ArticlePDF Available

Ternary hybrid systems of P3HT-CdSe-WS2 nanotubes for photovoltaic applications

Authors:

Abstract and Figures

Hybrid heterojunctions of conjugated polymers and inorganic nanomaterials are a promising combination for obtaining high performance solar cells (SC). In this work we have explored new possible uses of the WS2 nanotubes (NTs) both as the only acceptor material blended with a polymer and in ternary systems mixed with a polymer and quantum dots (QDs). In particular we have spectroscopically investigated binary blends of poly(3-hexylthiophene) (P3HT) and WS2 NTs, P3HT and CdSe QDs, and ternary blends of P3HT, CdSe QDs and WS2 NTs. We report fluorescence quenching effects of the QD signal in the P3HT-CdSe-WS2 system with the increase of NT concentration. Static and time-resolved fluorescence studies reveal efficient resonant energy transfer from the QDs to the NTs upon photoexcitation. The evidence of energetic interaction between WS2 NTs and QDs opens new fields of application of WS2 NTs and holds very promising potential for improving charge transfer phenomena in the active layer of hybrid solar cells.
Content may be subject to copyright.
17998 |Phys. Chem. Chem. Phys., 2014, 16, 17998--18003 This journal is ©the Owner Societies 2014
Cite this: Phys. Chem. Chem. Phys.,
2014, 16, 17998
Ternary hybrid systems of P3HT–CdSe–WS
2
nanotubes for photovoltaic applications
A. Bruno,*
ab
C. Borriello,
a
S. A. Haque,
b
C. Minarini
a
and T. Di Luccio
a
Hybrid heterojunctions of conjugated polymers and inorganic nanomaterials are a promising combination
for obtaining high performance solar cells (SC). In this work we have explored new possible uses of the
WS
2
nanotubes (NTs) both as the only acceptor material blended with a polymer and in ternary systems
mixed with a polymer and quantum dots (QDs). In particular we have spectroscopically investigated
binary blends of poly(3-hexylthiophene) (P3HT) and WS
2
NTs, P3HT and CdSe QDs, and ternary blends
of P3HT, CdSe QDs and WS
2
NTs. We report fluorescence quenching effects of the QD signal in the
P3HT–CdSe–WS
2
system with the increase of NT concentration. Static and time-resolved fluorescence
studies reveal efficient resonant energy transfer from the QDs to the NTs upon photoexcitation. The
evidence of energetic interaction between WS
2
NTs and QDs opens new fields of application of WS
2
NTs and holds very promising potential for improving charge transfer phenomena in the active layer of
hybrid solar cells.
1 Introduction
The elevated cost of silicon-based photovoltaics and depletion
of material stocks have given a strong input to search for new
and less expensive materials for solar energy conversion. A valid
alternative is represented by purely organic or hybrid organic–
inorganic solar cells as they do not require expensive high
temperature processes and can be easily processed on a large
scale.
1
In particular, bulk heterojunctions of conjugated polymers,
as donor materials, and inorganic nanomaterials, as acceptor
materials, can offer numerous advantages. On one hand, poly-
mers having large absorption coefficients permit us to realize
efficient thin film based devices. Polymeric thin films can be
deposited by different solution based deposition techniques
2
on rigid or flexible substrates also with special morphologies
targeted to enhance the cell performances. However, much
effort is being devoted to overcome major limitations of poly-
mer based photovoltaic devices i.e. degradation and limited
time stability. To this purpose air stable polymers, efficient
encapsulation materials and other solutions are under develop-
ment and have been extensively discussed by Jo
¨rgensen’s review.
3
On the other hand, inorganic nanomaterials do not easily degrade
over time
4
providing a longer lifetime to the hybrid blends and
related electronic devices.
5–7
Inorganic quantum dots (QDs) are
the most used inorganic nanomaterials in hybrid blends for
photovoltaics because of their tunable band-gap over visible and
infra-red spectral ranges. In addition, multiple exciton generation
effects occurring in several types of QDs can potentially lead to
high photovoltaic conversion efficiency.
8
Despite these promising properties one of the main factors
limiting the performances of polymer–QD heterojunction solar
cells is the poor charge transport among QDs in blends, due to
the low carrier mobility leading to cell efficiencies of still
around 4%.
9
The low carrier mobility of QDs is often ascribed
to the inefficient charge transfer processes between the nano-
crystals themselves due to the presence of organic capping
agents, which are used during their synthesis in order to avoid
nanocrystal aggregation and the loss of quantum properties.
These organic ligands are usually long alkyl chains that need to be
replaced by shorter ligands after the synthesis through suitable
ligand exchange processes that improve charge mobility while
retaining the good miscibility of the dots in the polymer solution
for a homogeneous blend.
Alternative non-toxic inorganic nanomaterials that do not
require the use of ligands during and after their synthesis are
tungsten disulfide (WS
2
) nanotubes (NTs). Moreover, they can
extend the absorption region of the blend into the infrared
region. Bulk WS
2
is a metal dichalcogenide compound char-
acterized by a direct bandgap at B1.95 eV and a small indirect
bandgap at B1.3 eV.
10
WS
2
nanotubes were firstly synthesized
in 1992
11
and nowadays pure multiwall WS
2
nanotubes are
a
ENEA, Italian National Agency for New Technologies, Energy and Sustainable
Economic Development, P.le Enrico Fermi 1, Portici (NA), Italy.
E-mail: annalisa.bruno@enea.it
b
Department of Chemistry, Imperial College London, South Kensington Campus,
London SW7 2AZ, UK
Electronic supplementary information (ESI) available: The optical scheme for
the fluorescence upconversion system, fluorescence decays of the P3HT–WS
2
and
P3HT–CdSe blends and the fluorescence peak position vs. QD concentrations.
See DOI: 10.1039/c4cp00594e
Received 10th February 2014,
Accepted 15th July 2014
DOI: 10.1039/c4cp00594e
www.rsc.org/pccp
PCCP
PAPER
Published on 15 July 2014. Downloaded by Imperial College London Library on 05/09/2014 06:18:09.
View Article Online
View Journal
| View Issue
This journal is ©the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 17998--18003 | 17999
available in macroscopic quantities on the market (about
0.5 kg per day).
12
WS
2
nanotubes are known mainly for their
excellent mechanical behavior
13
and reinforcing properties of
polymer composites
14
based on resins,
15
thermoplastic polymers,
16
and elastomers.
17
Just recently the first nanocomposites of WS
2
NTs and conductive polymers have been reported. They have
shown to increase polyaniline conductivity by efficient doping
18
and preserve OLED device performances until high percentage
loading(about10%).
19
Anyway,theiruseinhybridblendsisquite
challenging and unexplored.
In this work we have investigated the possibility of using
WS
2
both as the sole acceptor material blended with a polymer
and in systems of a polymer and QDs to improve the perfor-
mances of this better known system. We report a series of both
static and time-resolved spectroscopic studies on binary blend
systems, composed of poly(3-hexylthiophene) (P3HT) and WS
2
NTs (P3HT/WS
2
), and P3HT and cadmium selenide (CdSe) QDs
(P3HT/CdSe), and a ternary blend system (P3HT/CdSe/WS
2
). The
ternary blends were prepared keeping a fixed ratio between
P3HT and CdSe, and varying the concentrations of WS
2
nano-
tubes in a range between 9 and 33 wt% with respect to the total
weight of the blend.
2 Methods
2.1 Materials
The P3HT polymer was purchased from Sigma-Aldrich, WS
2
nanotubes from NanoMaterials Ltd, and CdSe QDs (absorbing
at 620 nm) from NN-Labs. All the solvents were furnished by
Sigma-Aldrich and used as received.
2.2 Binary and ternary blend preparation
WS
2
nanotubes were purified and disagglomerated before use.
1 g of material was dispersed in 1 l of 2-propanol and sonicated
for 2 hours at low power. Successively they were centrifuged at
1500 rpm for 15 minutes, and the supernatant was collected
and dried in a vacuum to give purified nanotubes. To prepare
the P3HT–WS
2
binary blends, P3HT was dissolved in chloro-
benzene at the concentration of 20 mg ml
1
and the solution
was stirred for 15 minutes at 60 1C and for one hour at room
temperature. It was then filtered through a 0.2 mm PTFE filter
and mixed with different amounts of WS
2
chlorobenzene
suspension (20 mg ml
1
) to obtain blends with WS
2
loading
of 5, 10, 25, 50, and 75 wt%.
The QDs were provided in toluene solution by the company.
In order to remove the excess ligand they were treated before
usage. Indeed after precipitation and washing in acetone, they
were dried and re-dissolved in chlorobenzene at a concentration
of 20 mg ml
1
. P3HT–CdSe binary blends were prepared by
mixing different amounts of CdSe QD chlorobenzene solution
with the polymer solution to obtain blends with QD loading of 25,
50, and 70 wt%. The resulting mixtures were stirred for 2 hours at
room temperature to obtain homogenous nanocomposites.
For the preparation of ternary blends we followed a two step
process. First a fixed volume of CdSe QD chlorobenzene solution
(20 mg ml
1
) was mixed with the appropriate amount of solid WS
2
nanotubes to obtain the CdSe : WS
2
weight ratio of 1:0.2, 1:0.5,
1:0.75, and 1:1. The mixtures were sonicated for 40 minutes.
Then the same volume of chlorobenzene polymer solution at
the same concentration (20 mg ml
1
) was added to each new
solution. In this way it was possible to maintain the P3HT:CdSe
weight ratio constant while varying the weight concentration of
the nanotubes in the ternary mixture (9, 20, 27, and 33 wt%).
All the solutions were deposited by spin coating on glass
substrates at 1000 rpm for 30 s, at room temperature. The
substrates were pre-cleaned by sonication in d.i. water and the
deconex detergent, acetone and subsequently in 2-propanol.
The films were homogeneous and their thickness was in the
range of 80–100 nm as measured using a Tencor profilometer.
2.3 Optical measurements
The pure polymer and the nanocomposite blend layers were
characterized by UV-Vis absorption spectroscopy using a Perkin
Elmer Lambda 900 Spectrophotometer. Photoluminescence
emission (PL) of the same films was measured using a Fluorolog
3 instrument, Horiba Jobin Yvon Instruments SA.
The ultra-fast fluorescence emission experiments were per-
formed using a femtosecond laser based system. The Second
Harmonic output of a mode-locked Ti:Sapphire oscillator with the
wavelength fixed at 800 nm was used as the excitation beam. The
pulse duration was 70 fs and the repetition rate 80 MHz. A portion
of this fundamental beam was frequency doubled to create the
excitation beam with a wavelength of 400 nm. Fluorescence from
the sample was focused on a beta barium borate (BBO) crystal
along with the 800 nm fundamental beam. The fluorescence
emitted at different emission wavelengths was mixed in the BBO
crystal with the gate beam to generate sum frequency photons,
which were detected using a photomultiplier tube. Data were
acquired using Lab-View software and subsequently analyzed.
The temporal reconstruction of the signal was obtained through
a micrometrical precise delay line on the gate beam. The resolu-
tion of the system was measured to be 150 fs. Sample degradation
was avoided by performing the measurements under flowing
nitrogen and using a translation stage to move the sample within
the beam, removing the effect of photobleaching and providing
data averaged across the whole sample. The system has been
already described in previous studies
20,21
and the optical scheme is
reported in the ESIsection (Fig. S1).
3 Results and discussion
In P3HT–WS
2
blends we have investigated the possibility of
using WS
2
NTs as an electron acceptor due to the favorable
band alignment between the polymer and the NTs. In Fig. 1(a)
the highest occupied molecular orbital (HOMO) and lowest
unoccupied molecular orbital (LUMO) levels are reported both
for WS
2
and P3HT.
10,22
Fig. 1(b) shows the UV-visible absorption spectra of P3HT/
WS
2
nanocomposites with an increasing amount of WS
2
NTs,
ranging from 5 to 50 wt%, together with those of pristine P3HT
Paper PCCP
Published on 15 July 2014. Downloaded by Imperial College London Library on 05/09/2014 06:18:09.
View Article Online
18000 |Phys. Chem. Chem. Phys., 2014, 16, 17998--18003 This journal is ©the Owner Societies 2014
and WS
2
films for comparison. As expected the absorption
signal between 400 and 630 nm, mainly due to the polymer,
reduces as a function of the NT concentration because the
relative amount of P3HT in the blend decreases accordingly.
Moreover, a clear effect of the NT addition is an increase of the
absorption of nanocomposites in the region between 630 and
800 nm, around the maximum absorption of NTs.
The emission spectra of pure P3HT and P3HT/WS
2
nano-
composite films are shown in Fig. 1(c). All the emissions have
been collected, after excitation at 400 nm, in whole visible and
near infrared regions. The fluorescence signals of the blends
are essentially dominated by the polymer contribution, since
the NTs are not emissive, being an indirect bandgap material.
23
The fluorescence spectra have been corrected by the number
of photons absorbed at 400 nm where the sample was excited.
It is clear that the fluorescence intensity of the P3HT is just very
slightly quenched by the presence of the WS
2
NTs indicating
that no transfer mechanism is taking place between the two
materials in the blends. Indeed, the quenching values are quite
small (around 10%) for all NT concentrations. This suggests
that such small variation can be attributed more to the different
thickness and uniformity of the films and/or experimental
uncertainty than to a real quenching effect. This result has
also been confirmed by the time resolved measurements per-
formed on all the blends showing that the lifetime of the
polymer does not change due to the presence of the NTs. These
data are reported in the ESI.
An alternative way to exploit the properties of WS
2
NTs could
be using them in ternary blends of polymer and inorganic QDs.
In previous studies ternary hybrid systems composed of polymers
blended with carbon nanotubes (CNT) and QDs have shown
interesting physical phenomena such as long-lived charge transfer
and luminescence quenching via energy transfer and reduced
blinking.
24
In these systems the NT backbone is expected to
promote carrier mobility
25
and QD dispersion that are both very
relevant factors for an efficient bulk heterojunction photovoltaic
solar cell.
Starting from these observations, we wish to explore the
spectroscopic properties of the ternary blend system P3HT–
QDs–WS
2
, with the aim of testing its application as an active layer
in hybrid solar cells with respect to the binary system P3HT–CdSe.
P3HT–CdSe hybrid blends are among the most studied
hybrid nanocomposites where the polymer photoluminescence is
quenched by the CdSe quantum dots.
26,27
Nevertheless, there is
room for addressing some critical points in this binary system. In
particular the efficiency of charge separation at the donor/acceptor
interface is crucial to the photocurrent generation in hydrid solar
cells, and a complete understanding of this process is essential to
optimize it.
28–30
In order to follow this route, different amounts of
WS
2
nanotubes have been added to the P3HT–CdSe blends. The
idea is to facilitate the charge generation at the interface and make
possible the energy transfer between the QDs and the NTs. In our
case, upon adding NTs, P3HT is the donor material and both CdSe
QDs and WS
2
NTs could act as the acceptors.
In the framework of Fo
¨rster theory of fluorescence resonance
energy transfer, the energy transfer phenomenon is mediated
by a long-range dipole–dipole interaction between donor and
acceptor molecules. The occurrence of an energy transfer event
requires spectral overlap of the donor emission spectrum with
the acceptor absorption spectrum for energy conservation.
The maximum distance over which Fo
¨rster energy transfer can,
typically, occur is 30–50 Å.
31,32
Fig. 2(a) clearly shows that the alignment of the energy levels
of P3HT, CdSe and WS
2
is very favorable for this kind of
transfer. In the same way CdSe QDs and P3HT emission and
WS
2
nanotube absorption peaks, Fig. 2(b), nicely overlap satis-
fying the requirement for Fo
¨rster energy transfer. Indeed the
WS
2
NTs can in principle efficiently absorb at around 650 nm
both the less intense light emitted by the P3HT and the more
intense light emitted by the CdSe QDs. Because the NTs are not
emissive the occurrence of this energy transfer could not be
Fig. 1 (a) P3HT and WS
2
energy levels. Pure P3HT, WS
2
and P3HT–WS
2
blends. (b) UV-Vis absorption spectra and (c) emission spectra after excitation
at 400 nm. The intensity of the excitation beam is kept constant for all the
measurements and the data are corrected for the number of absorbed
photons.
PCCP Paper
Published on 15 July 2014. Downloaded by Imperial College London Library on 05/09/2014 06:18:09.
View Article Online
This journal is ©the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 17998--18003 | 18001
probed directly by monitoring the WS
2
emission as is usually
done.
10,33
In an equivalent way we have proceeded to monitor
the changes of the static and dynamic emission of the QDs
induced by the presence of different amounts of NTs.
In Fig. 3(a) the normalized static emissions of the pure QDs,
pure P3HT and the P3HT–CdSe blends are reported. For the
blends and the polymer the normalization was done with
respect to the P3HT peak emission at 720 nm. This value has
been chosen because at this wavelength there is no contribu-
tion from the QDs or the nanotubes. As expected, the emission
from the QDs becomes more pronounced as their concen-
tration in the blends increases, as indicated by the arrow in
Fig. 3(a) even if the total emission is quenched (data are reported
in the ESIsection).
26–35
Interestingly, upon increasing the QD
concentrations in the blend the emission peak linearly blue shifts
from 660 nm, the position of the pure P3HT peak, to 640 nm,
where the bare CdSe emission is centered (data are reported in
the ESIsection). At this point we fixed the CdSe amount with
respect to the P3HT in the blend at 50 wt% and varied the
NT concentration. In Fig. 3(b) the emission spectra of the
P3HT–CdSe blend without NTs (namely sample CdSe–WS
2
0)
together with all the ternary blends P3HT–CdSe–WS
2
containing
different amounts of nanotubes (from 9 to 33%), but the same
amount of polymer and QDs, are reported. Also in this case the
spectra are shown to be normalized with respect to the polymer
emission at 720 nm. It is clear that the QD emission peak is
strongly reduced by the presence of the nanotubes. This result
suggests that the emission of the QDs is reabsorbed by the NTs
and so an energy transfer process is efficiently taking place
between the CdSe and the WS
2
nanotubes. Therefore, following
the exciton dynamics is a powerful method for evaluating these
processes inside the blends. In order to probe the nature of this
transfer in greater detail and evaluate the quenching effects,
also in the lifetime of the exciton, time resolved fluorescence
measurements have been performed on the ternary blend samples
containing different amounts of WS
2
nanotubes. Energy transfer
processes occur typically on time scales ranging from picoseconds
to nanoseconds for singlet energy transfer.
The study of the fluorescence signals in the femtosecond (fs)
to picosecond (ps) time range allows following the dynamics of
the fluorescent excitons generated immediately after absorption
right up until charge separation. The fluorescence decay time
strongly depends on the nanocomposition of the blends and the
energy transfer process inside the blends.
The emission decays have been measured exciting the
sample at 400 nm and collecting the fluorescence decays at
650 nm where the contribution of both the polymer and the
QDs is present. The fluorescence decays for the ternary blend
samples are reported in Fig. 4(a) with WS
2
contents varying
from 9 wt% to 33 wt% with respect to the P3HT–CdSe 50 wt%
blend (indicated as CdSe–WS
2
_0). Pristine P3HT has also been
reported for clarity.
It can be observed that the fluorescence decays are faster in
the ternary blends with respect to pure P3HT, and the lifetimes
decrease for increasing concentrations of nanotubes in the blends.
Fig. 2 (a) P3HT, CdSe and WS
2
energy levels and (b) CdSe and P3HT
emission spectra and WS
2
absorption peak. Fig. 3 Emission spectra normalized at 720 nm after excitation at 400 nm
of (a) P3HT–CdSe blends and (b) P3HT–CdSe–WS
2
blends.
Paper PCCP
Published on 15 July 2014. Downloaded by Imperial College London Library on 05/09/2014 06:18:09.
View Article Online
18002 |Phys. Chem. Chem. Phys., 2014, 16, 17998--18003 This journal is ©the Owner Societies 2014
It appears that for concentrations higher than 27 wt% the
values of lifetimes reach a saturation level.
The kinetics is well described by a double exponential decay,
a fast and a slow phase correlated with different phenomena as
described in ref. 34. Indeed the emission at this wavelength
(650 nm) is given by a contribution from both CdSe QDs and
P3HT polymer emission. The results of the double fitting procedure
(starting from the maximum of the signal, thus excluding processes
occurring within the response function) are reported in Fig. 4(b).
Lifetimes rapidly decrease as a function of nanotube concentration
from about 100 to 40 ps. These results indicate that the quenching
effect is most probably due to an energy transfer between the QDs
and the WS
2
nanotubes.
4 Conclusions
In this work we have reported unprecedented fluorescence
quenching effects in the hybrid ternary system P3HT–CdSe–WS
2
due to WS
2
NT addition to P3HT–CdSe QD blends. Static and
time-resolved fluorescence decays give a strong indication of an
efficient energy transfer process from the QDs to the WS
2
NTs in
the ternary blend, responsible for fluorescence quenching. On the
other hand, in the binary blend P3HT–WS
2
no fluorescence
quenching has been detected suggesting the absence of energy
or electron transfer processes in this system. The evidence of
energetic interaction between WS
2
NTs and QDs, observed in
purely inorganic systems,
23
and here established for the first time
in hybrid nanocomposites with P3HT opens new possible fields of
application to metal dichalcogenide nanomaterials, which are
non-toxic and nowadays available in large amounts. Moreover,
our results provide new insights into the physics of hybrid blends
based on conjugated polymers and QDs and promising improve-
ments of the corresponding hybrid solar cell performances.
Acknowledgements
The authors wish to thank Prof. R. Tenne for useful discussion
and suggestions and for reading the manuscript. MIUR Public–
Private Laboratory Project (PO2_00556_33069378 RELIGHT) and
the COST Action MP0902 Composites of Inorganic Nanotubes
and Polymers (COINAPO) are acknowledged for partial financial
support.
Notes and references
1(a) J. Liu, T. Tanaka, K. Sivula, A. P. Alivisatos and
J. M. J. Fre
´chet, J. Am. Chem. Soc., 2004, 126, 6550;
(b) Y. Zhou, Y. Li, H. Zhong, J. Hou, J. Y. Ding, C. Yang
and Y. Li, Nanotechnology, 2006, 17, 4041; (c) M. Afzaal and
P. O’Brien, J. Mater. Chem., 2006, 16, 1597; (d) M. Brinkmann,
D. Aldakov and F. Chandezon, Adv. Mater., 2007, 19, 3819;
(e)S.KumarandG.D.Scholes,Microchim. Acta, 2008, 160, 315.
2(a) R. Søndergaard, M. Ho
¨sel, D. Angmo, T. T. Larsen-Olsen
and F. C. Krebs, Mater. Today, 2012, 15, 36; (b) A. Bruno,
F. Villani, I. A. Grimaldi, F. Loffredo, P. Morvillo, R. Diana,
S. Haque and C. Minarini, Thin Solid Films, 2014, 560, 14–19.
3M.Jo
¨rgensen, K. Norman and F. C. Krebs, Sol. Energy Mater.
Sol. Cells, 2008, 92, 686.
4 I. Gur, N. A. Fromer, M. L. Geier and A. P. Alivisatos, Science,
2005, 310, 462.
5 L. Qian, J. Yang, R. Zhou, A. Tang, Y. Zheng, T.-K. Tseng,
D. Bera, J. Xue and P. H. Holloway, J. Mater. Chem., 2011,
21, 3814.
6 P. M. Sommeling, M. Spa
¨th, H. J. P. Smith, N. J. Bakker and
J. M. Kroon, J. Photochem. Photobiol., A, 2004, 164, 137.
7 B. Paci, G. D. Spyropoulos, A. Generosi, D. Bailo, V. Rossi
Albertini, E. Stratakis and E. Kymakis, Adv. Funct. Mater.,
2011, 21, 3573.
8 M. C. Beard, A. G. Midgett, M. C. Hanna, J. M. Luther,
B. K. Hughes and A. J. Nozik, Nano Lett., 2010, 10, 3019.
9 T. Xu and Q. Qiao, Energy Environ. Sci., 2011, 8, 2700;
L. Martinez, A. Stavrinadis, S. Higuchi, A. L. Diedenhofen,
M. Bernechea, K. Tajima and G. Konstantatos, Phys. Chem.
Chem. Phys., 2013, 15, 5482.
10 M. Thomalla and H. Tributsch, J. Phys. Chem. B, 2006,
110, 12167.
11 R. Tenne, L. Margulis, M. Genut and G. Hodes, Nature, 1992,
360, 444.
12 (a) A. Rothschild, J. Sloan and R. Tenne, J. Am. Chem. Soc.,
2000, 122, 5169; (b) A. Zak, L. Sallcan-Eckera, A. Margolin,
M. Genut and R. Tenne, Nano, 2009, 4, 91.
Fig. 4 (a) Normalized fluorescence decays and (b) lifetimes vs. WS
2
wt% in
the ternary blend.
PCCP Paper
Published on 15 July 2014. Downloaded by Imperial College London Library on 05/09/2014 06:18:09.
View Article Online
This journal is ©the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 17998--18003 | 18003
13 (a)J.Kaplan-Ashiri,S.R.Cohen,K.Gartsman,V.Ivanovskaya,
T. Heine, G. Seifert, I. Wiesel, H. D. Wagner and R. Tenne,
Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 523.
14 (a) E. Zohar, S. Baruch, M. Shneider, H. Dodiuk, S. Kenig,
R. Tenne and H. D. Wagner, J. Adhes. Sci. Technol., 2011,
25, 1603; (b) M. Naffakh, M. Remskar, C. Marco, M. A. Gomez-
Fatou and I. Jimenez, J. Mater. Chem., 2011, 21, 3574.
15 M. Shneider, L. Rapoport, A. Moshkovich, H. Dodiuk,
S. Kenig, R. Tenne and A. Zak, Phys. Status Solidi A, 2013,
210, 2298.
16 W. Zhang, S. Ge, Y. Wang, M. H. Rafailovich, O. Dhez,
D. A. Winesett, H. Ade, Kurikka V. P. M. Shafi, A. Ulman,
R. Popovitz-Biro, R. Tenne and J. Sokolov, Polymer, 2003,
44, 2109.
17 H. Dodiuk, O. Kariv, S. Kenig and R. Tenne, J. Adhes. Sci.
Technol., 2014, 28, 38.
18 A. Voldman, D. Zbaida, H. Cohen, G. Leitus and R. Tenne,
Macromol. Chem. Phys., 2013, 214, 2007.
19 T. Di Luccio, C. Borriello, A. Bruno, M. G. Maglione,
C. Minarini and G. Nenna, Phys. Status Solidi A, 2013, 210, 2278.
20 (a) A. Bruno, C. Dyer-Smith, L. X. Reynolds, J. Nelson and
S. A. Haque, J. Phys. Chem. C, 2013, 117, 19832; (b) A. Bruno,
C. Borriello, T. Di Luccio, G. Nenna, L. Sessa, S. A. Haque
and C. Minarini, J. Nanopart. Res., 2013, 15, 2085.
21 A. A. Y. Guilbert, L. X. Reynolds, A. Bruno, A. MacLachlan,
S. P. King, M. A. Faist, E. Pires, J. E. Macdonald, N. Stingelin,
S. A. Haque and J. Nelson, ACS Nano, 2012, 6, 3868.
22 M. Dante, J. Peet and T. Q. Nguyen, J. Phys. Chem. C, 2008,
112, 7241.
23 R. Kreizman, O. Schwartz, Z. Deutsch, S. Itzhakov, A. Zak,
S. R. Cohen, R. Tenne and D. Oron, Phys. Chem. Chem. Phys.,
2012, 14, 4271.
24 (a) J. Haremza, M. Hahn, T. Krauss, S. Chen and J. Calcines,
Nano Lett., 2002, 2, 1253; (b) S. Banerjee and S. Wong, Nano
Lett., 2002, 2, 195; (c) M. Olek, T. Buesgen, M. Hilgendorff
and M. Giersig, J. Phys. Chem. B, 2006, 110, 12901.
25 S. Jeong, H. Shim, S. Kimand and C. Han, ACS Nano, 2009,
4, 324.
26 G. Grancini, M. Biasiucci, R. Mastria, F. Scotognella,
F. Tassone, D. Polli, G. Gigli and G. Lanzani, J. Phys. Chem.
Lett., 2012, 3, 517.
27 M. D. Heinemann, K. von Maydell, F. Zutz, J. Kolny-Olesiak,
H. Borchert, I. Riedel and J. Parisi, Adv. Funct. Mater., 2009,
19, 3788.
28 J. J. Benson-Smith, J. Wilson, C. Dyer-Smith, K. Mouri,
S. Yamaguchi, H. Murata and J. Nelson, J. Phys. Chem. B,
2009, 113, 7794.
29 J. J. Benson-Smith, H. Ohkita, S. Cook, J. R. Durrant,
D. D. C. Bradley and J. Nelson, Dalton Trans., 2009, 10000.
30 S. Westenhoff, I. A. Howard, J. M. Hodgkiss, K. R. Kirov,
H. A. Bronstein, C. K. Williams, N. C. Greenham and
R. H. Friend, J. Am. Chem. Soc., 2008, 130, 13653.
31 (a) G. Bernardo, A. Charas and J. Morgado, J. Phys. Chem.
Solids, 2010, 71, 340; (b) M. Ichikawa, T. Aoyama, J. Amagai,
T. Koyama and Y. Taniguchi, J. Phys. Chem. C, 2009, 113, 11520.
32 G. Y. Zhong, J. He, S. T. Zhang, Z. Xu, Z. H. Xiong, H. Z. Shi,
X. M. Ding, W. Huang and X. Y. Hou, Appl. Phys. Lett., 2002,
80, 4846.
33 M. Ichikawa, T. Aoyama, J. Amagai, T. Koyama and Y. Taniguchi,
J. Phys. Chem. C, 2009, 113, 11520.
34 L. X. Reynolds, T. Lutz, S. Dowland, A. MacLachlan, S. Kinga
and S. A. Haque, Nanoscale, 2012, 4, 1561.
35 A. Bruno, T. Di Luccio, C. Borriello, F. Villani, S. A. Haque
and C. Minarini, Energy Procedia, 2014, 44, 167.
Paper PCCP
Published on 15 July 2014. Downloaded by Imperial College London Library on 05/09/2014 06:18:09.
View Article Online
... It is easily available at a low cost and is less hazardous than other transition metal dichalcogenides (TMDC) compounds. The development of WS 2 in thin film solar cells is still in its infancy compared with other photovoltaic materials [20]. Due to its superior optoelectronic properties, tungsten disulfide (WS 2 ) has become the primary material for thin film solar cells. ...
... Its tunable bandgap is an essential feature that is usually ignored. WS 2 has a large direct bandgap (>2 eV) and a small indirect bandgap (~1.3 eV) [20][21][22]. ...
... The potential of these materials, such as WS 2 and P3HT, to act as electron and hole transport layers in thin film solar cells could be one of the reasons for the high-power conversion efficiency [20][21][22]. Another reason could be optical effects caused by multireflection within the device between layers, which could result in the active layer absorbing more photons and producing larger short-circuit currents [50]. ...
Article
Full-text available
In the present work, lead-free perovskite solar cell has been structured using CH 3 NH 3 SnI 3 as an absorber layer, P3HT acting as a HTL, and TiO 2 as ETL material. Perovskite solar cell is the originating photovoltaic technology that shows great elevation in its performance during recent years. The fundamental n-i-p planar heterojunction structure of photovoltaic cells has been designed and simulated with solar cell capacitance simulation software (SCAPS-1D). In this study, different parameters like thickness, acceptor density, temperature, and defect density have been varied to increase the device performance. Optimum values of different parameters have been used to attain the good results of the photovoltaic device such as PCE, V OC , FF, and J SC of 27.54%, 1.0216 V, 86.56%, and 31.14 mA/cm ² , respectively. The impact of acceptor density has been varied from 1x10 ⁻¹² cm ⁻³ to 1x10 ⁻²⁰ cm ⁻³ for the proposed device structure. Therefore, the PCE of this device structure increases by using different charge transport materials. This simulation study shows that the proposed cell structure can be used to construct the photovoltaic cell with higher efficiency.
... It is easily available at a low cost and is less hazardous than other transition metal dichalcogenides (TMDC) compounds. The development of WS 2 in thin film solar cells is still in its infancy compared with other photovoltaic materials [20]. Due to its superior optoelectronic properties, tungsten disulfide (WS 2 ) has become the primary material for thin film solar cells. ...
... Its tunable bandgap is an essential feature that is usually ignored. WS 2 has a large direct bandgap (>2 eV) and a small indirect bandgap (~1.3 eV) [20][21][22]. ...
... The potential of these materials, such as WS 2 and P3HT, to act as electron and hole transport layers in thin film solar cells could be one of the reasons for the high-power conversion efficiency [20][21][22]. Another reason could be optical effects caused by multireflection within the device between layers, which could result in the active layer absorbing more photons and producing larger short-circuit currents [50]. ...
Article
Full-text available
Solar cells based on lead-free perovskite have demonstrated great potential for next-generation renewable energy. The SCAPS-1D simulation software was used in this study to perform novel device modelling of a lead-free perovskite solar cell of the architecture ITO/WS2/CH3NH3SnI3/P3HT/Au. For the performance evaluation, an optimization process of the different parameters such as thickness, bandgap, doping concentration, etc., was conducted. Extensive optimization of the thickness and doping density of the absorber and electron transport layer resulted in a maximum power-conversion efficiency of 33.46% for our designed solar cell. Because of the short diffusion length and higher defect density in thicker perovskite, an absorber thickness of 1.2 µm is recommended for optimal solar cell performance. Therefore, we expect that our findings will pave the way for the development of lead-free and highly effective perovskite solar cells.
... The upconversion CQD easily excites the lower energy photons and emits the higher energy photons. Mostly, organic polymers are used as donor molecules in organic solar cell (OSC) 5 applications. However, the PANI cannot be used for OSC because the band gap range is < 3.5 eV, it only absorbs visible regions (higher energy). ...
Article
Carbon quantum dots (CQDs) play a significant role in various applications, such as solar cells, bio-imaging, batteries, sensing, and therapy, thanks to their outstanding optical and electrical properties.
... In another work, a binary hybrid of the polymer poly(3 hexylthiophene) (P3HT) and CdSe quantum dots was prepared to which WS 2 nanotubes were added [79]. Energy transfer from the quantum dots to the nanotubes led to substantial fluorescence quenching, which was documented via transient fluorescence measurements. ...
Article
The birth of nanoscience more than 50 years ago fueled the renaissance in layered materials research leading to many materials discoveries with unprecedented scientific and technological impacts. Following the early reports on carbon fullerenes and nanotubes, the discovery of inorganic one-dimensional (1D) nanotubes and zero-dimensional (0D) fullerenes created a major playground for new physicochemical observations. The meteoric rise of two-dimensional (2D) materials in concert set off outstanding advances in the synthesis and manipulation of layered materials with atomic precision. This review identifies new directions in materials science that emerge through integrating the two layered systems—2D with inorganic 1D and 0D. Summarizing the key developments in the two distinct nanomaterials families, we highlight preliminary instances of integrating them into functional nanostructures. A few gedankenexperiments regarding prospective applications of the integrated system are then introduced to stimulate further experimental and theoretical investigations that can potentially result in unforeseen scientific observations.Graphical abstract
... Also, a significant increase in PCE can be up to 4.62% by adsorbing PbS QDs on ZnO nanorod arrays [17]. Moreover, 1D component provides a direct electron transport path to reduce the recombination rate [9,18,19]. In a recent report, the interfacial recombination rate of CdSe-QD/TiO 2 -nanowire heterostructure solar cells can be depressed effectively by in situ interface engineering [3]. ...
Article
Full-text available
0D-1D mixed-dimensional van der Waals (MvdW) heterostructures have shown great potential in electronic/optoelectronic applications. However, addressing the interface barrier modulation and charge-transfer mechanisms remain challenging. Here, we develop an analytic model to illustrate the open-circuit voltage and charge-transfer state energy in PbSxSe1-x-quantum dots (QDs)/MoS2-nanotube (NT) 0D-1D MvdW heterostructures based on atomic-bond-relaxation approach, Marcus theory and modified-detailed balance principle. We find that the band alignment of PbSxSe1-x-QDs/MoS2-NT heterostructures undergoes a transition from type II to type I, and the threshold of size is around 5.6 nm for x=1, which makes the system suitable for various devices including photocatalytic device, light-emission device and solar cell under different sizes. Our results not only clarify the underlying mechanism of interfacial charge-transfer in the heterostructures, but also provide unique insight and new strategy for designing multifunctional and high-performance 0D-1D MvdW heterostructure devices.
... 97 Efficient photocatalytic reduction of CO 2 in aqueous suspension containing the hybrid nanoparticles was reported. In one work, 102 hybrid constructs consisting of the organic polymer poly(3-hexylthiophene) (P3HT) and quantum dots (QD) of CdSe and WS 2 nanotubes were assembled. The carrier transfer was studied by following the quenching of the QD photoluminescence. ...
Article
Full-text available
The global impact of carbonaceous emissions from the internal combustion engine and coal-fired power plants has stimulated efforts to mitigate global warming and deterioration of the habitable biosphere. These efforts resulted in the maturation of new technologies to harvest and store the natural energies available-wind, waves, geothermal, sun and outer space radiation - for conservation and security. Inorganic layered compounds (2D materials), like MoS2, TiS2 and CoO2 played a major role in the upbringing of novel energy technologies, which can one day replace fossil fuel in the transport industry as well as in other energy-consuming sectors. In this perspective, the history of various concepts explored in energy-related research using bulk layered compounds (2D materials), is briefly reviewed, first. The recent addition of 2D nanostructures, in energy conversion and storage, has added significant momentum to this research field. Particularly, this perspective places a great emphasis on the study of inorganic fullerene-like nanoparticles and nanotubes from 2D materials, and to some extent also on the atomically thin single layer solids, that apparently surpasses the performance of other respective allotropes in the pursuit of their use in energy storage/conversion, catalysis, and sensing.
... The literature values for the CB (VB) edge positions for CdSe and CsPbBr 3 cubic phase thin films vary between −3.4 eV (−5.8 eV) and −3.5 eV (−5.8 eV), respectively, with respect to the vacuum level. [26][27][28][29][30] Our data for the native OA capped hybrid system are in line with the above, but interestingly for the Gly and PABA hybrid systems, values are significantly different, suggesting that a pronounced field across the interface and nearly zero band offsets evolved, as described in detail below. Figure 3 illustrates the energy level modifications upon hybridization. ...
Article
Formation of a p-n junction-like with a large built-in field is demonstrated at the nanoscale, using two types of semiconducting nanoparticles, CsPbBr3 nanocrystals and CdSe nanoplatelets, capped with molecular linkers. By exploiting chemical recognition of the capping molecules, the two types of nanoparticles are brought into mutual contact, thus initiating spontaneous charge transfer and the formation of a strong junction field. Depending on the choice of capping molecules, the magnitude of the latter field is shown to vary in a broad range, corresponding to an interface potential step as large as ∼1 eV. The band diagram of the system as well as the emergence of photoinduced charge transfer processes across the interface is studied here by means of optical and photoelectron based spectroscopies. Our results propose an interesting template for generating and harnessing internal built-in fields in heterogeneous nanocrystal solids.
Article
Full-text available
In the ever-increasing energy demand scenario, the development of novel photovoltaic (PV) technologies is considered to be one of the key solutions to fulfil the energy request. In this context, graphene and related two-dimensional (2D) materials (GRMs), including nonlayered 2D materials and 2D perovskites, as well as their hybrid systems, are emerging as promising candidates to drive innovation in PV technologies. The mechanical, thermal, and optoelectronic properties of GRMs can be exploited in different active components of solar cells to design next-generation devices. These components include front (transparent) and back conductive electrodes, charge transporting layers, and interconnecting/recombination layers, as well as photoactive layers. The production and processing of GRMs in the liquid phase, coupled with the ability to "on-demand" tune their optoelectronic properties exploiting wet-chemical functionalization, enable their effective integration in advanced PV devices through scalable, reliable, and inexpensive printing/coating processes. Herein, we review the progresses in the use of solution-processed 2D materials in organic solar cells, dye-sensitized solar cells, perovskite solar cells, quantum dot solar cells, and organic-inorganic hybrid solar cells, as well as in tandem systems. We first provide a brief introduction on the properties of 2D materials and their production methods by solution-processing routes. Then, we discuss the functionality of 2D materials for electrodes, photoactive layer components/additives, charge transporting layers, and interconnecting layers through figures of merit, which allow the performance of solar cells to be determined and compared with the state-of-the-art values. We finally outline the roadmap for the further exploitation of solution-processed 2D materials to boost the performance of PV devices.
Article
Upconversion materials play a significant role in different applications like bio-imaging, solar cells, and therapy treatment due to their light conversion behavior (lower energy photon to higher energy photon). This material is a doping material for the improvement of the light-harvesting properties of polymers. However, there are limited publications in this field. Deposited/fabricated upconversion carbon quantum dot doped polyaniline photoactive film by electropolymerization. First, we prepared the quantum dot from the source of ascorbic acid and studied the quantum dot optical and structural properties by ultraviolet(UV), fluorescence (FL), and high-resolution transmission electron microscopy (HR-TEM), Fourier transform infrared (FT-IR), spectroscopy. The quantum dot is doped with polyaniline (PANI) using the fluorine-doped tin oxide (FTO) plate for making a film for this research. The prepared film has been analyzed spectroscopically for photophysical properties (absorption, transmission, band gap), surface morphology, and thermal stability of this merit to the solar cell. The current density and impedance spectra are have well supported by such a material characterization that might be used for the application of organic solar cells. Finally, the Förster resonance energy transfer of the quantum dot (donor) to the polyaniline (acceptor) is confirmed. With the help of the above characterization, results are supporting the organic solar cell.
Article
We investigated the photophysical interactions between CdSe/CdS quantum dots (QDs) and a conjugated polymer (CP, P3HT). The photoluminescence intensity of P3HT in the QDs/P3HT hybrid system is significantly enhanced compared to that of the neat P3HT system. We found via transient absorption spectroscopy that the energy-level differences at the interfaces between P3HT and QDs resulted in delayed relaxation-dynamics of the P3HT singlet (S1)-excitons and suppressed polaron-formation. Thus, the radiative recombination of the S1 excitons occurs frequently in the hybrid system than in the neat P3HT system. Our findings on the CP-based hybrid system may provide important information to improve the efficiencies of optoelectronic devices, such as organic light-emitting diodes.
Article
Full-text available
We present a spectroscopic investigation on a new hyperbranched cadmium selenide nanocrystals (CdSe NC)/poly(3-hexylthiophene) (P3HT) blend, a potentially good active component in hybrid photovoltaics. Combined ultrafast transient absorption spectroscopy and morphological investigations by means of an ultrafast confocal microscope reveal a strong influence of the complex local structure on the photogenerated carrier dynamics. In particular, we map the electron-transfer process across the hybrid NC/polymer interface, and we reveal that charge separation occurs through a preferential pathway from the CdSe nanobranches to the P3HT chains. Efficient charge generation at the distributed heterojunction is also confirmed by scanning kelvin probe force microscopy measurements.
Article
Full-text available
The prospect of exploiting quantum dots (QDs) properties (tunable absorption spectrum, multiple exciton generation) while maintaining the flexible structure of polymer systems opens new possibilities in the photovoltaic field. Although charge transport dynamics in pristine polymer and QDs systems have been quite well established lately, a complete understanding of the charge transfer process between QDs and polymers when they are in blends is still lacking. In this work we used static and ultrafast fluorescence spectroscopy together with Atomic force Microscopy (AFM) to study the exciton dynamics in polymer/QDs films. Specifically we used poly(3-hexylthiophene) (P3HT) as the hole conducting donor material and the core shell CdSe(ZnS) QDs as the electron acceptor material. The QDs surface has been treated with two different capping ligands treatments: one based on the use of pyridine and the other one on hexanoic acid. The influence of the two different methods on the exciton dynamics and on the morphology will also be discussed. Blends containing differently treated P3HT/CdSe(ZnS) wt% ratios have been prepared producing films having uniform morphology and good intermixing, as proved by AFM measurements. Ultrafast fluorescence decays allowed us to compare the exciton dynamics in the polymer pristine respect to the treated P3HT/CdSe(ZnS) films. Efficient fluorescence quenching has been shown by both kind of blends respect to the pure polymer.
Article
Full-text available
In this work, we studied energetic and optical proprieties of a polyester-containing oxadiazole and carbazole units that we will indicate as POC. This polymer is characterized by high photoluminescence activity in the blue region of the visible spectrum, making it suitable for the development of efficient white-emitting organic light emission devices. Moreover, POC polymer has been combined with two red emitters InP/ZnS quantum dots (QDs) to obtain nanocomposites with wide emission spectra. The two types of QDs have different absorption wavelengths: 570 nm [InP/ZnS(570)] and 627 nm [InP/ZnS(627)] and were inserted in the polymer at different concentrations. The optical properties of the nanocomposites have been investigated and compared to the ones of the pure polymer. Both spectral and time resolved fluorescence measurements show an efficient energy transfer from the polymer to QDs, resulting in white-emitting nanocomposites.
Conference Paper
Thermoset-elastomer polyurethane (PU) nanocomposites were prepared using two types of tungsten disulphide (WS2) nanoparticles: inorganic fullerene-like (IF) and inorganic nanotubes (INT) through an in-situ polymerization process. The quality of dispersion was evaluated using scanning electron microscope (SEM) and the thermomechanical properties were analyzed using dynamic mechanical analysis (DMA). Addition of 1% and 3 wt.%. IFs resulted in enhancement in storage modulus of 45 and 100%, respectively, compared to the neat polymer. The enhancement using only 0.5% wt. INTs was more than 100% and then decreased with additional amount of nanotubes. While no significant change in the composite’s glass transition temperature (Tg) was observed with IF-WS2, 0.5% INT-WS2 showed a 20 °C increase of Tg. In both cases, analysis of the chemical structure using an attenuated total reflectance- Fourier transform infrared spectroscopy showed no effect of the nanoparticles on the chemical structure of the PU and wide-angle X-ray diffraction showed no change in morphology. In the case of IF-WS2 the highest peel strength was obtained with 1% wt. demonstrating a 44% improvement in peel strength. However, in the case of INT-WS2, incorporating 0.5% wt. improved the peel strength by more than 1000%. SEM analysis showed a unique development of a nodular morphology and a failure mechanism dominated by nanotube pull-out. It was concluded that the geometry of the nanoparticles (nanotubes or fullerene) has a dominant effect on the final PU nanocomposite properties.
Article
Recently large amounts of inorganic nanotubes (INT) and inorganic fullerene-like (IF) nanoparticles of WS2 became available and methods for their dispersion in different media were developed. In the present work the tribological properties of epoxy composite compounded with tungsten disulfide particles of different sizes and morphologies, including quasi-spherical IF nanoparticles, one-dimensional INT as well as micron-size platelets (2H) were investigated. The coefficient of friction and wear loss were measured under dry contact conditions using different tribological rigs. Remarkable reduction in wear and also friction (under high load) was demonstrated for the IF/INT epoxy nanocomposite. The reduced wear is attributed in general to the reinforcement of the polymer matrix by nanoparticles and the simultaneous reduction of the epoxy brittleness. Contrarily, the friction of the neat epoxy sample and epoxy mixed with platelets was accompanied with strong wear and transfer of a polymer film onto the rubbed surfaces. These results are consistent with the recently reported improvements in the fracture toughness, peel and shear strength of the epoxy-nanoparticles (IF/INT) composites.
Article
Emission quenching is studied in systems composed of [9,9-dioctylfluorene-co-N-(4-butylphenyl)-diphenylamine] (TFB) and various concentrations of three different types of acceptors: [6,6]-phenyl-C61 butyric acid methyl ester (mono-PCBM) and its multiadduct analogues bis-PCBM and tris-PCBM. We find that the degree of emission quenching for a given fullerene concentration decreases as the PCBM adduct number increases, while the microstructure, as observed with transmission electron microscopy, becomes more coarse. These observations are rationalized in terms of possible differences in the miscibility of fullerenes in the polymer and different excitation dissociation rates. We also extract a value for the exciton diffusion length in TFB of 9.0 ± 2 nm from ultrafast fluorescence decay measurements. The results have been confirmed with independent measurements.
Article
The growth mechanism of WS2 nanotubes in the large-scale fluidized-bed reactor is studied in greater detail. This study and careful parameterization of the conditions within the reactor lead to the synthesis of large amounts (50–100 g/batch) of pure nanotubes, which appear as a fluffy powder, and (400–500 g/batch) of nanotubes/nanoplatelets mixture (50:50), where nanotubes usually coming in bundles. The two products are obtained simultaneously in the same reaction but are collected in different zones of the reactor, in a reproducible fashion. The characterization of the nanotubes, which grow catalyst-free, by a number of analytical techniques is reported. The majority of the nanotubes range from 10 to 50 micron in length and 20–180 nm in diameter. The nanotubes reveal highly crystalline order, suggesting very good mechanical behavior with numerous applications.
Article
Polymer-inorganic nanoparticle composites are an area of great research interest. Inorganic fullerene-like (IF) structures and inorganic nanotubes (INT) are used for producing composite materials with specific goals of reinforcement of the polymer and ameliorating its thermal stability. Here, a composite material containing INT and conducting polymer (polyaniline (PANI)) is synthesized by an in situ oxidative polymerization of the monomer in the presence of WS 2 nanotubes. The structure and electrical behavior of PANI/INT composite is studied and compared with the results of pristine PANI (synthesized under the same conditions without INT). Most remarkably, INT-WS 2 are found to play the role of an active doping agent. The highest conductivity is obtained for 0.85 wt% (ca. 0.06 at%) nanotubes content, two orders of magnitude higher than that of pristine PANI. This work suggests a new approach to control the host–guest interaction in the polymer nanocomposite via (Lewis) acid–base equilibrium.
Article
Tungsten disulfide (WS 2) nanotubes are used to prepare polymer nanocomposites, similarly to other metal dichalcogenide materials, to improve lubricating and/or mechanical properties. In order to explore the possibility to extend these advantages to conductive polymers we realized new nanocomposites of WS2 nanotubes and polyfluorene conjugated polymer (WS2/PFO). The nanocomposites were prepared from solution processing at several nanotubes concentrations. The morphological and structural analyses by SEM and XRD proved that the density of nanotubes within the polymer increased according to the preparation conditions. The successful incorporation of WS2 nanotubes was also evidenced by UV–Vis absorbance spectroscopy. The WS2/PFO nanocomposites were tested in light emitting devices at relatively big load of nanotubes realizing a new class of devices with promising improved mechanical and thermal properties with out affecting substantially the device optoelectronic performances.