Sei sulla pagina 1di 6

www.advmat.

de
COMMUNICATION

Comparison of the Mobility–Carrier Density Relation in


Polymer and Single-Crystal Organic Transistors
Employing Vacuum and Liquid Gate Dielectrics
By Yu Xia, Jeong Ho Cho, Jiyoul Lee, P. Paul Ruden, and C. Daniel Frisbie*

Charge mobility is a key figure of merit for many applications of In this communication, we report a systematic comparison of
organic semiconductors. In organic field-effect transistors the field-effect mobilities in polymer thin-film transistors (TFTs)
(OFETs), higher charge mobility leads to larger ON-to-OFF and single-crystal OFETs as a function of tunable gate insulator
current ratios and higher switching speeds, which enhances the dielectric constant and gate-induced charge density. Experimen-
utility of these devices for electronic circuitry.[1] It is generally tally, we follow the approach pioneered by Rogers and Podzorov,
appreciated that the carrier mobility in organic semiconductors in which an organic semiconductor crystal or a polymer
correlates with the degree of structural order, for example, high semiconductor film is laminated to a polydimethylsiloxane
crystallinity often (but not always) favors higher mobilities (PDMS) stamp embossed with source, drain, and gate electrodes
because the intermolecular electronic coupling can be greater in a topographic relief, Figure 1a and b.[7] In earlier reports by
than in amorphous or semicrystalline materials.[2] In addition, Podzorov and others, the vacuum gap between the semiconductor
structural disorder leads to in-gap or ‘‘band-tail’’ states that can and the gate serves as the gate dielectric.[6–8,9] Application of a gate
trap charge. For amorphous or semicrystalline polymer semi- voltage induces highly mobile carriers at the pristine vacuum/
conductors, structural disorder is significant and the charge semiconductor interface.
mobility is consequently a strong function of total carrier Here, rather than simply using air or vacuum as the dielectric,
density.[3] Tanase et al. and Shimotani et al. have shown that the we filled the channel space between the semiconductor and the
carrier mobility in polymer semiconductors can increase by gate with liquids of varying dielectric constants. These
several orders of magnitude as carrier density is increased.[4,5] ‘‘liquid-gap’’ transistors allowed us to measure field-effect
The explanation is that at higher carrier concentrations, mobility in the same crystal or film as a function of tunable
disorder-induced traps are filled to a greater extent, which dielectric constant and charge density. We have found striking
increases the average mobility of the remaining carriers because differences in transport behavior for organic single crystals versus
they sample fewer and shallower traps. polymer semiconductor films using these liquid dielectric
In contrast, Morpurgo and colleagues reported that for OFETs transistors. For single crystals of rubrene, the carrier mobility
based on high-mobility rubrene single crystals, the trend is is not a function of charge density (up to 5  1013 cm2), but
opposite;[6] OFETs with higher gate-dielectric constants, and thus strongly depends on the liquid dielectric constant, in keeping
higher gate-induced charge densities, have lower field-effect with previous results reported by Morpurgo on the effects of
mobilities, which they attributed not to carrier-density effects (at dielectric polarizability.[6] For polymer semiconductors, on the
least for carrier densities <1013 cm2), but to coupling of the other hand, the effect of charge density is overwhelming; there is
charges in the channel with the polarizable dielectric. The a strong increase in charge mobility with increasing carrier
fundamental mobility–charge density relation is apparently very concentration, following a power law.
different for polymer semiconductors compared with small- Our conclusions from these experiments reinforce prior
molecule single crystals. It is expected that, to a first approxima- reports on the dielectric and charge-density dependence of
tion, this is because of the vastly different degrees of static transport in organic single crystals and polymer semiconduc-
disorder in the two systems. tors.[4–6,10–12] The novelty of our work is that we have directly
examined the different behaviors of crystals and polymers using
the same test-bed, so that their inherent differences are made
patently clear. In addition, our use of a highly polarizable ionic
[*] Prof. C. D. Frisbie, Y. Xia, Dr. J. H. Cho, Dr. J. Lee liquid as the gate dielectric has allowed us to measure the
Department of Chemical Engineering and Materials Science field-effect mobility over a four order of magnitude span in the
University of Minnesota carrier density, which is much larger than typically reported.
421 Washington Ave. SE
Minneapolis, MN 55455 (USA) For our initial work, we have employed two liquids as gate
E-mail: frisbie@cems.umn.edu dielectrics: a pure ionic liquid [EMIM][TFSI] (molecular structure
Prof. P. P. Ruden shown in Fig. 1c) and Dow Corning #704 silicone oil. The ionic
Department of Electrical and Computer Engineering liquid is strongly polarizable and has been employed previously in
University of Minnesota so-called electrical double-layer OFETs.[13–16] Application of a
200 Union St. SE
Minneapolis, MN 55455 (USA)
negative bias to the gate causes the negative [TFSI] ions in the
ionic liquid to drift to the liquid/semiconductor interface, which
DOI: 10.1002/adma.200803437 induces the accumulation of oppositely charged holes in the

2174 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 2174–2179
www.advmat.de

COMMUNICATION
similar gate-dielectric-dependent transfer char-
acteristics (see Supporting Information).
The ID–VG characteristic of a typical ionic-
liquid-gated rubrene single-crystal OFET is
displayed in Figure 2b. This transistor could
also be operated with very low gate bias.
However, the ON-to-OFF current ratio for the
rubrene OFET is nearly 100 times lower than
the ionic-liquid-gated polymer TFTs with
similar channel dimensions. Conversely, when
gated by vacuum or oil, rubrene single-crystal
transistors operated much better than the
polymer devices. As shown in the inset of
Figure 2b, the vacuum- and oil-gated single-
Figure 1. Device structure: a) cross-section of the polymer TFTs. b) Cross-section of the
crystal transistors could still be turned on at a
single-crystal OFETs. c) Chemical structure of the ionic liquid used in this study. d) Image of
a rubrene single-crystal OFET. A sharp wetting front can be clearly observed when the ionic very low gate voltage, and the channel-output
liquid was introduced into the gap. current was comparable to the current in the
ionic-liquid-gated device. Similar drain-current
levels were also observed for pentacene
single-crystal OFETs gated by the three differ-
semiconductor. An electrical double layer is thus formed ent dielectrics (see Supporting Information).
consisting of large negative ions and positively charged holes It is evident from the transfer characteristics in Figure 2a that
separated across the interface by a distance of the order of the slopes of the ID–VG curves for the polymer transistors
1 nm.[17] The capacitance associated with this double layer is continually increased with negatively increasing gate voltages for
extremely large. We have measured the specific capacitance for both liquid dielectrics and the vacuum gap, while in single-crystal
the [EMIM][TFSI] liquid sandwiched between metal plates to be rubrene OFETs (Fig. 2b) the slope dID/dVG was essentially
15 mF cm2, and this value is essentially independent of thickness constant beyond the threshold voltage. This difference in the
(because in an ionic liquid the bulk polarization is zero as the ID–VG curves reflects the different dependences of mobility (m)
double layers screen the field). On the other hand, silicone oil is on gate-induced sheet charge density ( p) for polymer versus
naturally much less polarizable (dielectric constant ¼ 2.9 at low single-crystal semiconductors;[18] understanding the m–p relation
frequency), and behaves like a conventional dielectric. Typical is thus our main focus.
gate capacitances for the OFETs we fabricated using silicone oil While m and p for the vacuum- and oil-gated transistors can
are 5.5  104 mF cm2. We note that when using ionic liquids, be calculated easily from the standard FET equations
there is also the possibility of electrochemical doping instead of ( p ¼ C0 (VGVth)/e, mlin ¼ (LIDlin)/(WpVD), where C0 is the
electrical double-layer charging.[17] This is particularly relevant for measured sheet capacitance of the gate dielectric, L and W are
permeable polymer semiconductors, as we will
discuss later.
Liquid-gated transistors were fabricated with
spun-cast films of regioregular poly-3-
hexylthiophene (P3HT), poly-3,3000 -didodecyl
quarter thiophene (PQT-12), or vapor-grown
single crystals of rubrene or pentacene, as
shown in Figure 1. We compared the current–
voltage results for liquid-gap transistors to
identical devices with a vacuum gap. Figure 2a
shows the transfer characteristics (drain cur-
rent vs. gate voltage ID–VG, with the source
grounded) of a typical ionic-liquid-gated
PQT-12 TFT acquired at a fixed drain bias
(VD ¼ 1 V). The device turned on near
VG ¼ 0 V, and the channel current increased
by nearly four orders of magnitude by
VG ¼ 1 V. However, when gated using sili-
cone oil as the dielectric (see the inset of
Fig. 2a), the PQT-12 TFT required much higher
operating voltages and yielded much lower
channel currents. The vacuum-gated PQT-12 Figure 2. Transfer characteristics of [EMIM][TFSI] gated a) PQT-12 thin film TFT and b) rubrene
device barely turned on over the same voltage single-crystal OFET at 75 mV s1 gate-voltage sweep rate. The insets show the transfer
range. P3HT-based transistors showed very characteristics of a similar PQT-12 and rubrene OFET gated by vacuum and oil, consecutively.

Adv. Mater. 2009, 21, 2174–2179 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2175
www.advmat.de

the channel length and width, respectively, Vth is the threshold


COMMUNICATION

voltage, and IDlin is the current when jVGVthj > jVDj),[19] the
determination of p in ionic-liquid-gated transistors requires extra
caution. It has been demonstrated elsewhere that under relatively
slow gate-voltage sweep rates (for example, 75 mV s1), an
electrochemical doping process can take place in polymer
transistors.[17] The [TFSI] ions in the dielectric can penetrate
into the bulk of the polymer semiconductor and hence induce a
considerably higher injected carrier density (a 3D instead of 2D
channel). Therefore, for the ionic-liquid-gated polymer devices,
we adopted an alternative method to estimate p. Namely, we
measured the gate-displacement current (IDisp) simultaneously
with the ID–VG characteristics of these transistors, and integrated
IDisp with time as[17,20]

R
Q IDisp dVG
p¼ ¼ (1)
eA rV eA

where rV is the constant gate-voltage sweep rate, dVG/dt.


Subsequently, the mobilities of the polymer TFTs were calculated
using m ¼ (L/W)(ID/pVD).
As we illustrate in the Supporting Information, the electro-
chemical doping of [TFSI] ions into single crystals of rubrene
and pentacene does not occur. Therefore, in our ionic-
liquid-gated single-crystal transistors the carrier densities were
calculated from the capacitance as discussed above, and the linear
mobilities were again obtained from mlin ¼ (LIDlin)/(WpVD) with
VD  0.2 V.
In Figure 3a we display the central result of this study, in which
the effective mobility for each type of transistor is plotted versus
the estimated carrier densities. By using two different liquid-gate
dielectrics and the vacuum gap, we were able to examine the
mobility–carrier density relation over carrier densities spanning
four orders of magnitude, from 3  1010 to 4  1014 cm2, for
each semiconductor material. It is apparent that the effective
mobilities m extracted for polymer TFTs show dependence on
the carrier concentration p, which is qualitatively different from
Figure 3. a) Mobility–carrier density relationship for transistors based on
those of the molecular-crystal OFETs. In polymer transistors, the rubrene and pentacene single crystals as well as PQT-12 and P3HT thin
hole mobility increased with increasing carrier density, though films. The background colors represent different types of gate-dielectric
there was a large initial drop (40) in mobility upon switching materials at various ranges of carrier densities. b) Temperature-dependent
from oil to ionic liquid (discussed below). Charge density was carrier mobility in rubrene single-crystal OFETs and P3HT TFTs gated by
varied by over one order of magnitude using the ionic liquid alone either vacuum or an ionic liquid.
(from 2  1013 to 4  1014 cm2), leading to a steep power-law
increase in mobility over this carrier-density range, with the Table 1. Maximum mobility of typical devices (unit: cm2 V1s1)
mobility eventually exceeding 1 cm2 V1s1 at carrier densities of
Gate dielectrics Single crystal OFETs Polymer TFTs
4  1014 cm2. This large mobility is comparable with polymer
semiconductor mobilities previously reported for transistors with Rubrene Pentacene PQT-12 P3HT
electrolyte gates.[14–17,21]
Vacuum [0.18 nF cm2] 13  1 1.5  0.4 0.006  0.003 0.005  0.002
In contrast, single-crystal OFETs had considerably higher
DC 704 oil [0.55 nF cm2] 8.8  1.7 0.6  0.2 0.04  0.01 0.03  0.02
mobilities when gated by vacuum, and the mobilities decreased
[EMIM][TFSI] [15 mF cm2] 0.31  0.06 0.020  0.004 1.7  0.8 1.2  0.7
when the dielectric was changed to oil and ionic liquid. For a
given dielectric, however, the mobilities in single-crystal rubrene
OFETs were not strongly affected by the carrier concentration. For
pentacene single-crystal OFETs, the mobility increased with gate dielectric material is summarized in Table 1, where one can
carrier concentration for vacuum and oil dielectrics, but this easily recognize the opposite trends for polymer TFTs and
effect was small compared to the mobility change accompanying single-crystal OFETs as a function of gate insulator type.
the change of dielectric. We examined more than ten devices for The large changes in mobility evident in Figure 3a for both
each organic semiconductor, and the maximum mobility for each polymer and single-crystal devices prompted us to investigate the

2176 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 2174–2179
www.advmat.de

COMMUNICATION
temperature dependence of transport. The effective mobilities of
vacuum- and ionic-liquid-gated P3HT TFTs, and rubrene
single-crystal OFETs were measured in the temperature range
200–300 K, and the results are plotted in Figure 3b (it was found
that at temperatures less than 220 K the ionic liquid was frozen
and the transistors were difficult to turn on). As shown in the
figure, the mobility in vacuum-gated rubrene single-crystal
transistors increased with decreasing temperature, consistent
with diffusive transport as observed previously by Podzorov
et al.[8] However, Figure 3b also shows that the mobility was
thermally activated for rubrene single-crystal OFETs gated by
ionic liquids. The change in both the magnitude and temperature
dependence of the mobility for rubrene OFETs suggests a
significant change in the transport behavior associated with
changing the dielectric from vacuum to ionic liquid, as discussed
further below. For P3HT transistors, activated mobilities were
observed for devices gated with vacuum and ionic liquid, but
the activation energy (EAct) was much smaller for gating with
the ionic liquid. An Arrhenius plot yielded EAct ¼ 100 meV for the Figure 4. Mobility–drain current relationship for PQT-12 and P3HT thin-
P3HT device with a vacuum gate, and the same procedure films transistors employing vacuum, oil, or ionic liquid as gate dielectric
gave EAct  15 meV for the P3HT gated with ionic liquid. The material. The background colors represent different types of gate-dielectric
materials that enable various ranges of drain current.
large difference in mobilities for rubrene and P3HT is also clearly
evident in Figure 3b.
The collective observations in Figure 3 can be explained in characteristic widths E0  100 meV for their trap DOS distribu-
terms of the role of disorder on the density of states for the tions.
different semiconductor classes, polymer versus single crystal. At relatively low carrier concentrations (<3  1013 cm2),
We focus first on the polymer TFT results. As described above, we the ionic-liquid-gated polymer TFTs had lower mobility compared
were able to extract activation energies for the mobilities in the to identical transistors gated by vacuum or oil, as shown
polymer transistors and found that the activation energies in Figures 3a and 4. This mobility-lowering effect may be
decreased with increasing carrier concentration. These observa- attributable to carrier localization associated with electrochemical
tions are consistent with a continuum of localized states above a doping of the [TFSI] anions into the transistor channel, as
(hole) transport level, Et. High densities of localized states previously addressed by Shimotani et al.[5] At low electro-
associated with disorder are expected in polymers, and holes chemical-doping concentrations, the [TFSI] in the bulk of
populating these localized states have negligible mobility. the polymer can have strong Coulombic interactions with the
Consequently, hole transport involves thermal excitation to states induced holes in the channel, and hence can significantly reduce
at or below Et. As the states above Et are filled with increasing the hole mobility. At higher doping levels, the holes can screen the
charge-carrier density, the quasi-Fermi level, EF, is lowered, and ionic charge, leading to strong increases in the hole mobility.[10]
the activation energy, EAct  EF–Et, decreases. The measured Another possible explanation is that penetration of [TFSI] into
field-effect mobility in this model is proportional to pt/p, where pt the polymer semiconductor increases the structural disorder,
is the concentration of (mobile) charge carriers with energy less thereby introducing more traps, which must be overcome by
than Et. Assuming that the density of states (DOS) varies further increases in carrier density. Though the precise reason for
exponentially with the energy, that is, gðEÞ / expðE=E0 Þ, where the initial decrease in mobility when employing ionic liquids is
E0 is the characteristic distribution width, one readily obtains not clear, it is evident that the overall m–p relation is qualitatively
that the measured mobility increases with increasing carrier the same for polymer films gated either with vacuum, oil, or
density approximately as m / pðE0 =kT1Þ (where it is assumed ionic-liquid dielectrics.
that E0 > kT and that EAct > kT), as shown in Figure 3a. In In rubrene single-crystal OFETs, the effects of disorder are
addition, using that form of the mobility in a standard gradual expected to be much less than in the polymer devices.
channel approximation model for the current versus voltage Consequently, the HOMO (highest occupied molecular orbital)-
characteristics of a transistor, one finds that the effective mobility band edges are reasonably sharp and mobility values at small hole
extracted in the linear regime is proportional to the drain current concentrations are much higher than in the polymer case.[8]
ðE kT Þ=E0
as IDlin0 (full derivation of the m–p and m–ID relationships However, surface-roughness scattering is likely to be important
can be found in the Supporting Information). We plotted the for molecular crystals, which inevitably are not completely
mobility of polymer transistors for each dielectric as a function of smooth on a molecular scale but have terraces.[22] As a result, the
the drain current in Figure 4, where the power-law relationship dependence of the mobility on the carrier concentration for a
between m and ID can be readily observed. This power law given dielectric involves the competing mechanisms of Cou-
relationship supports the view that continuous filling of an lombic screening and surface-roughness scattering. The former
exponential distribution of trap states is responsible for the ensures that potential fluctuations are minimized as the carrier
mobility increase in polymers. The dashed line in Figure 4 also concentration increases, thus increasing the mobility, while the
indicates that P3HT and PQT-12 transistors have similar latter leads to reductions in the mobility as the transverse electric

Adv. Mater. 2009, 21, 2174–2179 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2177
www.advmat.de

field increasingly confines the carriers to the crystal surface. The crystallized by means of horizontal physical vapor transport (PVT) [23].
COMMUNICATION

experimental results indicate that the two conflicting trends The obtained crystals were used as source material for a subsequent
growth, and this process was repeated twice to increase the purity of the
approximately cancel for the rubrene devices. Compared to
crystals. Poly-3-hexylthiophene (P3HT) and Poly-3,3000 -didodecyl quarter
rubrene, pentacene single crystals have relatively higher trap thiophene (PQT-12) were purchased from Rieke Metals, Inc and American
concentrations.[9] When the injected carrier density was low Dye Source, Inc, respectively, and purified by successive Soxhlet extractions
(when the transistor was gated by vacuum), these traps in the with methanol, acetone, and hexane.
pentacene crystal could still play an important role, and resulted Device Fabrication: A PDMS (Sylgard 184, Dow Corning Corp.) stamp
in a carrier density-dependent mobility similar to that observed in with a gap feature and a conductive coating similar to that described
previously was used as the substrate for all devices. The thickness of the
polymer transistors (Fig. 3a). Such dependency was much less
gap is approximately 5 mm. Single-crystal transistors were fabricated by
pronounced when the OFET was gated by oil or ionic liquid, placing the long axis of a single crystal across the PDMS gap. The
because the carrier density in the channel was considerably metal-coated features on the PDMS stamp were used as source, drain, and
higher. gate electrodes, respectively, while the gap served as a gate dielectric layer.
Employing different gate-dielectric materials that are increas- The channel length was varied from 100 to 500 mm, and the channel width
ingly polarizable introduces additional mechanisms that can was generally limited by the width of the crystals. In case of polymer TFTs,
reduce the mobility. Dynamic coupling of the Fröhlich type 1,2-dichlorobenzene was used to dissolve the polymer (3 mg mL1), and
the solution was spin-coated onto a SiO2/Si wafer (thickness around
between charge carriers in the rubrene or pentacene channels 20 nm). Prior to the spin coating, Cr/Au source and drain contacts were
and the polarization-field fluctuations in the dielectric layer will deposited on the SiO2 surface through a life-off process to assure a better
reduce the mobility of the charge carriers.[6,12] For single-crystal metal/organic contact. The channel length was 20 mm and the channel
OFETs, we tentatively associate the observed steps in the effective width was 200 mm. After the polymer was spun-coated and dried in vacuum
mobilities that coincide with changes in the dielectric layer from overnight, the transparent PDMS substrate was stamped, with the gap of
vacuum to oil and from oil to ionic liquid with that mechanism. the PDMS carefully aligned with the channel between source and drain
electrodes. Finally, the whole device was flipped over and placed on a piece
The activated transport we observed in ionic-liquid-gated rubrene
of glass, silver paint/Au wires were used to extend the source and drain
OFETs (Fig. 3b) could also be considered as reflecting this contacts for the sake of measurement, as illustrated in Figure 1a.
coupling, with the activation energy approximately 70 meV. Device Characterization: Electrical characterization of the transistors
Following the analysis illustrated by Morpurgo et al.,[6] we was performed using a Lakeshore TTP4 probe station with Keithley 237 and
obtained the steady-state dielectric constant of [EMIM][TFSI] 6517A electrometers. For each device, the transistor performance gated by
(er  24). The dielectric value further allowed us to estimate the vacuum (106 Torr, 1 Torr ¼ 133.32 Pa) was first measured. Then a drop
of ionic liquid (or oil) was cast on the PDMS stamp through a 1mL
thickness of the electrical double layer at the ionic liquid/
microsyringe under atmosphere. As can be seen in Figure 1d, a probe tip
semiconductor interface (7 Å, see Supporting Information). was applied to direct the liquid droplet to fill the gap between the
In summary, we have developed a simple platform that enables semiconductor and the gate contact. One can clearly observed a wetting
the systematic study of charge-carrier transport in polymer TFTs front propogated through the whole channel within 1 s, which indicates the
and single-crystal OFETs using vacuum, oil, or ionic liquid gate filling of the gap. The excess liquid was carefully removed using Kimwipe.
insulators, with specific capacitances varying over five orders of After that, the liquid-gated transistor was characterized again in vacuum.
magnitude. Using this broad range of capacitance, we have been The capacitance values of different gate dielectrics were measured using a
HP 4192A LF impedance analyzer using a metal–insulator–metal (MIM)
able to examine the transport over a larger range of carrier structure with similar gap features.
concentrations than typically reported. The carrier mobility in
polymer TFTs strongly increases with increasing carrier
concentration, and the quantitative dependence can be accounted
for by an exponential distribution of states. The transport in Acknowledgements
polymers is also thermally activated, but the activation energy
This work was supported primarily by the MRSEC Program of the National
decreases substantially at the high charge densities facilitated by Science Foundation under Award Numbers DMR-0212302 and
the ionic liquid gate dielectric. The charge transport in DMR-0819885. The authors acknowledge B. Boudouris and B. Kim for
single-crystal transistors, on the other hand, exhibits a completely polymer purification and M. Ha for helpful discussion. Y. X. acknowledges
different behavior, with the most striking phenomenon being the additional support through the Sundahl Fellowship at the University of
strong decrease in mobility with increased polarizability of the Minnesota. Supporting Information is available online from Wiley
gate dielectric. Charge density effects on the mobility are less InterScience or from the author.
important in single crystals up to 3  1013 carriers cm2. Received: November 22, 2008
Fundamentally, the very different state of structural order in Revised: January 17, 2009
single crystals versus polymer films leads to a very different Published online: March 9, 2009
mobility–carrier density relations for the two cases.

[1] a) H. Klauk, in: Organic Electronics, Materials, Manufacturing and Appli-


Experimental cations, Wiley-VCH, Weinheim, Germany 2006. b) Z. Bao, J. Locklin, in:
Materials Preparation: The single-component silicone diffusion Organic Field-Effect Transistors, CRC Press, New York 2007.
pump oil (DC 704) was purchased from Dow Corning Corp. 1-ethyl-3- [2] V. Coropceanu, J. Cornil, D. A. Da Silva Filho, Y. Oliver, R. Silbey, J.-L.
methylimidazolium bis(trifluoromethylsulfonyl)imide ([EMIM][TFSI]) was Brédas, Chem. Rev. 2007, 107, 926.
purchased from Solvent Innovation GmbH (Germany). Rubrene and [3] J.-F. Chang, H. Sirringhaus, M. Giles, M. Heeney, I. McCulloch, Phys. Rev. B
pentacene powder were purchased from Sigma-Aldrich Corp. and 2007, 76, 205204.

2178 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2009, 21, 2174–2179
www.advmat.de

COMMUNICATION
[4] C. Tanase, E. J. Meijer, P. W. M. Blom, D. M. de Leeuw, Phys. Rev. Lett. 2003, [15] J. H. Cho, J. Lee, Y. He, B.-S. Kim, T. P. Lodge, C. D. Frisbie, Adv. Mater.
91, 216601. 2008, 20, 3177.
[5] H. Shimotani, G. Diguet, Y. Iwasa, Appl. Phys. Lett. 2005, 86, 022104. [16] J. H. Cho, J. Lee, Y. Xia, B.-S. Kim, Y. He, M. J. Renn, T. P. Lodge, C. D.
[6] I. N. Hulea, S. Fratini, H. Xie, C. L. Mulder, N. N. Iossad, G. Rastelli, S. Frisbie, Nat. Mater. 2008, 7, 900.
Ciuchi, A. F. Morpurgo, Nat. Mater. 2006, 5, 982. [17] J. Lee, L. G. Kaake, J. H. Cho, X.-Y. Zhu, T. P. Lodge, C. D. Frisbie, J. Phys.
[7] E. Menard, V. Podzorov, S.-H. Hur, A. Gaur, M. E. Gershenson, J. A. Rogers, Chem. C 2008, submitted.
Adv. Mater. 2004, 16, 2097. [18] C. D. Dimitrakopoulos, P. R. L. Malenfant, Adv. Mater. 2002, 14, 99.
[8] V. Podzorov, E. Menard, J. A. Rogers, M. E. Gershenson, Phys. Rev. Lett. [19] S. M. Sze, in: Physics of Semiconductor Devices, 2nd edn. John Wiley & Sons,
2005, 95, 226601. New York 1981.
[9] Y. Xia, V. Kalihari, C. D. Frisbie, N. K. Oh, J. A. Rogers, Appl. Phys. Lett. 2007, [20] Y. Xia, J. H. Cho, B. Paulsen, C. D. Frisbie, M. J. Renn, Appl. Phys. Lett. 2009,
90, 162106. 94, 013304.
[10] V. I. Arkhipov, E. V. Emelianova, P. Heremans, H. Bässler, Phys, Rev. B 2005, [21] a) A. S. Dhoot, J. D. Yuen, M. Heeney, I. McCulloch, D. Moses, A. J. Heeger,
72, 235202. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 11834. b) M. J. Panzer, C. D. Frisbie,
[11] H. Houili, J. D. Picon, L. Zuppiroli, M. N. Bussac, J. Appl. Phys. 2006, 100, Adv. Funct. Mater. 2006, 16, 1051. c) M. J. Panzer, C. D. Frisbie, J. Am.
023702. Chem. Soc. 2007, 129, 6599. d) M. J. Panzer, C. D. Frisbie, Adv. Mater. 2008,
[12] S. Fratini, A. F. Morpurgo, S. Ciuchi, Phys. Stat. Sol. (c) 2008, 5, 718. 20, 686.
[13] S. Ono, S. Seki, R. Hirahara, Y. Tominari, J. Takeya, Appl. Phys. Lett. 2008, [22] E. Menard, A. Marchenko, V. Podzorov, M. E. Gershenson, D. Fichou, J. A.
92, 103313. Rogers, Adv. Mater. 2006, 18, 1552.
[14] J. Lee, M. J. Panzer, Y. He, T. P. Lodge, C. D. Frisbie, J. Am. Chem. Soc. 2007, [23] R. A. Laudise, C. Kloc, P. G. Simpkins, T. Siegrist, J. Cryst. Growth 1998, 187,
129, 4532. 449.

Adv. Mater. 2009, 21, 2174–2179 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2179

Potrebbero piacerti anche