Sunteți pe pagina 1din 9

This article has been accepted for inclusion in a future issue of this journal.

Content is final as presented, with the exception of pagination.

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS 1

Organic Photovoltaic Devices Using an Amorphous


Molecular Material With High Hole Drift Mobility,
Tris[4-(2-Thienyl)Phenyl]Amine
Hiroshi Kageyama, Hitoshi Ohishi, Masatake Tanaka, Yutaka Ohmori, Fellow, IEEE, and Yasuhiko Shirota

Abstract—A planar p-n heterojunction organic photovoltaic cent by the use of copper phthalocyanine [5], [6] or poly(3-
(OPV) device using an amorphous molecular material with a high hexylthiophene) [7], [8] as an electron donor, and fullerene
hole drift mobility of 1.1 × 10−2 cm2 /V·s at an electric field of (C60 ) [5], [6] or its derivative, [6,6]-phenyl-C61 -butyric acid
1.0 × 105 V/cm at 293 K, tris[4-(2-thienyl)phenyl]amine (TTPA),
as an electron donor, and C6 0 , as an electron acceptor, indium-tin- methyl ester ([6,6]-PCBM) [7]–[9], as an electron acceptor, in
oxide (ITO)/poly(3,4-ethylenedioxythiophene) doped with poly(4- devices with elaborate structures.
styrene sulfonate) (PEDOT:PSS) (ca.30 nm)/TTPA (30 nm)/C6 0 Polycrystalline materials have mostly been used for OPV de-
(40 nm)/LiF (0.1 nm)/Al (150 nm), exhibited high performance with vices; however, amorphous molecular materials, i.e., small or-
a fill factor of 0.62 and a power conversion efficiency (PCE) of 1.5% ganic molecules that readily form stable amorphous glasses with
under air-mass 1.5G illumination at an intensity of 100 mW/cm2 .
A p-i-n-type OPV device having a mixed interlayer of TTPA and well-defined glass-transition temperatures (Tg s) [10], [11], have
C6 0 , ITO/PEDOT (ca.30 nm)/TTPA (27 nm)/TTPA:C6 0 (1:4 mo- also been receiving growing attention as good candidates for
lar ratio, 20 nm)/C6 0 (23 nm)/LiF (0.1 nm)/Al (100 nm), exhibited OPV device materials. Grain-boundary-free amorphous molec-
higher performance with a PCE of 1.8% under the same irradi- ular materials that permit the formation of smooth, uniform
ation conditions. A bulk p-n heterojunction OPV devices fabri- amorphous thin films by both thermal deposition and solution
cated by spin coating from solution of TTPA and [6,6]-phenyl-
C6 1 -butyric acid methyl ester ([6,6]-PCBM), ITO/PEDOT:PSS methods are attractive from the standpoint of practical applica-
(ca. 30 nm)/TTPA:[6,6]-PCBM (1:4 molar ratio, ca. 73 nm)/LiF tions. As discussed later, PCEs of ∼1.85% have been reported
(0.1 nm)/Al (100 nm), exhibited a PCE of 1.3%. The high perfor- for the OPV devices using amorphous molecular materials.
mance of the present devices is attributed to the high charge-carrier Photocurrent generation in donor-/acceptor-based OPV de-
mobilities of the materials and the relatively high ionization poten- vices has been understood as consisting of the following se-
tial of TTPA.
quential steps [3], [4], [12], [13]:
Index Terms—Amorphous molecular material, bulk p-n het- 1) light absorption to form excitons;
erojunction device, charge transport, organic photovoltaic (OPV) 2) exciton diffusion to the donor/acceptor interface;
device, p-i-n-type device, planar p-n heterojunction device,
tris(oligoarylenyl)amine. 3) exciton dissociation that involves charge transfer between
the donor and the acceptor to generate geminate hole–electron
pairs, and then, charge separation into free holes and electrons
I. INTRODUCTION
in competition with the recombination of hole–electron pairs;
RGANIC photovoltaic (OPV) devices have been receiv-
O ing a great deal of attention as candidates for next-
generation solar cells and photodetectors due to their poten-
4) charge transport;
5) charge collection by electrodes.
The charge separation probability has been understood in
tially low cost, light weight, and the capability of large-area, terms of the Onsager theory [12], [13]. Recent studies have
thin-film devices [1]–[4]. Since it was reported that an OPV de- shown that the step of the charge separation of hole–electron
vice using copper phthalocyanine (CuPc) as an electron donor pairs remaining bound at the donor/acceptor interface greatly
and a perylene pigment as an electron acceptor exhibited a affects the conversion efficiency of OPV devices [12]–[17] and
power conversion efficiency (PCE) of ∼1% under simulated also the separation probability is largely influenced by the mo-
sunlight irradiation [1], conversion efficiencies of OPV devices bilities of charge carriers [17], i.e., a Monte Carlo model of
have recently been significantly improved up to several per- charge-carrier separation and recombination predicts that the
increase in the carrier mobility reduces the recombination pro-
Manuscript received November 11, 2009; revised December 21, 2009. cess of hole–electron pairs, resulting in the facilitation of the
H. Kageyama, H. Ohishi, and M. Tanaka are with the Department of Applied charge separation process, and hence, leading to an increase in
Chemistry, Faculty of Engineering, Osaka University, Suita 565-0871, Japan
(e-mail: kageyama@chem.eng.osaka-u.ac.jp; ohishi@chem.eng.osaka-u.ac.jp; the short-circuit current (JSC ) and the fill factor (FF) [17]. An-
masatake@chem.eng.osaka-u.ac.jp). other simulation model has also shown that for attaining a large
Y. Ohmori is with the Center for Advanced Science and Innovation, Osaka mean carrier distance, which leads to high JSC , carrier mobility
University, Suita 565-0871, Japan (e-mail: ohmori@casi.osaka-u.ac.jp).
Y. Shirota is with the Department of Environmental and Biological Chemistry, should be large [18].
Fukui University of Technology, Fukui 910-8505, Japan (e-mail: shirota@fukui- In the light of the OPV device material requirement of
ut.ac.jp). high charge-carrier mobility [4], [16], [17], we have aimed at
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org. developing high-performance OPV devices using amorphous
Digital Object Identifier 10.1109/JSTQE.2009.2039699 molecular materials with high charge-carrier mobilities. A few

1077-260X/$26.00 © 2010 IEEE


This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

2 IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS

molecular design concepts for high charge-carrier mobility have was 15–32 µm. The cell was mounted in a cryostat and irradiated
been obtained from the studies of the correlation between molec- with pulsed N2 laser light (wavelength: 337 nm; pulse duration:
ular structures and charge-carrier drift mobilities, which in- 3 ns).
clude the reduction of both molecular size and conformational Current density–voltage characteristics in the dark and
change caused by the bond rotation and the avoidance of the under air-mass (AM) 1.5G illumination (Yamashita Denso,
incorporation of polar groups [3], [10], [11], [19]–[21]. We USS-40) at room temperature were recorded using a Keithley
have found that several amorphous molecular materials of the 2400 source meter. Light intensity was measured using a power
tris(oligoarylenyl)amine family show high charge-carrier drift meter (NEOARK, PM-341). The intensity of the monochro-
mobilities and reported hole drift mobilities of several mate- matic light was measured using a calibrated Si photodiode
rials of this family, including tris[4-(2-thienyl)phenyl]amine (Hamamatsu Photonics, S1336-18BQ). The spectral mismatch
(TTPA), in a previous communication [22]. We have also factors for the devices were 1.06, as determined by the method
demonstrated that the amorphous molecular materials of the described in the literature [25].
tris(oligoarylenyl)amine family function as good candidates for
electron donors in bilayer p-n heterojunction OPV devices [23]. C. OPV Device Fabrication
This paper presents a full account of charge transport in the
TTPA amorphous glass, and the fabrication and performance of Planar p-n heterojunction OPV devices were fabricated as
a planar p-n heterojunction OPV device, including the exper- follows: an ITO-coated glass substrate was cleaned by suc-
imental details. In addition, this paper reports on a p-i-n-type cessive washing with neutral detergent, deionized water, THF,
OPV device fabricated by thermal deposition of TTPA and C60 , and trichloroethene in an ultrasonic bath, followed by exposure
and bulk p-n heterojunction OPV devices fabricated by spin to trichloroethene vapor. Finally, the substrate was irradiated
coating from the solution of TTPA as an electron donor and with ultraviolet light. Poly(3,4-ethylenedioxythiophene) doped
[6,6]-PCBM as an electron acceptor, and discusses the perfor- with poly(4-styrene sulfonate) (PEDOT:PSS) (BAYTRON P
mance of these devices. AI4038) was spin-coated onto the ITO-coated glass substrate.
TTPA amorphous film was prepared by thermal deposition onto
the PEDOT:PSS layer at ∼2 × 10−4 Pa at a deposition rate
of ∼0.1 nm/s at room temperature. Then C60 was vacuum-
deposited onto the TTPA film at ∼2 × 10−4 Pa at a deposition
rate of 0.1–0.2 nm/s at room temperature. The mixed inter-
layer was prepared by codeposition of TTPA and C60 . Finally,
lithium fluoride (0.1 nm) and aluminum (100–150 nm) were
thermally deposited onto C60 at deposition rates of ca. 0.01 and
ca. 0.2–0.5 nm/s, respectively. Bulk p-n heterojunction OPV
devices were fabricated using the film of a mixture of TTPA
and [6,6]-PCBM prepared by spin coating at 2000 r/min from
II. EXPERIMENTAL its chlorobenzene solution with a concentration of 3.5 wt% onto
the PEDOT:PSS-coated ITO glass substrate. All the fabricated
A. Materials
devices were sealed using glass plates with epoxy resin in an
TTPA was synthesized by the Grignard coupling re- argon-filled glove box. Atomic force microscopy images were
action of thiophen-2-ylmagnesium bromide with tris(4- taken using a scanning probe microscope (JEOL, JSPM-5200).
iodophenyl)amine in tetrahydrofuran (THF). It forms an amor-
phous glass with a glass-transition temperature of 70 ◦ C when
III. RESULTS AND DISCUSSION
the melted sample was cooled [24].
C60 and [6,6]-PCBM were purchased from Tokyo Chemical A. Properties of TTPA
Industry and Luminescence Technology Corporation, respec- TTPA undergoes reversible anodic oxidation, showing one
tively, and used without further purification. anodic and the corresponding cathodic waves in its cyclic
voltammogram (see Fig. 1). The ratio of the cathodic peak cur-
B. Measurements
rent (ip c ) to the anodic peak current (ipa ), ip c /ipa , was 0.966,
Cyclic voltammetry was carried out using Ag/AgNO3 as calculated according to the method described in the litera-
(0.01 mol/dm3 ) as a reference electrode in CH2 Cl2 . The half- ture [26]. E1/2 of TTPA, which was determined as the midpoint
wave oxidation potential (E1/2 ) versus ferrocene/ferrocenium of Epa and Ep c , namely, (Epa + Ep c )/2, where Epa and Ep c
(Fc/Fc+ ) was calibrated using E1/2 = 0.20 V versus Ag/AgNO3 are the potentials at which the anodic and cathodic currents,
(0.01 mol/dm3 ) for ferrocene in CH2 Cl2 . respectively, reach the maximum values, was 0.36 V versus fer-
Charge-carrier drift mobility was measured by a time-of-flight rocene/ferrocenium (Fc/Fc+ ) reference electrode [0.20 V ver-
(TOF) method for the TTPA single-layer device. An appropriate sus Ag/AgNO3 (0.01 mol/dm3 )] in dichloromethane (CH2 Cl2 ).
amount of TTPA was heated to melt on an indium-tin-oxide The half-wave oxidation potential of TTPA is 0.17 V less
(ITO) coated glass, which was pressed by another ITO glass, than the half-peak oxidation potential (Ep/2 , the potential at
and then cooled on standing in air. The thickness of the sample which the anodic current is half of the anodic peak current) of
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KAGEYAMA et al.: ORGANIC PHOTOVOLTAIC DEVICES USING AN AMORPHOUS MOLECULAR MATERIAL 3

Fig. 2. Typical transient photocurrent observed for the molecular glass of


Fig. 1. Cyclic voltammogram of TTPA. Solvent: CH2 Cl2 ; sweep rate : TTPA. Thickness of the sample: 22 µm; applied voltage: 220 V; temperature:
100 mV/s; supporting electrolyte: n-Bu4 NClO4 (0.1 mol/dm3 ). 293 K.

TABLE I
ELECTRONIC ABSORPTION BAND MAXIMA (λm a x ), MOLAR EXTINCTION
COEFFICIENTS (ε), HALF-WAVE OXIDATION POTENTIAL (E1 / 2 ), AND HOMO bility of TTPA was determined to be 1.1 × 10−2 cm2 /V·s at
AND LUMO ENERGY LEVELS OF TTPA an electric field of 1.0 × 105 V/cm and at 293 K. This value
is of the highest level among those reported for organic disor-
dered systems, which is comparable to the field-effect mobility
of polycrystalline CuPc [28].
The temperature and electric-field dependencies of the hole
drift mobilities were analyzed in terms of the disorder formal-
ism, which assumes that charge transport in disordered sys-
tems takes place by hopping through a manifold of states with
Gaussian-type density of states (DOS), subject to fluctuations
in both hopping site energies and intermolecular wave function
triphenylamine (TPA) (0.53 V versus Fc/Fc+ in CH2 Cl2 ). Thus, overlap, i.e., energetic disorder and positional disorder, respec-
TTPA has a stronger electron-donating property than TPA. The tively. According to the disorder formalism, the electric-field
highest occupied molecular orbital (HOMO) energy level of and temperature dependencies of charge-carrier drift mobilities
TTPA was estimated to be −5.60 eV on the basis of the oxida- are given by (1) [29], [30], where σ and Σ are the parameters that
tion potential and the solid-state ionization potential (Ip ) of characterize the energetic and positional disorder, respectively,
4,4 ,4 -tris[3-methylphenyl(phenyl)amino]triphenylamine (m- µ0 represents a hypothetical mobility in the energetic disorder-
MTDATA) (E1/2 = −0.14 V versus Fc/Fc+ in CH2 Cl2 and free system, E is the electric field, k is the Boltzmann constant,
Ip = 5.1 eV, as determined by ultraviolet photoemission spec- T is the temperature, and C is an empirical constant
troscopy in air [27]). The lowest unoccupied molecular orbital   2   

(LUMO) energy level of TTPA was estimated to be −2.50 eV 2σ σ 2


from its HOMO energy level and the optical bandgap deter- µ = µ0 exp − exp C −Σ E 2 1/2
.
3kT kT
mined from its long-wavelength absorption edge. The optical
(1)
bandgap of TTPA was determined to be 3.10 eV, which is ca.
In the experimental setup reported previously, a xenon stro-
0.40 eV smaller than that of TPA (3.50 eV). Thus, the LUMO
boscopic lamp with pulsewidths of 1–4 µs was used [24]. Since
level of TTPA is ca. 0.23 eV lower than that of TPA.
hole drift mobilities greater than 10−3 cm2 /V·s could not be
Table I summarizes the electronic absorption band maxima
measured, temperature dependence of hole drift mobilities was
(λm ax ), molar extinction coefficients (ε) at the band maxima,
examined at low temperature below 193 K. In the present TOF
the half-wave oxidation potential (E1/2 ), and the HOMO and
experimental setup, a pulsed N2 laser with a pulsewidth of 3 ns
LUMO energy levels of TTPA.
was employed as the light source; this enabled the measure-
ment of drift mobilities greater than 10−3 cm2 /V·s in the higher
B. Charge Transport in TTPA
temperature region.
The TOF method was used for the measurement of hole drift Fig. 3 shows the electric-field dependence of hole drift mo-
mobilities. As Fig. 2 shows, the molecular glass of TTPA ex- bilities in the glassy state of TTPA at a temperature region wider
hibited nondispersive transient photocurrents. The transit time (up to 293 K) than that of the previous temperature setup (up to
(tτ ) was determined from the cusp of the double logarithm plots 193 K). The logarithm of the hole drift mobilities was propor-
of the transient photocurrent. The hole drift mobility was cal- tional to the square root of the electric field. Fig. 4 shows the
culated from the formula µ = L2 /tτ V, where L is the thickness temperature dependence of the hole drift mobilities at the zero
of the sample and V is the applied voltage. The hole drift mo- electric field, µ(E = 0), which were obtained by the extrapolation
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

4 IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS

Fig. 5. Plots of S versus (σ/kT)2 . Here, S = ∂ ln(µ/µ 0 )/∂E 1 / 2 .


Fig. 3. Electric-field dependence of the hole drift mobilities of TTPA molec-
ular glass.

higher than that for the packet of charge carriers that reached
the thermodynamically equilibrium state.
The values of µ0 and σ in the thermodynamically equilibrium
state were obtained from the intersect and the slope, respectively,
of the linear dependence of the logarithm of µ(E = 0) versus 1/T2
in the temperature region above Tc in Fig. 4. Fig. 5 shows the
plot of the slope of the electric-field dependence of the hole drift
mobility (S = ∂ ln(µ/µ0 )/∂E1/2 ) versus (σ/kT)2 . The positional
disorder Σ was determined from the intersect (S = 0) under the
condition (σ/kT)2 = Σ 2 . The hole-transport parameters in terms
of the disorder formalism for the amorphous glass of TTPA
were determined as follows: µ0 = 1.1 × 10−1 cm2 /V·s, σ =
0.064 eV, Σ = 1.1, and C = 3.3 × 10−4 (cm/V)1/2 .
The results show that while the values of Σ and C for the
molecular glass of TTPA are more or less similar to those for
other amorphous molecular materials, the preexponential factor
µ0 for the TTPA glass is approximately one order of magnitude
Fig. 4. Temperature dependence of the hole drift mobilities of TTPA molecular
glass at zero electric field.
greater than for a number of other amorphous molecular materi-
als [3]. In addition, the value of σ for the TTPA molecular glass
is less than that for other amorphous molecular materials [3]. It
of the electric-field dependence of the hole drift mobilities to is, therefore, indicated that both the large preexponential factor
the zero electric field at each temperature. A linear dependence µ0 and the relatively small energetic disorder σ are responsible
of the logarithm of µ(E = 0) versus 1/T2 was observed in the for the high hole drift mobility of the TTPA molecular glass.
temperature region of ca. 293–195 K, a deviation from the lin- The preexponential factor µ0 , namely, the hypothetical mo-
ear dependence being observed in the temperature region below bility in the absence of energetic disorder, is determined by the
ca. 195 K. This kind of discontinuity has been reported as aris- attempt frequency and the degree of intermolecular wavefunc-
ing from the nondispersive-to-dispersive charge-transport tran- tion overlap [32]. The X-ray crystal structure analysis data of
sition [31], i.e., in the temperature region above this transition the dihedral angles of 6.5◦ , 20.2◦ , and 26.4◦ between the phenyl
temperature (Tc ), a packet of charge carriers, which was initially rings and the thiophene ring for TTPA [24] suggest that the
energetically random, relaxes to the thermodynamically equi- three thiophene-containing oligoarylenyl groups in TTPA have
librium state within the Gaussian DOS. The activation energy relatively high degree of planarity, which leads to a high degree
for charge transport above Tc , which is assumed to be the energy of intermolecular wavefunction overlap.
difference between the equilibration energy of charge carriers The energetic disorder σ, namely, the fluctuation of the hop-
and the center of the DOS, is σ 2 /kT, whereas in the tempera- ping site energy, is understood as the fluctuation of polarization
ture region below Tc , the apparent activation energy for charge energy resulting from van der Waals and charge-dipole interac-
transport becomes less than σ 2 /kT because the packet of charge tions [33], and is affected by the variation of molecular geometry
carriers has not yet reached the thermodynamically equilibrium caused by bond rotation. It is thought that small conformational
state, and hence, the equilibration energy of charge carriers is changes caused by the rotations of C–N and/or C–C bonds
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KAGEYAMA et al.: ORGANIC PHOTOVOLTAIC DEVICES USING AN AMORPHOUS MOLECULAR MATERIAL 5

TABLE II
PERFORMANCE OF DEVICES A AND B

Fig. 6. Configurations of planar p-n heterojunction and p-i-n-type OPV


devices.

Fig. 8. J–V characteristics of ITO/PEDOT:PSS (ca. 30 nm)/TTPA


(30 nm)/C6 0 (40 nm)/LiF (0.1 nm)/Al (150 nm) (device A, circle) and
ITO/PEDOT:PSS (ca. 30 nm)/TTPA (27 nm)/TTPA:C6 0 (1:4, 20 nm)/C6 0
(23 nm)/LiF (0.1 nm)/Al (100 nm) (device B, square) in dark (dashed line)
and under AM1.5G illumination at an intensity of 100 mW/cm2 (solid line).
Fig. 7. EQE action spectrum of ITO/PEDOT:PSS (ca. 30 nm)/TTPA
(30 nm)/C6 0 (40 nm)/LiF (0.1 nm)/Al (150 nm) (device A, circle) and
ITO/PEDOT:PSS (ca. 30 nm)/TTPA (27 nm)/TTPA:C6 0 (1:4, 20 nm)/C6 0
(23 nm)/LiF (0.1 nm)/Al (100 nm) (device B, square) for transmitted light Fig. 8 shows the current density (J)–voltage (V) characteris-
through ITO electrode and electronic absorption spectra of vapor-deposited
films of TTPA (30 nm) and C6 0 (40 nm). tics of device A in the dark and under AM1.5G illumination at an
incident light intensity of 100 mW/cm2 . The performance of de-
vice A is listed in Table II. Device A exhibited an open-circuit
voltage (VOC ) of 0.95 ± 0.01 V, a short-circuit photocurrent
together with the absence of polar moieties in TTPA lead to the (JSC ) of 2.5 ± 0.1 mA/cm2 , an FF of 0.62 ± 0.01, and a PCE
smaller value of σ. of 1.5% for the incident AM1.5G light at 100 mW/cm2 .
It has recently been reported that a p-n heterojunction
OPV device using N,N -diphenyl-N,N -di(α-naphthyl)benzidine
C. Fabrication and Performance of Planar p-n Heterojunction (α-NPD) as an electron donor and C60 as an electron
and p-i-n-Type OPV Devices Using TTPA as an Electron Donor
acceptor, ITO/PEDOT:PSS (30 nm)/α-NPD (10 nm)/C60
A planar p-n heterojunction OPV device using TTPA (48 nm)/MgAg (170 nm), gave 1% PCE under AM1.5G il-
as an electron donor and C60 as an electron acceptor, lumination at intensities of 97 mW/cm2 [34]. Later, it was
ITO/PEDOT:PSS (ca. 30 nm)/TTPA (30 nm)/C60 (40 nm)/LiF also reported that a p-n heterojunction OPV device us-
(0.1 nm)/Al (150 nm) (device A), was fabricated, and the per- ing N,N -diphenyl-N,N -bis(3-methylphenyl)benzidine (TPD)
formance was examined. Fig. 6 shows the device configuration. as an electron donor and C60 as an electron acceptor gave
Fig. 7 shows the external quantum efficiency (EQE) action PCEs of 0.34%–1.0%, depending on the heat treatment
spectrum for the photogeneration of charge carriers for de- [35]. Other amorphous molecular materials that have hith-
vice A under illumination with transmitted monochromatic light erto been studied as OPV materials include 4,4 ,4 -tris[5-
through the ITO electrode, together with the electronic absorp- (dimesitylboryl)thiophen-2-yl]triphenylamine [36], 4,4 ,4 -
tion spectra of TTPA and C60 as their vapor-deposited film. tris[4-nitrophenyl(4-methylphenyl)amino]triphenylamine [36],
EQE for the photogeneration of charge carriers for transmitted m-MTDATA [37], [38], siloles containing carbazolyl groups
monochromatic light through the ITO electrode was calculated [39], π-conjugated systems consisting of triphenylamine and
from the equation EQE = (JSC /e)/(Iλ/hc), where JSC is the oligothiophene [40], triarylamine-substituted carbazole-based
short-circuit current, e is the elementary electronic charge, I and dendrimers with an oligothiophene core [41] perylene di-
λ are the intensity and the wavelength of transmitted monochro- imide derivatives [42], diphenylaminofluorenyl-capped thia-
matic light, respectively, h is the Planck constant, and c is the diazoloquinoxaline [43], π-conjugated systems consisting of
velocity of light in vacuum. The action spectra for the device triphenylamine and benzothiadiazole [44], and dithiafulvenyl-
resembled the superposition of the absorption spectra of TTPA derivatized triphenylamines [45]. The devices using these amor-
and C60 , exhibiting a peak EQE of ∼38%. phous molecular materials as electron donors have been reported
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

6 IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS

to exhibit JSC of 1–4 mA/cm2 , VOC of 0.3–0.9 V, FFs of 0.3–


0.5, and PCEs of 0.1%–1.3% under white-light irradiation. A
p-n heterojunction device using dicyanovinyl-substituted TTPA
as an electron donor and C60 as an electron acceptor has been re-
ported to give a PCE of 1.85% under AM1.5G illumination [46].
As compared with the reported performance of OPV de-
vices using the aforementioned amorphous molecular materi-
als [34]–[36], [40]–[45], the performance, in particular, the FF
value of 0.62 for device A is very high, which is comparable
to those reported for p-n heterojunction OPVs using polycrys-
talline CuPc [1].
It has been reported that the performance of OPV devices
Fig. 9. (a) Absorption spectra of mixed films of TTPA and [6,6]-PCBM pre-
is improved by the incorporation of a mixed interlayer be- pared from chlorobenene solution (film thickness: ca. 73 nm). (b) EQE action
tween electron donor and acceptor materials due to the in- spectra of ITO/PEDOT:PSS (ca. 30 nm)/TTPA:[6,6]-PCBM (ca. 73 nm)/LiF
crease in the fraction of photogenerated excitons that diffuse (0.1 nm)/Al (100 nm) for transmitted light through ITO electrode. TTPA:[6,6]-
PCBM ratios are 1:0.8 (solid line), 1:2 (dotted line), and 1:4 (dashed line).
to the donor/acceptor interface [6], [47]. In the present study,
a p-i-n-type OPV device, ITO/PEDOT:PSS (ca. 30 nm)/TTPA
(20 nm)/TTPA:C60 (1:4 molar ratio, 20 nm)/C60 (23 nm)/LiF
(0.1 nm)/Al (100 nm) (device B), was fabricated by the thermal
deposition method, and its performance was examined.
As shown in Fig. 7, device B, in which a mixed interlayer
with a thickness of 20 nm and a molar ratio of TTPA:C60 = 1:4
is incorporated into the TTPA and C60 layers, showed higher
EQEs than device A with a simple bilayer structure. The J–V
characteristics and the performance of device B are shown in
Fig. 8 and Table II, respectively. Device B exhibited a higher
JSC of 3.8 mA/cm2 and a higher PCE of 1.8% under AM1.5
illumination at an intensity of 100 mW/cm2 than device A.

D. Fabrication and Performance of Bulk p-n Heterojunction Fig. 10. J–V characteristics of ITO/PEDOT:PSS (ca. 30 nm)/TTPA:[6,6]-
PCBM (ca. 73 nm)/LiF (0.1 nm)/Al (100 nm) in dark (dashed line) and under
OPV Devices AM1.5G illumination at an intensity of 100 mW/cm2 (solid line). TTPA:[6,6]-
PCBM ratios are 1:0.8 (circle), 1:2 (square), and 1:4 (triangle).
It is expected that in the future, organic thin-film OPV devices
will be fabricated by preparing organic thin films by solution TABLE III
methods. Amorphous molecular materials are attractive in that PERFORMANCE OF BULK p-n HETEROJUNCTION OPV DEVICES USING TTPA
they permit the formation of smooth, uniform thin films by solu- AND [6,6]-PCBM AS ELECTRON DONOR AND ACCEPTOR MATERIALSa

tion methods. We have fabricated bulk p-n heterojunction OPV


devices, where the mixed-layer thin films of TTPA and C60
were prepared by spin coating from chlorobenzene solution,
ITO/PEDOT:PSS (ca. 30 nm)/TTPA:[6,6]-PCBM (ca. 73 nm
with molar ratios of 1:0.8, 1:2, and 1:4)/LiF (0.1 nm)/Al
(100 nm), and examined the performance.
Fig. 9 shows the EQE action spectrum for the devices together
with the electronic absorption spectra of the spin-coated mixed
films of TTPA and [6,6]-PCBM. It was found that both the EQE of [6,6]-PCBM to TTPA. The device using the active layer with
and the cell performance greatly depend on the molar ratios of a molar ratio of TTPA:[6,6]-PCBM = 1:4 exhibited a JSC of
TTPA and [6,6]-PCBM, and increase with the increasing molar 3.3 mA/cm2 , an FF of 0.40, and a PCE of 1.3% under the same
fraction of [6,6]-PCBM. The observed peak EQE was ca. 47% irradiation conditions.
at around 345 nm for the OPV device with a TTPA:[6,6]-PCBM
ratio of 1:4. Fig. 10 shows the J–V characteristics of the devices
in the dark and under AM1.5G illumination at an incident light E. Discussion
intensity of 100 mW/cm2 . The device performance is listed in Fig. 11 shows the schematic energy level diagram for the
Table III. The device using an active layer with a molar ratio device. It has been understood that VOC corresponds to the
of TTPA:[6,6]-PCBM = 1:0.8 exhibited poor performance with difference between the HOMO level of the donor and the LUMO
a VOC of 0.95 V, a JSC of 1.1 mA/cm2 , an FF of 0.27, and a level of the acceptor [4]. The VOC value of 0.95 V obtained for
PCE of 0.3% for the incident AM1.5G light at 100 mW/cm2 ; the present device corresponds to the difference between the
however, JSC and FF increased with the increasing molar ratio HOMO energy level of TTPA and the LUMO energy level of
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KAGEYAMA et al.: ORGANIC PHOTOVOLTAIC DEVICES USING AN AMORPHOUS MOLECULAR MATERIAL 7

the amorphous TTPA film (20 nm) on the PEDOT:PSS film.


The surface of the PEDOT:PSS film on the ITO glass substrate
exhibited a granular shape with a diameter of ∼50 nm and an
rms roughness of ca. 0.8 nm. The rms roughness of the surface
of the TTPA film was smaller than that of the PEDOT:PSS film,
which is ca. 0.1 nm. This result suggests that the amorphous
TTPA film smoothly covers the rough surface of PEDOT:PSS,
serving to prevent direct contact between the two electrodes.
Higher performance of the p-i-n-type OPV device (device B)
with the thin mixed interlayer relative to the planar p-n het-
Fig. 11. Schematic energy level diagram for the present device (values are
erojunction device (device A) is thought to be due to more
given in electronvolts with respect to vacuum level). The energy levels of efficient charge transfer between TTPA and C60 to generate
PEDOT:PSS, C6 0 , and Al were taken from [2] and [50]. hole–electron pairs.
It is of interest to note that bulk p-n heterojunction devices
consisting of a single layer of a mixture of TTPA and [6,6]-
PCBM, which is prepared by spin coating from solution, ex-
hibited high performance, greatly depending on the molar ratio
of TTPA and [6,6]-PCBM. It is suggested that charge transfer
between the donor and the acceptor to generate hole–electron
pairs becomes more efficient and that the recombination of hole–
electron pairs is not significant. The improvement in perfor-
mance with the increasing molar fraction of [6,6]-PCBM from
1:0.8 to 1:4 (TTPA:[6,6]-PCBM) may be attributed partly to
Fig. 12. Topographic AFM images of (a) PEDOT:PSS film (ca. 30 nm) on the improvement of the light absorption efficiency and partly to
ITO glass substrate and (b) TTPA amorphous film (20 nm) on the PEDOT:PSS more facile electron transport, which leads to better balanced
film. The size of the images is 1 µm × 1 µm.
hole and electron mobilities.
The present results have shown that amorphous molecular
C60 . The relatively high solid-state Ip of TTPA leads to the high materials, which permit the formation of uniform thin films by
VOC for the present device. both thermal deposition and solution methods, are promising
The photocurrent results mainly from the photoabsorption by candidates for materials in OPV devices. The poor overlap be-
C60 or [6,6]-PCBM, as evidenced from the absorption spectra tween the absorption spectrum of TTPA and the solar spectrum
of TTPA, C60 , and [6,6]-PCBM in Figs. 7 and 9. Therefore, the limits the performance of the present OPV devices. Amorphous
high EQEs for the photogeneration of charge carriers, relatively molecular materials with both high charge-carrier mobilities
high JSC , and the high FF for the present devices are suggested to and the absorption spectrum that overlapped well with the solar
be due to the large exciton diffusion length of C60 (ca. 40 nm) [2] spectrum are expected to give much higher PCE, and the de-
and the high charge-carrier mobilities of TTPA, C60 [48], and velopment of such materials is required for further research and
[6,6]-PCBM [49]. development of OPV devices. In addition, higher Tg s are desir-
The observed current density (Jobs ) under illumination is ex- able from the standpoint of long-term stability of OPV devices
pressed as the subtraction of the dark current density from the for practical applications. The incorporation of aryl groups into
photocurrent density. The dark current density is the sum of TTPA or the replacement of the phenyl group in TTPA by an-
the diode current density and the leakage current density. For other rigid moiety can be molecular design concepts for high Tg
attaining high FF, Jobs under given forward voltage should be materials, according to the proposed molecular design concepts
large. Since Jobs decreases when the series resistance is large, for attaining high Tg [10], [11].
the series resistance should be small. It has also been reported
that series resistance of OPV devices plays an important role
in ensuring high FF [34]. The high charge-carrier mobilities of IV. CONCLUSION
TTPA and C60 may contribute to reduce the series resistance of Charge transport in the TTPA amorphous glass and the per-
the present devices. formance of OPV devices using TTPA as an electron donor and
Leakage current density, which flows through the shunt, in- C60 or [6,6]-PCBM as an electron acceptor are discussed in
creases linearly with the increasing forward voltage, and the this paper. The TTPA amorphous glass exhibited a hole drift
existence of the leakage current density at a given forward volt- mobility of 1.1 × 10−2 cm2 /V·s at an electric field of 1.0 ×
age decreases Jobs , and hence, FF. Therefore, for attaining high 105 V/cm and at 293 K, as determined by the TOF method.
FF, the shunt resistance should be as large as possible. The ho- This value is the highest level among amorphous molecular ma-
mogeneous, grain-boundary-free films of TTPA are suggested to terials. A planar p-n heterojunction OPV device using TTPA
prevent any “shorts” between electrodes, leading to high shunt as an electron donor and C60 as an electron acceptor showed
resistance. Fig. 12 shows the topographic AFM images of the high performance, exhibiting an FF of 0.62 and a PCE of 1.5%
PEDOT:PSS film (ca. 30 nm) on the ITO glass substrate and under AM1.5G illumination at 100 mW/cm2 . A p-i-n-type OPV
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

8 IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS

device having a mixed interlayer of TTPA and C60 (TTPA:C60 [18] P. Schilinsky, C. Waldauf, J. Hauch, and C. J. Brabec, “Simulation of
= 1:4 molar ratio with a thickness of 20 nm) exhibited a higher light intensity dependent current characteristics of polymer solar cells,”
J. Appl. Phys., vol. 95, no. 5, pp. 2816–2819, 2004.
PCE of 1.8% than the planar p-n heterojunction device. A bulk [19] Y. Shirota, S. Nomura, and H. Kageyama, “Charge transport in amorphous
p-n heterojunction OPV device using the film of a mixture of molecular materials,” Proc. SPIE, vol. 3476, pp. 132–141, 1998.
TTPA and [6,6]-PCBM (1:4 molar ratio) prepared by spin coat- [20] K. Okumoto, K. Wayaku, T. Noda, H. Kageyama, and Y. Shirota, “Amor-
phous molecular materials: Charge transport in the glassy state of N,N -
ing from chlorobenzene solution exhibited a PCE of 1.3%. The diphenyl-[1,1 -biphenyl]-4,4 -diamines,” Synth. Metal, vol. 111/112,
high performance of the present devices is thought to be at- pp. 473–476, 2000.
tributable to the high charge-carrier mobilities of TTPA and [21] Y. Shirota, K. Okumoto, H. Ohishi, M. Tanaka, M. Nakao, K. Wayaku,
S. Nomura, and H. Kageyama, “Charge transport in amorphous molecular
C60 , the relatively high Ip of TTPA, and also its capability of materials,” Proc. SPIE, vol. 5937, pp. 132–141, 1998.
forming homogeneous, grain-boundary-free films without pin- [22] H. Ohishi, M. Tanaka, H. Kageyama, and Y. Shirota, “Amorphous molec-
holes. The present results have demonstrated that amorphous ular materials with high carrier mobilities: Thiophene- and selenophene-
containing tris(oligoarylenyl)amines,” Chem. Lett., vol. 33, no. 10,
molecular materials, as well as crystalline molecular materials, pp. 1266–1267, 2004.
are also promising candidates for OPV materials. [23] H. Kageyama, H. Ohishi, M. Tanaka, Y. Ohmori, and Y. Shirota, “High
performance organic photovoltaic devices using amorphous molecular
materials with high charge-carrier drift mobilities,” Appl. Phys. Lett.,
REFERENCES vol. 94, no. 6, pp. 063304-1–063304-3, 2009.
[24] J. Sakai, H. Kageyama, S. Nomura, H. Nakano, and Y. Shirota, “Photo- and
[1] C. W. Tang, “Two-layer organic photovoltaic cell,” Appl. Phys. Lett.,
electro-active amorphous molecular material: Morphology, structures, and
vol. 48, no. 2, pp. 183–185, 1986.
hole transport properties of tris[4-(2-thienyl)phenyl]amine,” Mol. Cryst.
[2] P. Peumans, A. Yakimov, and S. R. Forrest, “Small molecular weight
Liquid Cryst., vol. 296, pp. 445–463, 1997.
organic thin-film photodetectors and solar cells,” J. Appl. Phys., vol. 93,
[25] J. M. Kroon, M. M. Wienk, W. J. H. Verhees, and J. C. Hummelen,
no. 7, pp. 3693–3723, 2003.
“Accurate efficiency determination and stability studies of conjugated
[3] Y. Shirota and H. Kageyama, “Charge carrier transporting molecular ma-
polymer/fullerene solar cells,” Thin Solid Films, vol. 403/404, pp. 223–
terials and their applications in devices,” Chem. Rev., vol. 107, no. 4,
228, 2002.
pp. 953–1010, 2007.
[26] A. J. Bard and L. R. Faulkner, Electrochemical Methods. New York:
[4] S. Günes, H. Neugebauer, and N. S. Sariciftci, “Conjugated polymer-
Wiley, 2001.
based organic solar cells,” Chem. Rev., vol. 107, no. 4, pp. 1324–1338,
[27] Y. Shirota, Y. Kuwabara, H. Inada, T. Wakimoto, H. Nakada,
2007.
Y. Yonemoto, S. Kawami, and K. Imai, “Multilayered organic elec-
[5] J. Xue, S. Uchida, B. P. Rand, and S. R. Forrest, “Asymmetric tandem
troluminescent device using a novel starburst molecule, 4,4 ,4 -tris(3-
organic photovoltaic cells with hybrid planar-mixed molecular hetero-
methylphenylphenylamino)triphenylamine, as a hole transport material,”
junctions,” Appl. Phys. Lett., vol. 85, no. 23, pp. 5757–5759, 2004.
Appl. Phys. Lett., vol. 65, no. 7, pp. 807–809, 1994.
[6] J. Xue, B. P. Rand, S. Uchida, and S. R. Forrest, “A hybrid planar-mixed
[28] Z. Bao, A. J. Lovinger, and A. Dodabalapur, “Organic field-effect tran-
molecular heterojunction photovoltaic cell,” Adv. Mater., vol. 17, no. 1,
sistors with high mobility based on copper phthalocyanine,” Appl. Phys.
pp. 66–71, 2005.
Lett., vol. 69, no. 20, pp. 3066–3068, 1996.
[7] M. Reyes-Reyes, K. Kim, J. Dewald, R. Lopez-Sandoval, A. Avadhanula,
[29] H. Bässler, “Localized states and electronic transport in single component
S. Curran, and D. L. Carroll, “Meso-structure formation for enhanced
organic solids with diagonal disorder,” Phys. Status Solidi B, vol. 107,
organic photovoltaic cells,” Org. Lett., vol. 7, no. 26, pp. 5749–5752,
no. 1, pp. 9–53, 1981.
2005.
[30] H. Bässler, “Charge transport in disordered organic photoconductors,”
[8] J. Y. Kim, K. Lee, N. E. Coates, D. Moses, T.-Q. Nguyen, M. Dante, and
Phys. Status Solidi B, vol. 175, no. 1, pp. 15–56, 1993.
A. J. Heeger, “Efficient tandem polymer solar cells fabricated by all-
[31] P. M. Borsenberger, L. T. Pautmeier, and H. Bässler, “Nondispersive-
solution processing,” Science, vol. 317, no. 5835, pp. 222–225, 2007.
to-dispersive charge-transport transition in disordered molecular solids,”
[9] Y. Liang, D. Feng, Y. Wu, S.-T. Tsai, G. Li, C. Ray, and L. Yu, “Highly
Phys. Rev. B, vol. 46, no. 19, pp. 12145–12153, 1992.
efficient solar cell polymers developed via fine-tuning of structural and
[32] N. F. Mott, “Electrons in non-crystalline materials,” in Electronic and
electronic properties,” J. Amer. Chem. Soc., vol. 131, no. 22, pp. 7792–
Structural Properties of Amorphous Semiconductors, P. G. L. Comber
7799, 2009.
and J. Mort, Eds. New York: Academic, 1973.
[10] Y. Shirota, “Organic materials for electronic and optoelectronic devices,”
[33] P. M. Borsenberger and J. J. Fitzgerald, “Effects of the dipole moment on
J. Mater. Chem., vol. 10, no. 1, pp. 1–25, 2000.
charge transport in disordered molecular solids,” J. Phys. Chem., vol. 97,
[11] Y. Shirota, “Photo- and electroactive amorphous molecular materials—
no. 18, pp. 4815–4819, 1993.
Molecular design, synthesis, reactions, properties, and applications,” J.
[34] G. P. Kushto, W. Kim, and Z. H. Kafafi, “Flexible organic photovoltaics
Mater. Chem., vol. 15, no. 1, pp. 75–93, 2005.
using conducting polymer electrodes,” Appl. Phys. Lett., vol. 86, no. 9,
[12] P. Peumans and S. R. Forrest, “Separation of geminate charge-pairs
pp. 093502-1–093502-3, 2005.
at donor–acceptor interfaces in disordered solids,” Chem. Phys. Lett.,
[35] T. Osasa, S. Yamamoto, and M. Matsumura, “Organic solar cells by an-
vol. 398, no. 1–3, pp. 27–31, 2004.
nealing stacked amorphous and microcrystalline layers,” Adv. Funct.
[13] V. D. Mihailetchi, L. J. A. Koster, J. C. Hummelen, and P. W. M. Blom,
Mater., vol. 17, no. 15, pp. 2937–2942, 2007.
“Photocurrent generation in polymer-fullerene bulk heterojunctions,”
[36] M. Kinoshita, N. Fujii, T. Tsuzuki, and Y. Shirota, “Creation of novel light
Phys. Rev. Lett., vol. 93, no. 21, pp. 216601-1–216601-4, 2004.
sensitive amorphous molecular materials and their photovoltaic proper-
[14] C. R. McNeill, S. Westenhoff, C. Groves, R. H. Friend, and N. C.
ties,” Synth. Metal, vol. 121, no. 1–3, pp. 1571–1572, 2001.
Greenham, “Influence of nanoscale phase separation on the charge gen-
[37] Y. Shirota, T. Kobata, and N. Noma, “Starburst molecules for amorphous
eration dynamics and photovoltaic performance of conjugated polymer
molecular materials. 4,4 ,4 -Tris(N,N-diphenylamino)triphenylamine
blends: Balancing charge generation and separation,” J. Phys. Chem. C,
and 4,4 ,4 -tris[N-(3-methylphenyl)-N-phenylamino]triphenylamine,”
vol. 111, no. 51, pp. 19153–19160, 2007.
Chem. Lett., vol. 18, no. 7, pp. 1145–1148, 1989.
[15] H. Ohkita, S. Cook, Y. Astuti, W. Duffy, S. Tierney, W. Zhang, M. Heeney,
[38] Z. R. Hong, C. S. Lee, S. T. Lee, W. L. Li, and Y. Shirota, “Bifunctional
I. McCulloch, J. Nelson, D. D. C. Bradley, and J. R. Durrant, “Charge car-
photovoltaic and electroluminescent devices using a starburst amine as an
rier formation in polythiophene/fullerene blend films studied by transient
electron donor and hole-transporting material,” Appl. Phys. Lett., vol. 81,
absorption spectroscopy,” J. Amer. Chem. Soc., vol. 130, no. 10, pp. 3030–
no. 15, pp. 2878–2880, 2002.
3042, 2008.
[39] B. Mi, Y. Dong, Z. Li, J. W. Y. Lam, M. Häußler, H. H. Y. Sung,
[16] C. Yin, T. Kietzke, D. Neher, and H.-H. Hörhold, “Photovoltaic properties
H. S. Kwok, Y. Dong, I. D. Williams, Y. Liu, Y. Luo, Z. Shuai, D. Zhu, and
and exciplex emission of polyphenylenevinylene-based blend solar cells,”
B. Z. Tang, “Making silole photovoltaically active by attaching carbazolyl
Appl. Phys. Lett., vol. 90, no. 9, pp. 092117-1–092117-3, 2007.
donor groups to the silolyl acceptor core,” Chem. Commun., no. 28,
[17] R. A. Marsh, C. Groves, and N. C. Greenham, “A microscopic model for
pp. 3583–3585, 2005.
the behavior of nanostructured organic photovoltaic devices,” J. Appl.
[40] A. Cravino, S. Roquet, O. Alévêque, P. Leriche, P. Frère, and
Phys., vol. 101, no. 8, pp. 083509-1–083509-7, 2007.
J. Roncali, “Triphenylamine–oligothiophene conjugated systems as
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KAGEYAMA et al.: ORGANIC PHOTOVOLTAIC DEVICES USING AN AMORPHOUS MOLECULAR MATERIAL 9

organic semiconductors for optoelectronics,” Chem. Mater., vol. 18, Masatake Tanaka received the B.Eng. and M.Eng. degrees from the Depart-
no. 10, pp. 2584–2590, 2006. ment of Applied Chemistry, Faculty of Engineering, Osaka University, Suita,
[41] J. Lu, P. F. Xia, P. K. Lo, Y. Tao, and M. S. Wong, “Synthesis and prop- Japan, in 2002 and 2004, respectively.
erties of multi-triarylamine-substituted carbazole-based dendrimers with He is currently with the Department of Applied Chemistry, Faculty of Engi-
an oligothiophene core for potential applications in organic solar cells and neering, Osaka University.
light-emitting diodes,” Chem. Mater., vol. 18, no. 26, pp. 6194–6203,
2006.
[42] W. S. Shin, H.-H. Jeong, M.-K. Kim, S.-H. Jin, M.-R. Kim, J.-K. Lee,
J. W. Lee, and Y.-S. Gal, “Effects of functional groups at perylene diimide
derivatives on organic photovoltaic device application,” J. Mater. Chem.,
vol. 16, no. 4, pp. 384–390, 2006.
[43] M. Sun, L. Wang, X. Zhu, B. Du, R. Liu, W. Yang, and Y. Cao, “Near-
infrared response photovoltaic device based on novel narrow band gap Yutaka Ohmori (M’79–SM’97–F’05) received the
small molecule and PCBM fabricated by solution processing,” Solar B.Eng. degree in 1972 and the Dr. Eng. degree in
Energy Mater. Solar Cells, vol. 91, no. 18, pp. 1681–1687, 2007. 1979 from the Department of Electrical Engineer-
[44] C. He, Q. He, Y. Yi, G. Wu, F. Bai, Z. Shuai, and Y. Li, “Improving ing, Faculty of Engineering, Osaka University, Suita,
the efficiency of solution processable organic photovoltaic devices by Japan.
a star-shaped molecular geometry,” J. Mater. Chem., vol. 18, no. 34, In 1977, he joined Nippon Telegraph Telephone
pp. 4085–4090, 2008. Public Corporation (now NTT Corporation), where
[45] O. Alévêque, P. Leriche, N. Cocherel, P. Frère, A. Cravino, and J. Roncali, he was mainly involved in optical semiconductor de-
“Star-shaped conjugated systems derived from dithiafulvenyl-derivatized vices. In 1989, he became an Associate Professor in
triphenylamines as active materials for organic solar cells,” Solar Energy the Department of Electronic Engineering, Faculty of
Mater. Solar Cells, vol. 92, no. 9, pp. 1170–1174, 2008. Engineering, Osaka University, where he has been a
[46] A. Cravino, P. Leriche, O. Alévêque, S. Roquet, and J. Roncali, “Light- Professor in the Collaborative Research Center for Advanced Science and Tech-
emitting organic solar cells based on a 3D conjugated system with internal nology (Center for Advanced Science and Innovation since 2004), Electronic
charge transfer,” Adv. Mater., vol. 18, no. 22, pp. 3033–3037, 2006. Materials and Systems Engineering, since 2000, and was engaged in optical and
[47] M. Hiramoto, H. Fujiwara, and M. Yokoyama, “Three-layered organic electrical devices utilizing organic materials, including conducting polymers.
solar cell with a photoactive interlayer of codeposited pigments,” Appl. Prof. Ohmori is a Fellow of the Institute of Electronics, Information and
Phys. Lett., vol. 58, no. 10, pp. 1062–1064, 1991. Communication Engineers in Japan, and a member of the Japan Society of
[48] E. Frankevich, Y. Maruyama, and H. Ogata, “Mobility of charge carriers Applied Physics, the American Physical Society, and the Materials Research
in vapor-phase grown C6 0 single crystal,” Chem. Phys. Lett., vol. 214, Society.
no. 1, pp. 39–44, 1993.
[49] V. D. Mihailetchi, J. K. J. van Duren, P. W. M. Blom, J. C. Hummelen,
R. A. J. Janssen, J. M. Kroon, M. T. Respens, W. J. H. Verhees, and
M. M. Wienk, “Electron transport in a methanofullerene,” Adv. Funct.
Mater., vol. 13, no. 1, pp. 43–46, 2003.
[50] A. J. Mäkinen, I. G. Hill, R. Shashidhar, N. Nikolov, and Z. H. Kafafi,
“Hole injection barriers at polymer anode/small molecule interfaces,”
Appl. Phys. Lett., vol. 79, no. 5, pp. 557–559, 2001. Yasuhiko Shirota was born in 1940. He received the
B.Eng. degree in 1963 and the Dr. Eng. degree in 1968
from the Department of Applied Chemistry, Faculty
of Engineering, Osaka University, Suita, Japan.
In 1968, he was appointed as a Research Associate
at Osaka University, where he was promoted to an
Associate Professor in 1972, a full Professor in 1986,
Hiroshi Kageyama was born in 1969. He received and has been a Professor Emeritus since 2003. Since
the B.Eng., M.Eng., and Ph.D. degrees from the De- 2003, he has also been a Professor at Fukui University
partment of Applied Chemistry, Faculty of Engineer- of Technology, Fukui, Japan. His current research
ing, Osaka University, Suita, Japan, in 1992, 1994, interests include a wide field of organic materials
and 1997, respectively. science, including the synthesis, structures, reactions, properties, functions of
In 1997, he was appointed as a Research Associate both molecular materials and polymers, and their applications in devices.
at Osaka University, where he has been an Assistant Prof. Shirota is a member and a Fellow of the Chemical Society of Japan, and a
Professor since 2006. His current research interests member of the Society of Polymer Science, Japan, the Japanese Photochemistry
include charge transport in organic disordered sys- Association, the Society of Synthetic Organic Chemistry, and the Japan Society
tems. of Applied Physics.
Dr. Kageyama is a member of the Chemical Soci-
ety of Japan, the Society of Polymer Science, Japan, the Japanese Photochem-
istry Association, and the Japan Society of Applied Physics.

Hitoshi Ohishi received the B.Eng. and M.Eng. degrees from the Department
of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Japan,
in 1999 and 2001, respectively.
He is currently with the Department of Applied Chemistry, Faculty of Engi-
neering, Osaka University.

S-ar putea să vă placă și