J. Am. Chem. Soc.
2006, 128, 4970.
[56] K. Demirkan, A. Matthew, C. Weiland, Y. Yao, A. M. Rawlett, J. M. Tour, R.
L. Opila, J. Chem. Phys. 2008, 128, 074705.
[57] S. Lenfant, D. Guerin, F. Tran Van, C. Chevrot, S. Palacin, J. P. Bourgoin,
O. Bouloussa, F. Rondelez, D. Vuillaume, J. Phys. Chem. B. 2006, 110,
13947.
[58] C. Miramond, D. Vuillaume, J. Appl. Phys. 2004, 96, 1529.
[59] N. Peor, R. Sfez, S. Yitzchaik, J. Am. Chem. Soc. 2008, 130, 4158.
[60] H. Vazquez, W. Gao, F. Flores, A. Kahn, Phys. Rev. B Rapid Commun. 2005,
71, 041306.
[61] A. Natan, L. Kronik, H. Haick, R. Tung, Adv. Mater. 2007, 19, 4103.
[62] G. Heimel, L. Romaner, E. Zojer, J.-L. Bredas, Nano Lett. 2007, 7, 932.
[63] D. M. Alloway, M. Hofmann, D. L. Smith, N. E. Gruhn, A. L. Graham, R. J.
Colorado, V. H. Wysocki, R. T. Lee, P. A. Lee, N. R. Armstrong, J. Phys.
Chem. B 2003, 107, 11690.
[64] G. Dholakla, M. Meyyappan, A. Facchetti, T. J. Marks, Nano Lett. 2006, 6,
2447.
[65] J. Veres, S. Ogier, G. Lloyd, Chem. Mater. 2004, 16, 4543.
[66] G. E. Poirier, E. D. Pylant, Science 1996, 272, 1145.
[67] A. Cassaro, R. Mazzarello, R. Rousseau, L. Casalis, A. Verdini, A. Kohl-
meyer, L. Floreano, S. Scandolo, A. Morgante, M. L. Klein, G. Scoles,
Science 2008, 321, 943.
[68] A. Ulman, Chem. Rev. 1996, 96, 1533.
[69] J. Sagiv, J. Am. Chem. Soc. 1980, 10, 92.
[70] D. K. Aswal, S. Lenfant, D. Guerin, J. V. Yakhmi, D. Vuillaume, Anal. Chim.
Acta 2006, 568, 84.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1429
R
E
V
I
E
W
www.advmat.de
[71] A. Ulman, An Introduction to Ultrathin Organic Films from Langmuir-
Blodgett to Self-Assembly, Academic Press, San Diego 1991.
[72] F. Schreiber, J. Phys.: Condens. Matter 2004, 16, R881.
[73] C. Oncins, C. Vericat, F. Sanz, J. Chem. Phys. 2008, 128, 044701.
[74] T. Y. B. Leung, M. C. Gerstenberg, D. J. Lavrich, G. Scoles, F. Schreiber, G.
E. Poirier, Langmuir 2000, 16, 549.
[75] C. Vericat, M. E. Vela, G. A. Benitez, J. A. M. Gago, X. Torrelles, R. C.
Salvarezza, J. Phys.: Condens. Matter 2006, 18, R867.
[76] S.-S. Li, L.-P. Xu, L.-J. Wan, S.-T. Wang, L. Jiang, J. Phys. Chem. B 2006, 110,
1794.
[77] J. Noh, H. Kato, M. Kawai, M. Hara, J. Phys. Chem. B 2006, 110,
2793.
[78] D. S. Everhart, in Handbook of Applied Surface and Colloid Chemistry, (Ed.
K. Holmberg) Vol. 2 Wiley, Chichester 2003, pp. 22116.
[79] M. Twardowski, R. G. Nuzzo, Langmuir 2002, 18, 5529.
[80] D. Syomin, J. Wang, B. E. Koel, Surf. Sci. 2001, 495, L827.
[81] P. C. Rusu, G. Giovannetti, G. Brocks, J. Phys. Chem. C 2007, 111, 14448.
[82] P. E. Laibinis, G. M. Whitesides, D. L. Allara, Y. T. Tao, A. N. Parikh, R. G.
Nuzzo, J. Am. Chem. Soc. 1991, 113, 7152.
[83] A. Shaporenko, A. Ulman, M. Zharnikov, J. Phys. Chem. B 2005, 109,
3898.
[84] E. A. Weiss, R. C. Chiechi, G. K. Kaufman, J. K. Kriebel, Z. Li, M. Duati, M.
A. Rampi, G. M. Whitesides, J. Am. Chem. Soc. 2007, 129, 4336.
[85] E. Hoque, J. A. DeRose, R. Houriet, P. Hoffmann, H. J. Mathieu, Chem.
Mater. 2007, 19, 798.
[86] G. K. Jennings, T.-H. Yong, J. C. Munro, P. E. Laibinis, J. Am. Chem. Soc.
2003, 125, 2950.
[87] S. Vollmer, G. Witte, C. Woell, Langmuir 2001, 17, 7560.
[88] J. J. Stapleton, T. A. Daniel, S. Uppili, O. M. Cabarcos, J. Naciri, R.
Shashidhar, D. L. Allara, Langmuir 2005, 21, 11061.
[89] J. C. Love, D. B. Wolfe, R. Haasch, M. L. Chabinyc, K. E. Paul, G. M.
Whitesides, R. G. Nuzzo, J. Am. Chem. Soc. 2003, 125, 2597.
[90] H. Noguchi, M. Ito, K. Uosaki, Chem. Lett. 2005, 34, 950.
[91] S. Lee, J. Park, R. Ragan, S. H. Kim, Z. Lee, D. K. Lim, D. A. A. Ohlberg, R.
S. Williams, J. Am. Chem. Soc. 2006, 128, 5745.
[92] J. C. Love, L. A. Estroff, J. K. Kriebel, R. G. Nuzzo, G. M. Whitesides, Chem.
Rev. 2005, 105, 1103.
[93] F. Schreiber, A. Eberhardt, T. Y. B. Leung, P. Schwartz, S. M. Wetterer, D. J.
Lavrich, L. Berman, P. Fenter, P. Eisen Berger, G. Scoles, Phys. Rev. B:
Condens. Matter Mater. Phys. 1998, 57, 12476.
[94] C. D. Bain, E. B. Troughton, Y. T. Tao, J. Evall, G. M. Whitesides, R. G.
Nuzzo, J. Am. Chem. Soc. 1989, 111, 321.
[95] Y. Han, K. Uosaki, Electrochim. Acta 2008, 53, 6196.
[96] J. Zhang, Q. Chi, J. Ulstrup, Langmuir 2006, 22, 6203.
[97] V. B. Engelkes, J. M. Beebe, C. D. Frisbie, J. Am. Chem. Soc. 2004, 126,
14287.
[98] X. D. Cui, A. Primak, X. Zarate, J. Tomfohr, O. F. Sankey, A. L. Moore, T. A.
Moore, D. Gust, L. A. Nagahara, S. M. Lindsay, J. Phys. Chem. B 2002, 106,
8609.
[99] B. Xu, N. J. Tao, Science 2003, 301, 1221.
[100] X. Xiao, B. Xu, N. J. Tao, Nano Lett. 2004, 4, 267.
[101] M. Tsutsui, Y. Teramae, S. Kurokawa, A. Sakai, Appl. Phys. Lett. 2006, 89,
163111.
[102] R. B. Pontes, F. D. Novaes, A. Fazzio, A. J. R. Da Silva, J. Am. Chem. Soc.
2006, 128, 8996.
[103] D. J. Wold, R. Haag, M. A. Rampi, C. D. Frisbie, J. Phys. Chem. B 2002, 106,
2813.
[104] J. G. Kushmerick, D. B. Holt, S. K. Pollack, M. A. Ratner, J. C. Yang, T. L.
Schull, J. Naciri, M. H. Moore, R. Shashidar, J. Am. Chem. Soc. 2002, 124,
10654.
[105] Z.-F. Shi, L.-J. Wang, H. Wang, X.-P. Cao, H.-L. Zhange, Org. Lett. 2007, 9,
595.
[106] D. S. Seferos, R. Y. Lai, K. W. Plaxco, G. C. Bazan, Adv. Funct. Mater. 2006,
16, 2387.
[107] R. A. Wassel, C. B. Gorman, Angew. Chem. Int. Ed. 2004, 43, 5120.
[108] J. F. Smalley, S. B. Sachs, C. E. D. Chidsey, S. P. Dudek, H. D. Sikes, S. E.
Creager, C. J. Yu, S. W. Feldberg, M. D. Newton, J. Am. Chem. Soc. 2004,
126, 14620.
[109] A. K. Flatt, Y. Yuxing, M. Francisco, J. M. Tour, J. Org. Chem. 2004, 69,
1752.
[110] L. Patrone, S. Palacin, J. P. Bourgoin, J. Lagoute, T. Zambelli, S. Gauthier,
Chem. Phys. 2002, 281, 325.
[111] D. K. James, J. M. Tour, Top. Curr. Chem. 2005, 257, 33.
[112] F. Svedberg, Y. Alaverdyan, P. Johansson, M. Kaell, J. Phys. Chem. B 2006,
110, 25671.
[113] S. Yasuda, T. Nakamura, M. Matsumoto, H. Shigekawa, J. Am. Chem. Soc.
2003, 125, 16430.
[114] B.-C. Yu, Y. Shirai, J. M. Tour, Tetrahedron 2006, 62, 10303.
[115] J. Henzl, T. Bredow, K. Morgenstern, Chem. Phys. Lett. 2007, 435, 278.
[116] G. Pace, V. Ferri, C. Grave, M. Elbing, C. von Haenisch, M. Zharnikov, M.
Mayor, M. A. Rampi, P. Samori, Proc. Natl. Acad. Sci. 2007, 104, 9937.
[117] A. S. Kumar, T. Ye, T. Takami, B.-C. Yu, A. K. Flatt, J. M. Tour, P. S. Weiss,
Nano Lett. 2008, 8, 1644.
[118] M. L. Chabinyc, X. Chen, R. E. Holmlin, H. Jacobs, H. Skulason, C. D.
Frisbie, V. Mujica, M. A. Ratner, M. A. Rampi, G. M. Whitesides, J. Am.
Chem. Soc. 2002, 124, 11730.
[119] G. K. Ramachandran, J. K. Tomfohr, J. Li, O. F. Sankey, X. Zarate, A.
Primak, Y. Terazono, A. L. Moore, D. Gust, L. A. Nagahara, S. M. Lindsay,
J. Phys. Chem. B 2003, 107, 6162.
[120] H. B. Akkerman, B. de Boer, J. Phys.: Condens. Matter 2008, 20, 013001.
[121] E. Tran, M. Duati, V. Ferri, K. Mullen, M. Zharnikov, G. M. Whitesides, M.
A. Rampi, Adv. Mater. 2006, 18, 1323.
[122] C. Haensch, C. Ott, S. Hoeppener, U. S. Schubert, Langmuir 2008, 24,
10222.
[123] J. Cui, Q. Huang, J. C. G. Veinot, H. Yan, Q. Wang, G. R. Hutchison,
A. G. Richter, G. Evmenenko, P. Dutta, T. J. Marks, Langmuir 2002, 18,
9958.
[124] J. Li, J. Lui, G. A. Evmenenko, P. Dutta, T. J. Marks, Langmuir 2008, 24,
5755.
[125] S. Liu, R. Maoz, G. Schmid, J. Sagiv, Nano Lett. 2002, 2, 1055.
[126] R. Maoz, S. R. Cohen, J. Sagiv, Adv. Mater. 1999, 11, 55.
[127] E. P. Plueddemann, Silane Coupling Agents, 2
nd
Ed., Plenum Press, New
York 1991.
[128] M. E. McGovern, K. M. R. Kallury, M. Thompson, Langmuir 1994, 10,
3607.
[129] D. M. Spori, N. V. Venkatarman, S. G. P. Tosatti, F. Durmaz, N. D.
Spencer, Langmuir 2007, 23, 8053.
[130] J. P. Folkers, C. B. Gorman, P. E. Laibinis, S. Buchholz, G. M. Whitesides,
R. G. Nuzzo, Langmuir 1994, 11, 813.
[131] P. E. Laibinis, J. J. Hickman, M. S. Wrighton, G. M. Whitesides, Science
1989, 245, 845.
[132] N. Tsud, M. Yoshitake, Surf. Sci. 2007, 601, 3060.
[133] C. M. Yam, A. J. Dickie, A. K. Kakkar, Langmuir 2002, 18, 8481.
[134] D. A. Offord, S. B. Sachs, M. S. Ennis, T. A. Eberspacher, J. H. Grifn, C. E.
D. Chidsey, J. P. Collman, J. Am. Chem. Soc. 1998, 120, 4478.
[135] R. E. Schaak, T. E. Mallouk, Chem. Mater. 2000, 12, 2513.
[136] H. Lee, L. J. Kepley, H.-G. Hong, T. E. Mallouk, J. Am. Chem. Soc. 1988, 92,
2311.
[137] M. E. van der Boom, P. Zhu, G. Evmenenko, J. E. Malinsky, L. Wenbin, P.
Dutta, T. J. Marks, Langmuir 2002, 18, 3704.
[138] M. R. Linford, C. E. D. Chidsey, J. Am. Chem. Soc. 1993, 115, 12631.
[139] J. M. Buriak, Chem. Rev. 2002, 102, 1271.
[140] S. F. Bent, Surf. Sci. 2002, 500, 879.
[141] R. Basu, J.-C. Lin, C.-Y. Kim, M. J. Schmitz, J. A. Yoder, J. A. Kellar, M. J.
Bedzyk, M. C. Hersam, Langmuir 2007, 23, 1905.
[142] A. Aviram, M. A. Ratner, Chem. Phys. Lett. 1974, 29, 277.
[143] A. Aviram, M. A. Ratner, Molecular Electronics: Science and Technology,
New York Academy of Sciences, New York 1999.
1430 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
[144] R. L. McCreery, Chem. Mater. 2004, 16, 4477.
[145] A. Salomon, D. Cahen, S. Lindsay, J. Tomfohr, V. B. Engelkes, C. D. Frisbie,
Adv. Mater. 2003, 15, 1881.
[146] A. Hanciuc, R. M. Metzger, A. Gong, C. W. Spangler, J. Am. Chem. Soc.
2007, 129, 8310.
[147] J. Chen, M. A. Reed, A. M. Rawlett, J. M. Tour, Science 1999, 286, 1550.
[148] R. M. Metzger, T. Xu, I. R. Peterson, J. Phys. Chem. B 2001, 105, 7280.
[149] R. P. Kalakodimi, A. M. Nowak, R. L. McCreery, Chem. Mater. 2005, 17,
4939.
[150] J. ODwyer, Theory of Electrical Conduction and Breakdown in Solid Dielec-
trics, Clarendon, Oxford 1973.
[151] J. G. Simmons, J. Phys. D: Appl. Phys. 1971, 4, 613.
[152] S. M. Sze, Physics of Semiconductor Devices, 2nd Edn., John Wiley and
Sons, New York. 1981.
[153] J. J. Davis, N. Wang, A. Morgan, T. Zhang, J. Zhao, Faraday Discuss. 2006,
131, 167.
[154] Y. A. Berlin, A. L. Burin, M. A. Ratner, Chem. Phys. 2002, 275, 61.
[155] D. K. James, J. M. Tour, Chem. Mater. 2004, 16, 4423.
[156] W. Wang, T. Lee, M. A. Reed, Proc. IEEE 2005, 95, 1815.
[157] M. Duati, C. Grave, N. Tcbeborateva, W. Jishan, K. Mullen, A. Shaporenko,
M. Zharnikov, J. K. Kriebel, G. M. Whitesides, M. A. Rampi, Adv. Mater.
2006, 18, 329.
[158] A. L. Burin, M. A. Ratner, in Computational Materials Chemistry: Methods
and Aplications (Eds: L. A. Curtiss, M. S. Gordon ), Kluwer Academic,
Netherlands 2004, p. 308.
[159] J. K. N. Mbindyo, T. E. Mallouk, J. B. Mattzela, I. Kratochvilova, B. Razavi,
T. N. Jackson, T. S. Mayer, J. Am. Chem. Soc. 2002, 124, 4020.
[160] V. Mujica, M. A. Ratner, Chem. Phys. 2001, 264, 365.
[161] B. Ulgut, H. D. Abruna, Chem. Rev. 2008, 108, 2721.
[162] E. Tran, C. Grave, G. M. Whitesides, M. A. Rampi, Electrochim. Acta 2005,
50, 4850.
[163] C. Grave, C. Risko, A. Shaporenko, Y. Wang, C. Nuckolls, M. A. Ratner, M.
A. Rampi, M. Zharnikov, Adv. Funct. Mater. 2007, 17, 3816.
[164] A. Nitzan, M. A. Ratner, Science 2003, 300, 1384.
[165] Y. Selzer, A. Salomon, D. Cahen, J. Phys. Chem. B 2002, 106, 10432.
[166] F.-R. F. Fan, J. Yang, L. Cai, D. W. Price, S. M. Dirk, D. V. Kosykin, Y. Yao,
A. M. Rawlett, J. M. Tour, A. J. Bard, J. Am. Chem. Soc. 2002, 124, 5550.
[167] H. B. Akkerman, P. W. M. Blom, D. M. de Leeuw, B. de Boer, Nature 2006,
441, 69.
[168] N. B. Zhitenev, H. Erbe, Z. Bao, Nanotechnol 2003, 14, 254.
[169] M. Galperin, M. A. Ratner, A. Nitzan, A. Troisi, Science 2008, 319, 1056.
[170] A. Nitzan, Annu. Rev. Phys. Chem. 2001, 52, 681.
[171] W. Wang, L. Takhee, M. A. Reed, Phys. Rev. B 2003, 68, 035416.
[172] S. Banerjee, B. Shen, I. Chen, J. Bohlman, G. Brown, R. Doering, J. Appl.
Phys. 1989, 65, 1140.
[173] M. Lenzinger, E. H. Snow, J. Appl. Phys. 1969, 40, 278.
[174] J. M. Beebe, B. Kim, J. W. Gadzuk, C. D. Frisbie, J. G. Kushmerick, Phys.
Rev. Lett. 2006, 97, 026801.
[175] D. K. Aswal, S. Lenfant, D. Guerin, J. V. Yakhmi, D. Vuillaume, Small 2005,
1, 725.
[176] S. Maisch, F. Effenberger, J. Am. Chem. Soc. 2005, 127, 17315.
[177] T. Shamai, A. Ophir, Y. Selzer, Appl. Phys. Lett. 2007, 91, 102108.
[178] F. Chen, X. Li, J. hihath, Z. Huang, N. Tao, J. Am. Chem. Soc. 2006, 128,
15874.
[179] Y. S. Park, A. C. Whalley, M. Kamenetska, L. L. Steigerwald, M. S.
Hybersten, C. Nuckolls, L. Venkataraman, J. Am. Chem. Soc. 2007,
129, 15768.
[180] G. Wang, T.-W. Kim, Y. H. Jang, T. Lee, J. Phys. Chem. C 2008, 112,
13010.
[181] G. Wang, T.-W. Kim, H. Lee, T. Lee, Phys. Rev. B 2007, 76, 205320.
[182] C. Boulas, J. V. Davidovits, F. Rondelez, D. Vuillaume, Phys. Rev. Lett.
1996, 76, 4797.
[183] Y. Gu, B. Akhremitchev, G. C. Walker, D. H. Waldeck, J. Phys. Chem. B
1999, 103, 5220.
[184] R. E. Holmlin, R. Haag, M. L. Chabinyc, R. F. Ismagilov, A. E. Cohen,
A. Terfort, M. A. Rampi, G. M. Whitesides, J. Am. Chem. Soc. 2001, 123,
5075.
[185] N. I. Cracium, J. Wildeman, P. W. M. Blom, Phys. Rev. Lett. 2008, 100,
056601.
[186] S. Barth, U. Wolf, H. Bassler, P. Muller, H. Riel, H. Vestweber, P. F. Seidler,
W. Reiss, Phys. Rev. B 1999, 60, 8791.
[187] M. A. Lampert, P. Mark, Current Injection in Solids, Academic Press, New
York 1970.
[188] M. Pope, C. E. Swenberg, Electronic Processes in Organic Crystals and
Polymers, 2nd Edn., Oxford University Press, New York 1999.
[189] Y. Selzer, M. A. Cabassi, T. S. Mayer, D. L. Allara, Nanotechnol 2004, 15,
S483.
[190] D. Segal, A. Nitzan, M. A. Ratner, W. B. Davis, J. Phys. Chem. B 2000, 104,
2790.
[191] V. D. Lakhno, V. B. Sultanov, B. M. Pettitt, Chem. Phys. Lett. 2004, 400,
47.
[192] T. Renger, R. A. Marcus, J. Phys. Chem. A 2003, 107, 8404.
[193] M. Bixon, J. Jortner, Chem. Phys. 2002, 281, 393.
[194] D. Segal, A. Nitzan, J. Phys. Chem. B 2000, 104, 3817.
[195] E. G. Petrov, V. May, P. Hanggi, Chem. Phys. 2002, 281, 211.
[196] E. A. Weiss, M. R. Wasielewski, M. A. Ratner, in Topics in Current
Chemistry, Springer, Vol. 257, Heidelber 2005, p. 103.
[197] Y. A. Berlin, F. C. Grozema, L. D. A. Siebbeles, M. A. Ratner, J. Phys. Chem.
C 2008, 112, 10988.
[198] E. A. Weiss, M. J. Tauber, R. F. Kelley, M. J. Ahrens, M. A. Ratner, M. R.
Wasielewski, J. Am. Chem. Soc. 2005, 127, 11842.
[199] R. H. Goldsmith, O. DeLeon, T. M. Wilson, D. Finkelstein-Shapiro, M. A.
Ratner, M. R. Wasielewski, J. Phys. Chem. A 2008, 130, 4708.
[200] F. Anariba, R. L. McCreery, J. Phys. Chem. B 2002, 106, 10355.
[201] R. L. McCreery, J. Wu, R. P. Kalakodimi, Phys. Chem. Chem. Phys. 2006, 8,
2572.
[202] Y. A. Berlin, A. L. Burin, M. A. Ratner, Chem. Phys. 2002, 275, 61.
[203] B. Xu, P. Zhange, X. Li, N. J. Tao, Nano Lett. 2004, 4, 1105.
[204] S. H. Choi, B. Kim, C. D. Frisbie, Science 2008, 320, 1482.
[205] C. Zhou, M. R. Deshpande, M. A. Reed, L. I. Jones, J. M. Tour, Appl. Phys.
Lett. 1997, 71, 611.
[206] J. Chen, L. C. Calvet, M. A. Reed, D. W. Carr, D. S. Grubisha, D. W.
Bennett, Chem. Phys. Lett. 1999, 313, 741.
[207] L. E. Calvet, R. G. Wheeler, M. A. Reed, Appl. Phys. Lett. 2002, 80,
1761.
[208] V. Adamec, J. H. Calderwood, J. Phys. D: Appl. Phys. 1975, 8, 551.
[209] S. Jeong, D. Kim, S. Lee, B.-K. Park, J. Moon, Appl. Phys. Lett. 2006, 89,
092101.
[210] L.-L. Chua, J. Zaumseil, J.-F. Chang, E. C.-W. Ou, P. K.-H. Ho, H.
Sirringhaus, R. H. Friend, Nature 2005, 434, 194.
[211] S. Jeong, D. Kim, J. Moon, J. Phys. Chem. C 2008, 112, 5245.
[212] T. Sekitani, Y. Noguchi, U. Zschieschang, H. Klauk, T. Someya, Proc. Natl.
Acad. Sci. 2008, 105, 4976.
[213] A. Briseno, S. C. B. Mannsfeld, M. M. Ling, S. Liu, R. J. Tseng, C. Reese, M.
E. Roberts, Y. Yang, F. Wudl, Z. Bao, Nature 2006, 444, 913.
[214] A. K. Chauhan, D. K. Aswal, S. P. Koiry, S. K. Gupta, J. V. Yakjmi, C.
Suergers, D. Guerin, S. Lenfant, D. Vuillaume, Appl. Phys. A: Mater. Sci.
Proc. 2008, 90, 581.
[215] N. J. Tao, Phys. Rev. Lett. 1996, 76, 4066.
[216] N. P. Guisinger, M. E. Greene, R. Basu, A. S. Baluch, M. C. Hersam, Nano
Lett. 2004, 4, 55.
[217] P. D. Byrne, A. Facchetti, T. J. Marks, Adv. Mater. 2008, 20, 2319.
[218] B. N. Limketkai, M. A. Baldo, Phys. Rev. B 2005, 71, 085207.
[219] M. A. Baldo, S. R. Forrest, Phys. Rev. B 2001, 64, 085201.
[220] P. Marmont, N. Battaglini, P. Lang, G. Horowitz, J. Hwang, A. Kahn, C.
Amato, P. Calas, Org. Electron. 2008, 9, 419.
[221] H. E. Katz, J. Johnson, A. J. Lovinger, W. Li, J. Am. Chem. Soc. 2000, 122,
7787.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1431
R
E
V
I
E
W
www.advmat.de
[222] M. L. Swiggers, G. Xia, J. D. Slinker, A. A. Gorodetsky, G. G. Malliaras, R. L.
Headrick, B. T. Weslowski, R. N. Shashidar, C. S. Dulcey, Appl. Phys. Lett.
2001, 79, 1300.
[223] I. Kymissis, C. D. Dimitrakopoulos, S. Purushothaman, IEEE Trans.
Electron. Dev. 2001, 48, 1060.
[224] W. Wang, T. Lee, M. A. Reed, J. Phys. Chem. B 2006, 108, 18398.
[225] C. Bock, D. V. Pham, U. Kunze, D. Kafer, G. Witte, C. Woll, J. Appl. Phys.
2006, 100, 114517.
[226] C. D. Zangmeister, S. W. Robey, R. D. van Zee, J. G. Kushmerick, J. Naciri,
J. M. Tour, B. Varughese, B. Xu, J. E. Reutt-Robey, J. Phys. Chem. B 2006,
110, 17138.
[227] J.-P. Hong, A.-Y. Park, S. Lee, J. Kang, N. Shin, D. Y. Yoon, Appl. Phys. Lett.
2008, 92, 143311.
[228] D. J. Gundlach, L. Jia, T. N. Jackson, IEEE Electron. Dev. Lett. 2001, 22,
571.
[229] G. S. Tulevski, Q. Miao, A. Afzali, T. O. Graham, C. R. Kagan, C. Nuckolls,
J. Am. Chem. Soc. 2006, 128, 1788.
[230] S. Gowrisanker, Y. Ai, M. A. Quevedo-Lopez, H. Jia, H. N. Alshareef, E.
Vogel, B. Gnade, Appl. Phys. Lett. 2008, 92, 153305.
[231] C. D. Dimitrakopoulos, D. J. Mascaro, IBM J. Res. Dev. 2001, 45, 11.
[232] M.-H. Yoon, C. Kim, A. Facchetti, T. J. Marks, J. Am. Chem. Soc. 2006, 128,
12851.
[233] F. Todescato, R. Capelli, F. Dinelli, M. Murgia, N. Camaioni, M. Yang, R.
Bozio, M. Muccini, J. Phys. Chem. B 2008, 112, 10130.
[234] Y.-Y. Lin, D. J. Gundlach, S. F. Nelson, T. N. Jackson, IEEE Trans. Electron.
Dev. 1997, 44, 1325.
[235] G. Horowitz, X.-Z. Peng, D. Fichou, F. Garnier, Synth. Met. 1999, 51, 419.
[236] M. A. Loi, E. da Como, F. Dinelli, M. Murgia, R. Zamboni, F. Biscarini, M.
Muccini, Nat. Mater. 2005, 4, 81.
[237] T. W. Kelley, L. D. Boardman, T. D. Dunbar, D. V. Muyres, M. J. Pellerite, T.
P. Smith, J. Phys. Chem. B 2003, 107, 5877.
[238] M. Mottaghi, G. Horowitz, Org. Electron. 2006, 7, 528.
[239] D. Knipp, R. A. Street, A. Volkel, A. Ho, J. Appl. Phys. 2003, 93, 347.
[240] J. B. Koo, S. H. Kim, J. H. Lee, C. H. Ku, S. C. Lim, T. Zyung, Synth. Met.
2006, 156, 99.
[241] A. Salleo, M. L. Chabinyc, M. S. Yang, R. A. Street, Appl. Phys. Lett. 2002,
81, 4383.
[242] D. J. Gundlach, C.-C. S. Kuo, C. D. Sheraw, J. A. Nichols, T. N. Jackson,
Proc. SPIE 2001, 4466, 54.
[243] S. Y. Yang, K. Shin, C. E. Park, Adv. Funct. Mater. 2005, 15, 1806.
[244] D. Kumaki, S. Ando, S. Shimono, Y. Yamashita, T. Umeda, S. Tokito, Appl.
Phys. Lett. 2007, 90, 053506.
[245] Y. Bai, X. Liu, L. Chen, M. A. Khizar-ul-Haq, Khan, W. Q. Zhu, X. Y. Jiang, Z.
L. Zhang, Microelectron. J. 2007, 38, 1185.
[246] L. Li, Q. Tang, H. Li, W. Hu, J. Phys. Chem. B 2008, 112, 10405.
[247] M. Halik, H. Klauk, U. Zschieschang, G. Schmid, S. Ponomarenko, S.
Kirchmeyer, W. Weber, Adv. Mater. 2003, 15, 917.
[248] H. Yang, T. J. Shin, M.-M. Ling, K. Cho, C. Y. Ryu, Z. Bao, J. Am. Chem. Soc.
2005, 127, 11542.
[249] M. Shtein, J. Mapel, J. B. Benziger, S. R. Forrest, Appl. Phys. Lett. 2002, 81,
268.
[250] J. E. McDermott, M. McDowell, I. G. Hill, J. Hwang, A. Kahn, S. L.
Bernasek, J. Schwartz, J. Phys. Chem. A 2007, 111, 12333.
[251] Y. D. Park, J. A. Lim, H. S. Lee, K. Cho, Mater. Today 2007, 10, 46.
[252] I. Yagi, K. Tsukagoshi, Y. Aoyagi, Appl. Phys. Lett. 2005, 86, 103502.
[253] T. Yokoyama, C. B. Park, K. Nagashio, K. Kita, A. Toriumi, Appl. Phys. Exp.
2008, 1, 041801.
[254] C. Kim, A. Facchetti, T. J. Marks, Science 2007, 318, 76.
[255] D. Kumaki, M. Yahiro, Y. Inoue, S. Tokito, Appl. Phys. Lett. 2007, 90,
133511.
[256] J. Veres, S. D. Ogier, S. W. Leeming, D. C. Cupertino, S. M. Khaffaf, Adv.
Funct. Mater. 2003, 13, 199.
[257] B. Stadholder, U. Hass, H. Maresch, A. Haase, Phys. Rev. B 2006, 74,
165302.
[258] S. Kobayashi, T. Nishikawa, T. Takenobu, S. Mori, T. Shimoda, T. Mitani,
H. Shimotani, N. Yoshimoto, S. Ogawa, A. Iwasa, Nat. Mater. 2004, 3,
317.
[259] H. S. Lee, D. H. Kim, J. H. Cho, M. Hwang, Y. Jang, K. Cho, J. Am. Chem.
Soc. 2008, 130, 10556.
[260] T. Miyadera, S. D. Wang, T. Minari, K. Tsukagoshi, Y. Aoyagi, Appl. Phys.
Lett. 2008, 93, 033304.
[261] N. Benson, C. Melzer, R. Schmechel, H. van Seggern, Phys. Stat. Sol.(a)
2008, 205, 475.
[262] R. T. Weitz, K. Amsharov, U. Zschieschang, E. B. Villas, D. K. Goswami, M.
Burghard, H. Dosch, M. Jansen, K. Kern, H. Klauk, J. Am. Chem. Soc. 2008,
130, 4637.
[263] C. Huang, H. E. Katz, J. E. West, Langmuir 2007, 23, 13223.
[264] M. F. Calhoun, J. Sanchez, D. Olaya, M. E. Gershenson, V. Podzorov, Nat.
Mater. 2008, 7, 84.
[265] V. Podzorov, E. Menard, A. Borissov, V. Kiryukhin, J. A. Rogers, M. E.
Gershenson, Phys. Rev. Lett. 2004, 93, 086602.
[266] P. Fontaine, D. Goguenheim, D. Deresmes, D. Vuillaume, M. Garet, F.
Rondelez, Appl. Phys. Lett. 1993, 62, 2256.
[267] J. Collet, O. Tharaud, A. Chapoton, D. Vuillaume, Appl. Phys. Lett. 2000,
76, 1941.
[268] E. K. Evangelou, W. C. Weimer, M. Fanciulli, M. Sethu, W. Cranton, J. Appl.
Phys. 2003, 94, 318.
[269] S. A. DiBenedetto, D. Frattarelli, M. A. Ratner, A. Facchetti, T. J. Marks, J.
Am. Chem. Soc. 2008, 130, 7528.
[270] S. Keinan, M. A. Ratner, T. J. Marks, Chem. Phys. Lett. 2006, 417,
293.
[271] S. A. DiBenedetto, I. Paci, A. Facchetti, T. J. Marks, M. A. Ratner, J. Phys.
Chem. B 2006, 110, 22394.
[272] A. von Muhlenen, M. Castellani, M. Schaer, L. Zuppiroli, Phys. Stat. Sol.(b)
2008, 245, 1170.
[273] O. Acton, G. Ting, H. Ma, J. W. Ka, H.-L. Yip, N. M. Tucker, A. K.-Y. Jen,
Adv. Mater. 2008, 20, 3697.
[274] B. H. Lee, M. K. Ryu, S.-Y. Choi, K.-H. Lee, S. Im, M. M. Sung, J. Am. Chem.
Soc. 2007, 129, 16034.
[275] S.-H. Hur, M.-H. Yoon, A. Gaur, M. Shim, A. Facchetti, T. J. Marks, J. A.
Rogers, J. Am. Chem. Soc. 2005, 127, 13808.
[276] R. T. Weltz, U. Zschieschang, F. Effenberger, H. Klauk, M. Burghard, K.
Kern, Nano Lett. 2007, 7, 22.
[277] L. Wang, M.-H. Yoon, G. Lu, A. Facchetti, T. J. Marks, Nat. Mater. 2006, 5,
893.
[278] S. Ju, K. Lee, D. B. Janes, M.-H. Yoon, A. Facchetti, T. J. Marks, Nano Lett.
2005, 5, 2281.
[279] S. Ju, K. Lee, D. B. Janes, R. C. Dwivedi, H. Abffour-Awuah, R. Wilkins,
M.-H. Yoon, A. Facchetti, T. J. Marks, Appl. Phys. Lett. 2006, 89, 073510.
[280] P. Patel-Predd, IEEE Spectrum 2008, 45, 17.
[281] S. Ju, S. Kim, S. Mohammadi, D. B. Janes, y.-G, Ha, A. Facchetti, T. J.
Marks, Appl. Phys. Lett. 2008, 92, 022104.
[282] K. Lee, G. Lu, A. Facchetti, D. B. Janes, T. J. Marks, Appl. Phys. Lett. 2008,
92, 123509.
[283] H. C. Lin, P. D. Ye, Y. Xuan, A. Facchetti, T. J. Marks, Appl. Phys. Lett. 2006,
89, 142101.
[284] F. Seker, K. Meeker, T. F. Kuech, A. B. Ellis, Chem. Rev. 2000, 100,
2505.
[285] C. A. Mead, W. G. Spitzer, J. Appl. Phys. 1963, 34, 3061.
[286] H. Klauk, U. Zschieschang, J. Paum, M. Halik, Nature 2007, 445,
745.
[287] U. Zschieschang, M. Halik, H. Klauk, Langmuir 2008, 24, 1665.
[288] S. Ju, J. Li, J. Lui, P.-C. Chen, Y.-G. Ha, F. Ishikawa, H. Change, C. Zhou, A.
Facchetti, D. B. Janes, T. J. Marks, Nano Lett. 2008, 8, 997.
[289] T. Uchida, S. Kaneta, M. Ichihara, M. Ohtsuka, T. Otomo, D. R. Marx, Jpn.
J. Appl. Phys. 2005, 44, L282.
[290] J. Gao, J. B. Xu, M. Zhu, D. Ma, J. Phys. D: Appl. Phys. 2007, 40, 5666.
[291] B.-N. Park, S. Seo, P. G. Evans, J. Phys. D: Appl. Phys. 2007, 40, 3506.
1432 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 14071433
R
E
V
I
E
W
www.advmat.de
[292] M. Rittner, M. S. Martin-Gonzalez, A. Flores, H. Schweizer, F. Effenberger,
M. H. Pilkuhn, J. Appl. Phys. 2005, 98, 054312.
[293] J. H. Schon, H. Meng, Z. Bao, Nature 2001, 413, 713.
[294] J. H. Schon, Z. Bao, Appl. Phys. Lett. 2002, 80, 847.
[295] R. Ruiz, A. Papadimitratos, A. C. Mayer, G. C. Malliaras, Adv. Mater. 2005,
17, 1795.
[296] M. Mottaghi, P. Lang, F. Rodriguez, A. Rumyantseva, A. Yassar, G. Horowitz,
S. Lefant, D. Tondelier, D. Vuillaume, Adv. Funct. Mater. 2007, 17, 597.
[297] J. Huang, J. Sun, H. E. Katz, Adv. Mater. 2008, 20, 2567.
[298] F. Dinelli, M. Murgia, P. Levy, M. Cavallini, F. Biscarini, D. M. de Leeuw,
Phys. Rev. Lett. 2004, 92, 116802.
[299] E. C. P. Smits, S. G. J. Mathijssen, P. A. van Hal, S. Setayesh, T. C. T. Geuns,
K. A. H. A. Mutsaers, E. Cantatore, H. J. Wondergem, O. Werzer,
R. Resel, M. Kemerink, S. Kirchmeyer, A. M. Muzafarov, S. A. Ponomar-
enko, B. de Boer, P. W. M. Blom, D. M. de Leeuw, Nature 2008, 455,
956.
Adv. Mater. 2009, 21, 14071433 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1433

You might also like