Sie sind auf Seite 1von 7

Structural and Electrical Properties of Nb-Doped Anatase TiO

2
Nanowires by Electrospinning
Panikar S. Archana,
z
Rajan Jose,
w,z,y
Tan Mein Jin,
z
Chellapan Vijila,
z
Mashitah M. Yusoff,
y
and
Seeram Ramakrishna
w,z,J
z
National University of Singapore, 117576 Singapore, Singapore
y
Faculty of Industrial Sciences and Technology (FIST), Universiti Malaysia Pahang, 26300 Pahang, Malaysia
z
Institute of Materials Research and Engineering (IMRE), A-STAR, 117602 Singapore, Singapore
J
King Saud University, Riyadh 11451, Kingdom of Saudi Arabia
One-dimensional nanostructures of niobium-doped anatase TiO
2
(Nb:TiO
2
) up to 5 at.% Nb were synthesized by electrospinning
a polymeric solution containing titanium and niobium precursors
and subsequent annealing. Thus obtained bers had diameter
B150 nm. The undoped TiO
2
bers were constituted by larger
single crystalline grains of size B50 nm, whereas the doped ones
had decreased grain sizes (B30 nm) under similar processing
conditions. The Nb doping decreased the BET surface area of
TiO
2
. A strain-induced lattice contraction was observed in
Nb:TiO
2
. The continuous nanobers were shortened to nano-
wires (NW) of aspect ratio 10:1 by ultrasonically dispersing them
in acetic acid, which were developed as lms of thicknessB813lm
onto conducting glass substrates. The TiO
2
and Nb:TiO
2
nano-
wire lms were further sensitized by a dye; the amount of dye
anchored was found to decrease with increase in the dopant
concentration. The dye-sensitized solar cells fabricated using the
doped bers, although with a nominally increased current
density (J
SC
), have reduced efciency due to lower ll factor
and open circuit voltage (V
OC
). The electron diffusion coef-
cient (D
n
) and mobility (l
n
) of the TiO
2
and Nb:TiO
2
NW in the
presence of iodide/triiodide ions were an order of magnitude
higher compared with the undoped samples.
I. Introduction
O
NE-DIMENSIONAL (1D) metal oxide nanostructures, such as
continuous nanobers (NF), large aspect ratio nanowires
(NW), and nanorods, are attracting much attention recently due
to their dependence of physical properties on directionality.
14
Most commonly NW are produced by the bottom-up approach,
i.e., by self-assembly of the target material from vapor, liquid, or
solid phases through nucleation and growth under optimized
conditions of temperature, pressure, or concentration.
1
Electro-
spinning is a top-down approach to produce continuous NF of
polymers. The process works under the principle of bending of a
charged liquid jet, usually a polymeric solution of optimum
viscosity, while passing through a uniform electric eld. In the
electrospinning technique, a polymeric solution is injected
through a syringe needle connected to a power supply. If the
force due to electric eld acting on the pendant drop at the
needle tip overcomes the forces due to viscosity and electric eld,
then a jet is formed which upon subsequent symmetric and
asymmetric bending produces solid continuous NF on a collec-
tor surface. If the polymeric solution contains metal ions
required for forming an inorganic solid, then appropriate post-
annealing produces continuous NF of reduced diameters than
the parent polymeric ber. Therefore, synthesis of metal oxide
nanostructures by electrospinning process recently gained
increasing attention due to the prospective of the process in
producing 1D nanostructures on a commercial scale.
58
A large
number of metal oxides are synthesized as NF by electrospin-
ning; an account of which is available in recent reviews.
911
The
electrospun metal oxide NF are potential candidates for a range
of applications such as sensing,
1215
lithium ion batteries,
1618
catalysis,
1922
and photovoltaics.
2332
Recent discovery of metallic conductivity and high optical
transparency in Nb-doped anatase TiO
2
(Nb:TiO
2
) stimulated
intense research on this material
33,34
for their applications as
transparent conducting oxides.
3540
The Nb
51
is an n-type dopant
for TiO
2
, modies the band structure of TiO
2
, and generates
additional carriers in its conduction band.
34,41
The band structure
calculations report that the 4d orbitals of Nb strongly hybridize
with the 3d ones of Ti and form a d-natured conduction band,
which is thought to be the origin of metallic conductivity in this
material.
34
Sheppard et al.
42
veried the metallic conductivity in
bulk Nb:TiO
2
by thermoelectric power measurements and
concluded that the properties of this material arise from defects
such as oxygen vacancies and Nb-doping in Ti sites. These defects
are compensated by either electrons or Ti interstitials. Lee et al.
37
doped 6 at.% Nb in anatase TiO
2
using conventional powder
method and used the resulting material as a compact layer in dye-
sensitized solar cells (DSCs). They observed that the photocurrent
density (J
SC
) is increased for the Nb:TiO
2
. Hasin et al.
40
synthe-
sized Nb:TiO
2
using a template assisted chemical method and
used them as counter electrodes for DSCs. They observed lower
charge transfer resistance and larger exchange current density
when the Pt sputtered Nb:TiO
2
was used as the counter electrode
in DSCs. Jose et al.
43
reported enhancement in J
SC
in DSCs
fabricated using Nb:TiO
2
microcrystals. Very recently Huang
and colleagues
44
developed nanocrystals of Nb:TiO
2
using a
chemical method; the DSCs fabricated using them gave B18%
higher photoelectric conversion efciency compared with the un-
doped ones. The 1D Nb:TiO
2
nanostructures developed by elect-
rospinning are also shown to have superior electrochemical
stability compared with the undoped samples.
45
In this paper, we report synthesis of 1D nanostructures of
Nb:TiO
2
using electrospinning technique and their structural
and electrical properties. Two dopant concentrations, viz. 2 and
5 at.% were studied, which is schematically shown in Fig. 1. The
S. Bhandarkarcontributing editor
This work was nancially supported by the National Research Foundation (grants
NRF-CRP4-2008-03 and NRF2007EWT-CERP01-0531), Singapore.
w
Author to whom correspondence should be addressed. e-mails: joserajan@ump.edu.my
and seeram@nus.edu.sg
Manuscript No. 27707. Received March 26, 2010; approved June 25, 2010.
J
ournal
J. Am. Ceram. Soc., 93 [12] 40964102 (2010)
DOI: 10.1111/j.1551-2916.2010.04003.x
r 2010 The American Ceramic Society
4096
2 at.% doping replaces one in 50 atoms of Ti by Nb and 5 at.%
replaces one in 20 atoms. We observed from the X-ray diffrac-
tion (XRD) studies that the lattice strain increases with the
dopant concentration that lead in a lattice contraction of the
resulting nanostructures. The optical bandgap of the Nb:TiO
2
increased with 5 at.% doping. The continuous NF of the
Nb:TiO
2
were shortened to NW of aspect ratio 10:1 and fabri-
cated DSCs. It was observed that the defects derived from dop-
ing play dominant role in their photovoltaic properties.
Nevertheless, up to 5 at.% Nb-doping in TiO
2
increases the
diffusion coefcient (D
n
) of electrospun NW by an order of
magnitude compared with that of the undoped samples. Our
results are discussed herewith.
II. Experimental Details
The TiO
2
NF were prepared by a commercial electrospinning
machine (NANON, MECC, Fukuoka, Japan) using a previously
adopted procedure.
46
The precursor solution for electrospinning
was prepared from polyvinyl acetate (PVAc, M
w
500000), dime-
thyl formamide (DMF), titanium (IV) isopropoxide, niobium
ethoxide (Sigma Aldrich, Singapore; 99.9%) and acetic acid. In a
typical synthesis, 4.5 g of PVAc was dissolved in 11.5 wt% DMF
together with titanium (IV) isopropoxide and niobium ethoxide.
The amounts of Ti and Nb precursors required were calculated
based on the number of atoms required for replacing a given
number of Ti atoms by Nb atoms. The sol was prepared by the
dropwise addition of 0.5 g of acetic acid. The resulting solution
was stirred well and electrospun at 25 kV accelerating voltage
and at 1 mL/h ow rate. The polymeric bers containing Ti
41
and Nb
51
ions were collected on a grounded rotating drum
placed B10 cm below the spinneret.
Decomposition of the composite ber and crystallization
behavior were studied by simultaneous differential thermal and the-
rmogravimetric analyzer (STA, DTG-60 AH, Shimadzu, Tokyo,
Japan) in the range 3018001C. The samples were sintered in air at
5001Cfor 24 h to remove PVAc and allownucleation and growth of
TiO
2
particles in the ber structure. The morphology of the as-spun
and annealed bers were examined by scanning electron microscopy
(SEM; Quanta 200 FEG System: FEI Company, Oregon) and
transmission electron microscopy (TEM; JEOL 2010Fas, Tokyo,
Japan). Crystal structure of the NF was examined by XRD (Sie-
mens D5005, Bruker, Karlsruhe, Germany) technique. The XRD
patterns were tted to anatase TiO
2
with additional Nb ions using
the PowderCell 2.4 program
47
for Rietveld analysis. Particulate
properties of the materials were studied using BET surface area
measurements (Micromeritics, Tristar 3000, Norcross). Presence of
Nb was conrmed by X-ray uorescence measurements (EDX-720,
Shimadzu). The optical bandgap of the bers were determined from
the absorption spectral measurements (UV-Visible-NIR Spectro-
photometer, UV-3600, Shimadzu). The annealed bers were ultra-
sonically dispersed in acetic acid to develop high aspect ratio
NWs.
48
Polyethylene oxide (0.05 g) and ethyl cellulose (0.025 g)
were added to the above dispersion and resulting paste was coated
on FTO glass substrates (1.5 cm1 cm; 25 O/&, Asahi Glass Co.
Ltd., Tokyo, Japan), duly spun coated with a thin layer (B100
nm) of TiO
2
, by doctor blade technique. The lms were then
annealed at 5001C for 1 h. The surface homogeneity of the sin-
tered lms was examined by SEM and surface prolometry. The
DSCs were prepared by soaking a 0.28 cm
2
TiO
2
NW electrode
in a dye solution containing D131 (received from Mitsubishi
paper mill, Tokyo, Japan)
24,25
in a 1:1 volume mixture of ace-
tonitrile and tert-butanol for 24 h at room temperature. The
dye-sensitized samples were then washed in ethanol to remove
unanchored dye and dried in air. Samples were sealed using a 50
mm spacer. Aetonitrile containing 0.1M lithium iodide, 0.03M
iodine, 0.5M 4-tert-butylpyridine, and 0.6M 1-propyl-2,3-
dimethyl imidazolium iodide was used as the electrolyte. A Pt
sputtered FTO glass was used as the counter electrode. Photo-
current measurements of the assembled DSCs were performed
using a solar simulator (San Ei, Osaka, Japan) at AM1.5G con-
dition. IV curves were obtained using a potentiostat (Autolab
PGSTAT30, Eco Chemie B.V., Utrecht, the Netherlands).
The electron transport in the DSCs was studied using
transient photocurrent technique. The DSCs were excited with
a low-intensity laser pulse (532 nm Nd:YAG laser, pulse width
o5 ns) superimposed on a large background white light illumi-
nation. The intensity of the white bias light was varied in order to
study the effect of photocarrier density on the D
n
. The intensity
of the laser light was controlled with neutral density lters to
keep the magnitude of photocurrent transients less than the dc
level due to the white bias light. The cells were illuminated
through the substrate as well as the counter electrode sides.
The photocurrent transients were recorded using a digital oscil-
loscope (Agilent, Santa Clara; 1 GHz) under short circuit con-
ditions. The RC time constant of the instrument was o20 ms.
Fig. 1. Schematic representation of the compositions studied. The a, b, and c represents the unit cell parameters of the supercell. The left gure is the
unit cell of anatase and the right one is a supercell showing 5 at.% Nb doping in TiO
2
.
December 2010 Structural and Electrical Properties of Nb-Doped Anatase TiO
2
Nanowires by Electrospinning 4097
III. Results and Discussion
Figure 2 shows the result of simultaneous thermal analysis of the
as-prepared bers. The as-spun composite polymeric bers
showed usual decomposition of the polymer and formation of
inorganic phase in the electrospun polymeric ber template.
Formation of metal oxide NF from composite polymeric bers
containing metal ions involve at least three steps, viz. (i) nucle-
ation of the crystals, (ii) growth, and (iii) directional mass trans-
port, in other words, sintering of the grains. The rst two
processes typically occur during growth of nanocrystals from
solutions. The polymeric ber containing the niobium ions
showed an endothermic peak in the differential thermal analy-
sis curve and a weight loss (B10%) in the thermogravimetric
analysis curve at B601C, which is attributed to the liberation of
the adsorbed solvents. Following the endothermic event at 601C,
a major exothermic event and substantial weight loss (40%) was
observed at B3001C, which results from the decomposition of
the polymer. The crystallization and growth of the anatase
phase were revealed by an exothermic reaction at starting at
B4001C; after which weight of the sample remains constant.
The SEM images (data not shown) showed that the bers
have average diameter of B150 nm. The bers were continuous
with smooth surface and consisted of grains of similar size.
Figure 3 shows the TEM and HRTEM images of the undoped
and doped samples. The undoped TiO
2
bers had grains of
larger diameter (B50 nm) as reported in our previous publica-
tion.
48
The doped bers were constituted by grains of lower di-
ameter (B1030 nm) implying that the doping reduced the grain
growth under similar processing conditions. Despite having sim-
ilar ber diameter, the BET surface area decreased with increase
in the Nb concentration. The BET surface area of the pure,
2 at.% Nb, and 5 at.% Nb-doped samples were 45, 33, 24 m
2
/g,
respectively. Consequently the pore size increased from B3 to
B4 nm upon 5 at.% doping. Reason for the lowering of surface
area for bers processed at similar conditions is assigned to the
increased pore size. It is worth mentioning that synthesis of
mesoporous Nb
2
O
5
with high surface area is a challenge till to-
day, probably due to the presence of the heavier Nb ions.
43
Presence of Nb in the samples was conrmed by XRF, XRD,
and HRTEM techniques. Figure 4 shows the XRF patterns of
samples containing Nb. The intensities of the X-ray lines, both
Ka and Kb, of Ti decreased and those of the Nb increased with
doping concentration. This observation indicates the Nb is
dissolved in the Ti sites as expected.
Figure 5 shows the XRD patterns of the TiO
2
and Nb:TiO
2
samples. A Rietveld renement plot of 5 at.% Nb in TiO
2
is in
Fig. 5(B). The samples retained the anatase structure; however,
a small peak corresponding to the rutile phase (volume fraction
o1%) was observed in the XRD pattern of the 5 at.% Nb in
TiO
2
sample. The tted crystal structure parameters such as
lattice parameters, unit cell volume, particle size, and lattice
strain are in Table I. The XRD line shape was affected by 2 at.%
Nb doping with apparently no shift in the lattice parameters.
Samples containing 5 at.% Nb showed a clear shift in the lattice
parameters as seen in Table I which is consistent with the DFT
calculations of Liu et al.
41
We note that the lattice strain
nominally increased with the Nb concentration. The observed
lattice strain is thought to cause the shift in the lattice parameters.
The strain induced lattice contraction in nanomaterials is not a
new phenomenon and widely discussed in the open literature. The
nucleation of the rutile phase also thought to arise from the lattice
strain. Strain-induced nucleation of the rutile phase in electro-
spun NF is reported in electrospun NF of reduced diameters.
46
Rietveld analysis of 5 at.% Nb:TiO
2
gave a site occupancy factor
of 0.0438 for Nb and 0.9561 for Ti ions, which closely matches to
the amount of precursor initially added in the reaction mixture.
The lattice strain induced packing defects were clearly observed in
the HRTEM patterns (Fig. 3) indicating the presence of Nb in the
anatase lattice. The size of the particles decreased from 50 nm for
the undoped samples to 30 nm for the doped samples.
Figure 6 shows the absorption spectra of the pure as well as
Nb:TiO
2
from which the optical bandgaps of the materials were
evaluated. The undoped TiO
2
gave a bandgap of 3.2 eV consis-
tent with the literature reports. The bandgap of 5 at.% doped
TiO
2
was B3.3 eV. The bandgap of the samples increased with
increasing doping which is consistent with the increased optical
transparency reported in Nb:TiO
2
thin lms.
35
Huang and col-
leagues also reported enhancement in the at band potential as a
result of Nb doping in TiO
2
.
All these characterizations lead to the conclusion that 1D
nanostructures of Nb:TiO
2
could be produced by electrospin-
ning a polymeric solution containing the precursors of titanium
and niobium and subsequent sintering. We shall now proceed to
discuss the JV characteristics of the DSCs fabricated using
them and electrical diffusivity of the resultant devices. The
SEM
48
and surface prolometric studies (data not shown)
show that surface of the sintered lms were homogeneous;
therefore, any variation in the measured photovoltaic proper-
ties is not arise from sample nonuniformity.
The dye loading, measured from the absorption spectra of the
desorbed dyes, found to decrease with increase in the Nb content.
Undoped samples gave a dye loading of 1.010
7
mol/cm
2
,
which decreased to 9.710
8
and 7.9 10
8
mol/cm
2
for 2 and
5 at.% Nb doping, respectively. Lowered dye loading with in-
crease in the dopant concentration is consistent with the BET
surface area measurements. Figure 7 shows the JV curves of
the DSCs fabricated using these devices. Photoelectric conver-
sion efciencies (Z) of the cells were calculated from J
SC
, open
circuit voltage (V
OC
), and ll factor (FF) determined from the
JV curves measured at AM 1.5 conditions. A standard silicon
cell certied at ISE Callab, Fraunhofer was used for light in-
tensity calibration. The DSCs fabricated using undoped TiO
2
showed relatively high open circuit voltage (0.750 V) with
J
SC
B8.9 mA/cm
2
, FFB70%, and ZB4.67%. The DSCs fabri-
cated using the Nb:TiO
2
showed nominal enhancement in the
J
SC
owing to the high concentrations of additional electrons due
to doping; however, the cells showed decreased efciencies due
to lower FF and V
OC
. The J
SC
of the 2 and 5 at.% Nb doped
TiO
2
were 9.74 and 10 mA/cm
2
, respectively. The J
SC
in highly
efcient (ZB11%) DSCs
49
is about 50% higher than the values
reported here. However, those DSCs were fabricated using
mesoporous particles of B2025 nm and with much larger spe-
cic surface area, i.e., about three times higher than that of the
present ones. Furthermore, the larger diameter of the present
ones additionally contributes to scattering rather than absorp-
tion. It is expected that doped bers with diameters in the 2030
nm range with enhanced crystallinity would increase the J
SC
considerably; therefore, such efforts are currently underway.
Decreased FF and V
OC
were observed for the doped samples
indicating increased carrier recombination consistent with the
ndings of Huang and colleagues
44
The recombination with the
Fig. 2. Simultaneous thermal analysis showing the differential
thermal analysis and thermogravimetric analysis curves of the as-spun
composite bers.
4098 Journal of the American Ceramic SocietyArchana et al. Vol. 93, No. 12
iodide ions in the electrolyte accounts majority of the photo-
electron loss in DSCs. Thermal scattering of the photoelectrons,
i.e., the phonon relaxation, is the typical loss occur in metal
oxides in DSCs.
43
However, in the present case, the doping in-
duced packing defects could also contribute to recombination.
The radii of Ti
41
and Nb
51
are 0.56 and 0.62 A

, respectively.
i.e., there is a mist of B11%for Nb
51
ions in the anatase lattice,
which induces lattice distortion as could be clearly visible in the
HRTEM images (Fig. 3). These packing defects are expected
to additionally trap and/or scatter the photoelectrons. This
argument predicts that W
61
with similar ionic radius to Ti
41
could be a reasonable choice such that packing defect-induced
Fig. 3. TEM and HRTEM images of (top panel) pure, (middle panel) 2 at.% Nb:TiO
2
, and (bottom panel) 5 at.% Nb:TiO
2
, respectively.
December 2010 Structural and Electrical Properties of Nb-Doped Anatase TiO
2
Nanowires by Electrospinning 4099
recombination could be neglected. Another source of charge
recombination in the metal oxide arises due to charge neutrality
upon doping. The charge neutrality when the Nb
51
doping in
the Ti
41
sites is achieved either by the creation of one Ti vacancy
per four Nb
51
ions doped or by the stoichiometric reduction of
Ti
41
to Ti
31
per Nb
51
doped. The holes thus produced are
expected to act as recombination centers thereby reducing the
FF and V
OC
.
To understand which of the above, i.e., scattering at defects
or recombination at the holes, is the dominant factor behind the
lower FF and V
OC
, we measured the differences in D
n
of the
doped and undoped samples. Scattering results in the loss of
kinetic energy of the electrons but they are still collected. The
V
OC
and hence FF will be lowered in the resulting devices due to
loss of electrical energy upon scattering. As charge diffusion
depends mainly on the concentration, the doped samples should
show enhanced D
n
. On the other hand, lost electrons are no
longer collected and should show a decrease in D
n
. Figure 8
shows the photocurrent transients observed for the three devices
fabricated using TiO
2
and Nb:TiO
2
NWs. The photocurrent was
observed immediately after laser excitation in the device fabri-
cated using TiO
2
and reaches the peak at around 40 ms followed
by an exponential decay. The rise times of the photocurrent
transients of the 2% and 5% Nb:TiO
2
were 20 and 6 ms,
respectively. The difference in the photocurrent rise time is at-
tributed to the difference in the thickness of the TiO
2
lms (see
caption of Fig. 8 for details). The t
c
was extracted by tting the
photocurrent transient using a single exponential decay. The D
n
was estimated using the relation, t
c
5d
2
/2.35D
n
, where d is the
thickness of the samples. The photoexcitation density for
various white light illumination intensities were calculated by
numerically integrating the photocurrent transients.
Figure 8(B) show the calculated D
n
from the photocurrent
transients. The D
n
systematically increased upon doping which
is attributed to the number of additional electrons generated in
the bottom of the conduction band.
34
For a photoexcitation
density of 3.43 10
16
(1.12 10
16
) cm
3
5%Nb:TiO
2
(TiO
2
)
gave D
n
of 3.84 10
3
(4.67 10
4
) cm
2
/s. The observed
enhancement in the D
n
of Nb:TiO
2
resulted from an order
of magnitude lowering in the charge collection time t
c
. These
observations indicate that scattering at the defects is the major
Fig. 4. XRF spectra of the Nb:TiO
2
bers. The TiKa and TiKb lines
are at 4.5006 and 4.9236 keV, respectively. The NbKa and NbKb lines
are at 16.6086 and 18.1648 keV, respectively.
Fig. 5. (A) XRD patterns of the TiO
2
and Nb:TiO
2
bers; (B) Rietveld
renement plot of a sample containing 5 at.% Nb in TiO
2
. The observed
spectra are shown as open circles. The solid lines are the simulated and
subsequently tted spectra. Bottom continuous line is the difference
between the experimental and simulated spectra. The vertical lines show
the peak positions of the doped sample.
Table I. Rened Crystal Structure Parameters of Undoped
and Nb-Doped TiO
2
Phase
Lattice
parameters (A

)
Unit cell
volume (A

3
)
Density
(g/cm
3
)
Nb
51
occupancy
factor
TiO
2
a 53.7853 (45) 136.137 3.899 0
c 59.5012 (55)
2 at.%
Nb:TiO
2
a 53.7837 (28) 136.023 3.935 0.0194
c 59.5012 (55)
5 at.%
Nb:TiO
2
a 53.7834 (00) 135.947 4.010 0.0496
c 59.4974 (00)
The Rietveld analyses were done assuming the standard atomic positions of
I41/amd space group.
Fig. 6. Absorption spectra of the TiO
2
and Nb:TiO
2
bers. Inset: a
magnied portion displaying the shift in the absorption peak.
4100 Journal of the American Ceramic SocietyArchana et al. Vol. 93, No. 12
source of decreased V
OC
and FF. The highest m
n
calculated using
the Einsteins equation for the 5%Nb:TiO
2
NWs was
0.16 cm
2
(V s)
1
. This value is only several times less that mea-
sured for single crystalline 5% Nb:TiO
2
lms on LaAlO
3
sub-
strate (10 cm
2
(V s)
1
).
35
In contrast, the m
n
and the D
n
of
mesoporous TiO
2
nanoparticles are several orders of magnitude
less than corresponding single crystalline lms. The Nb-doping
accumulates excess electrons at the bottom of the conduction
band and contributes to the metallic conductivity of anatase
Nb:TiO
2
.
34,35,41
Alternatively, additional free electrons resulting
from doping lls the trap states. As a consequence, photo-injected
electrons move more or less freely through the conduction band.
Therefore, the presented high D
n
and m
n
in Nb:TiO
2
NW in con-
sistent with the calculations
13,14
and experimental ndings.
17
IV. Conclusions
1D nanostructures of Nb:TiO
2
such as continuous NF and high
aspect ratio NW have been synthesized by electrospinning tech-
nique. Precursor for the doping was initially included in the
electrospinning solution; the Nb:TiO
2
nucleated and grown in
the polymeric nanober. XRD measurements and analyses,
X-ray uorescence, and TEM investigations detected the pres-
ence of Nb in the samples. The BET surface area Nb:TiO
2
sam-
ples showed a decrease in trend with increase in the Nb
concentration, which affected the dye-loading adversely. A nom-
inal increase in the J
SC
was observed when Nb:TiO
2
was used as
charge separation and transport medium in DSCs; however,
lower FF and V
OC
due to enhanced recombination decreased
the overall efciency. We observed that Nb doping in 1D TiO
2
nanostructures can further increase the charge mobility (m
n
) and
diffusivity (D
n
) thereby providing opportunities to fabricate
higher performance photoelectrochemical devices. In the present
experiment, we show that 5% Nb:TiO
2
NW increases the D
n
by
an order of magnitude. The measured D
n
of Nb:TiO
2
NW is
B10
3
cm
2
/s for lower photoexcitation densities (B10
16
cm
3
).
However, the wires studied here (150 nm) are much larger than
the diameter of the conventional nanoparticles (B25 nm); and
therefore, have a reduced specic surface area than the latter.
Attaining similar D
n
in wires of diameter in the 2030 nm range
could provide a way to further enhance the properties of devices
fabricated using the Nb:TiO
2
.
Acknowledgments
Dr. Tan Tech Beng, Shimadzu is acknowledged for the XRF measurements.
References
1
Y. N. Xia, P. D. Yang, Y. G. Sun, Y. Y. Wu, B. Mayers, B. Gates, Y. D. Yin,
F. Kim, and Y. Q. Yan, One-Dimensional Nanostructures: Synthesis, Charac-
terization, and Applications, Adv. Mater., 15 [5] 35389 (2003).
2
A. I. Hochbaum, R. Chen, R. D. Delgado, W. Liang, E. C. Garnett, M.
Najarian, A. Majumdar, and P. Yang, Enhanced Thermoelectric Performance of
Rough Silicon Nanowires, Nature, 451 [7175] 1637 (2008).
3
J. R. Heath, Superlattice Nanowire Pattern Transfer (SNAP), Acc. Chem.
Res., 41 [12] 160917 (2008).
4
L. C. Palmer and S. I. Stupp, Molecular Self-Assembly into One-Dimensional
Nanostructures, Acc. Chem. Res., 41 [12] 167484 (2008).
5
D. Li and Y. N. Xia, Electrospinning of Nanobers: Reinventing the
Wheel?, Adv. Mater., 16 [14] 115170 (2004).
6
W. E. Teo and S. Ramakrishna, A Review on Electrospinning Design and
Nanobre Assemblies, Nanotechnology, 17 [14] R89106 (2006).
7
A. Greiner and J. H. Wendorff, Electrospinning: A Fascinating Method for
the Preparation of Ultrathin Fibres, Angew. Chem. Int. Ed., 46 [30] 5670703
(2007).
8
D. H. Reneker and A. L. Yarin, Electrospinning Jets and Polymer Nano-
bers, Polymer, 49, 2387425 (2008).
9
R. Ramasheshan, S. Sundarrajan, R. Jose, and S. Ramakrishna, Nanostruc-
tured Ceramics by Electrospinning, J. Appl. Phys., 102 [12] 111101, 17pp (2007).
10
L. Dan, T. M. Jesse, and Y. Xia, Electrospinning: A Simple and Versatile
Technique for Producing Ceramic Nanobers and Nanotubes, J. Am. Ceram.
Soc., 89 [6] 18619 (2006).
11
W. Sigmund, J. Yuh, H. Park, V. Maneeratana, G. Pyrgiotakis, A. Daga, J.
Taylor, and J. C. Nino, Processing and Structure Relationships in Electro-
spinning of Ceramic Fiber Systems, J. Am. Ceram. Soc., 89 [2] 395407
(2006).
Fig. 7. JV curves of the DSCs fabricated using TiO
2
and Nb:TiO
2
nanowires.
Fig. 8. (A) Typical photocurrent transients observed for pure TiO
2
,
2%Nb:TiO
2
, 5%Nb:TiO
2
NW lms. The transients were not normalized
for the difference in the thickness of the lms, which were 13, 7.8, 8.2 mm
for TiO
2
, 2%Nb:TiO
2
, and 5%Nb:TiO
2
, respectively. (B) Variation of
D
n
as a function of photocurrent density of respective samples deter-
mined from the photocurrent transients.
December 2010 Structural and Electrical Properties of Nb-Doped Anatase TiO
2
Nanowires by Electrospinning 4101
12
Y. Wang, W. Z. Jia, T. Strout, A. Schempf, H. Zhang, B. K. Li, J. H. Cui, and
Y. Lei, Ammonia Gas Sensor Using Polypyrrole-Coated TiO
2
/ZnO Nanobers,
Electroanalysis, 21 [12] 14328 (2009).
13
Y. Wang, W. Z. Jia, T. Strout, Y. Ding, and Y. Lei, Preparation, Charac-
terization and Sensitive Gas Sensing of Conductive Core-Sheath TiO
2
-PEDOT
Nanocables, Sensors, 9 [9] 675263 (2009).
14
R. L. V. Wal, G. M. Berger, M. J. Kulis, G. W. Hunter, J. C. Xu, and L.
Evans, Synthesis Methods, Microscopy Characterization and Device Integration
of Nanoscale Metal Oxide Semiconductors for Gas Sensing, Sensors, 9 [10] 7866
902 (2009).
15
Q. Qi, Y. L. Feng, T. Zhang, X. J. Zheng, and G. Y. Lu, Inuence of Crys-
tallographic Structure on the Humidity Sensing Properties of KCl-Doped TiO
2
Nanobers, Sens. Actuators BChem., 139 [2] 6117 (2009).
16
M. V. Reddy, R. Jose, T. H. Teng, B. V. R. Chowdari, and S. Ramakrishna,
Stable Electrochemical Cycling in Electrospun TiO
2
Nanonets, Electrochim.
Acta, 55 [9] 3109117 (2010).
17
H. W. Lu, W. Zeng, Y. S. Li, and Z. W. Fu, Fabrication and Electrochemical
Properties of Three-Dimensional Net Architectures of Anatase TiO
2
and Spinel
Li
4
Ti
5
O
12
Nanobers, J. Power Sources, 164 [2] 8749 (2007).
18
A. L. Viet, M. V. Reddy, R. Jose, B. V. R. Chowdari, and S. Ramakrishna,
Nanostructured Nb
2
O
5
Polymorphs by Electrospinning for Rechargeable Lith-
ium Batteries, J. Phys. Chem. C, 114 [1] 66471, (2010).
19
X. W. Zhang, S. Y. Xu, and G. R. Han, Fabrication and Photo-
catalytic Activity of TiO
2
Nanober Membrane, Mater. Lett., 63 [21] 17613
(2009).
20
Y. Yang, H. Y. Wang, X. Li, and C. Wang, Electrospun Mesoporous W61-
Doped TiO
2
Thin Films for Efcient Visible-Light Photocatalysis, Mater. Lett.,
63 [2] 3313 (2009).
21
J. S. Im, M. Il Kim, and Y. S. Lee, Preparation of PAN-Based Electrospun
Nanober Webs Containing TiO
2
for Photocatalytic Degradation, Mater. Lett.,
62 [2122] 36525 (2008).
22
S. H. Lee and W. M. Sigmund, Synthesis of Anatase-Silver Nanocomposite
Fibers Via Electrospinning, J. Nanosci. Nanotechnol., 6 [2] 5547 (2006).
23
K. Mukherjee, T. H. Teng, R. Jose, and S. Ramakrishna, Electron Transport
in Electrospun TiO
2
Nanober Dye-Sensitized Solar Cells, Appl. Phys. Lett., 95
[1] 012101, 3pp (2009).
24
R. Jose, A. Kumar, V. Thavasi, and S. Ramakrishna, Conversion Efciency
Versus Sensitizer for Electrospun Nanobers for the Dye-Sensitized Solar Cells,
Nanotechnology, 19, 424004, 7pp (2008).
25
R. Jose, A. Kumar, V. Thavasi, K. Fujihara, S. Uchida, and S. Ramakrishna,
Relationship Between the Molecular Structure of the Dyes and Photocurrent Den-
sity in the Dye-Sensitized Solar Cells, Appl. Phys. Lett., 93, 023125, 3pp (2008).
26
K. Fujihara, A. Kumar, R. Jose, S. Ramakrishna, and S. Uchida, Spray
Deposition of Electrospun TiO
2
Nanorods for Dye-Sensitized Solar Cell, Nano-
technology, 18 [36] 365709, 5pp (2007).
27
K. Onozuka, B. Ding, Y. Tsuge, T. Naka, M. Yamazaki, S. Sugi, S. Ohno, M.
Yoshikawa, and S. Shiratori, Electrospinning Processed Nanobrous TiO
2
Membranes for Photovoltaic Applications, Nanotechnology, 17 [4] 102631
(2006).
28
M. Y. Song, Y. R. Ahn, S. M. Jo, D. Y. Kim, and J. P. Ahn, TiO
2
Single-
Crystalline Nanorod Electrode for Quasi-Solid-State Dye-Sensitized Solar Cells,
Appl. Phys. Lett., 87, 113113, 3pp (2005).
29
M. Y. Song, D. K. Kim, K. J. Ihn, S. M. Jo, and D. Y. Kim, New Appli-
cation of Electrospun TiO
2
Electrode to Solid-State Dye-Sensitized Solar Cells,
Synth. Met., 153 [13] 7780, 5pp (2005).
30
H. Kokubo, B. Ding, T. Naka, H. Tsuchihira, and S. Shiratori, Multi-Core
Cable-Like TiO
2
Nanobrous Membranes for Dye-Sensitized Solar Cells, Nano-
technology, 18, 165604, 7pp (2007).
31
S. Chuangchote, T. Sagawa, and S. Yoshikawa, Efcient Dye-Sensitized
Solar Cells Using Electrospun TiO
2
Nanobers as a Light Harvesting Layer,
Appl. Phys. Lett., 93 [3], 033310, 3pp (2008).
32
B. H. Lee, M. Y. Song, S.-Y. Jang, S. M. Jo, S.-Y. Kwak, and D. K. Kim,
Charge Transport Characteristics of High Efciency Dye-Sensitized Solar Cells
Based on Electrospun TiO
2
Nanorod Photoelectrodes, J. Phys. Chem. C, 113,
2143557 (2009).
33
Y. Furubayashi, T. Hitosugi, Y. Yamamoto, K. Inaba, G. Kinoda, Y. Hirose,
T. Shimada, and T. Hasegawa, A Transparent Metal: Nb-Doped Anatase TiO
2
,
Appl. Phys. Lett., 86 [25] 2521013 (2005).
34
T. Hitosugi, H. Kamisaka, K. Yamashita, H. Nogawa, Y. Furubayashi,
S. Nakao, N. Yamada, A. Chikamatsu, H. Kumigashira, M. Oshima, Y.
Hirose, T. Shimada, and T. Hasegawa, Electronic Band Structure of Trans-
parent Conductor: Nb-Doped Anatase TiO
2
, Appl. Phys. Exp., 1, 111203,
3pp (2008).
35
S. X. Zhang, D. C. Kundaliya, W. Yu, S. Dhar, S. Y. Young, L. G.
Salamanca-Riba, S. B. Ogale, R. D. Vispute, and T. Venkatesan, Niobium
Doped TiO
2
: Intrinsic Transparent Metallic Anatase Versus Highly Resistive
Rutile Phase, J. Appl. Phys., 102 [1] 0137014 (2007).
36
J. H. Noh, S. Lee, J. Y. Kim, J. K. Lee, H. S. Han, C. M. Cho, I. S. Cho,
H. S. Jung, and K. S. Hong, Functional Multilayered Transparent Conducting
Oxide Thin Films for Photovoltaic Devices, J. Phys. Chem. C, 113 [3] 10837
(2009).
37
S. Lee, J. H. Noh, H. S. Han, D. K. Yim, D. H. Kim, J. K. Lee, J. Y. Kim, H.
S. Jung, and K. S. Hong, Nb-Doped TiO
2
: A New Compact Layer Material
for TiO
2
Dye-Sensitized Solar Cells, J. Phys. Chem. C, 113 [16] 687882
(2009).
38
Y. Hirose, N. Yamada, S. Nakao, T. Hitosugi, T. Shimada, and T. Hasegawa,
Large Electron Mass Anisotropy in a d-Electron-Based Transparent Conducting
Oxide: Nb-Doped Anatase TiO
2
Epitaxial Films, Phys. Rev. B, 79 [16] 165108,
5pp (2009).
39
M. A. Gillispie, M. van Hest, M. S. Dabney, J. D. Perkins, and D. S. Ginley,
Sputtered Nb- and Ta-Doped TiO
2
Transparent Conducting Oxide Films on
Glass, J. Mater. Res., 22 [10] 28327 (2007).
40
P. Hasin, M. A. Alpuche-Aviles, Y. G. Li, and Y. Y. Wu, Mesoporous Nb-
Doped TiO
2
as Pt Support for Counter Electrode in Dye-Sensitized Solar Cells,
J. Phys. Chem. C, 113 [17] 745660 (2009).
41
X. D. Liu, E. Y. Jiang, Z. Q. Li, and Q. G. Song, Electronic Structure and
Optical Properties of Nb-Doped Anatase TiO
2
, Appl. Phys. Lett., 92 [25] 252104,
3pp (2008).
42
L. R. Sheppard, T. Bak, and J. Nowotny, Electrical Properties of Niobium-
Doped Titanium Dioxide. 3. Thermoelectric Power, J. Phys. Chem. C, 112 [2]
6117 (2008).
43
R. Jose, V. Thavasi, and S. Ramakrishna, Metal Oxides for Dye-Sensitized
Solar Cells, J. Am. Ceram. Soc., 92 [2] 289301 (2009).
44
X. J. Lu, X. L. Mou, J. J. Wu, D. W. Zhang, L. L. Zhang, F. Q. Huang, F. F.
Xu, and S. M. Huang, Improved-Performance Dye-Sensitized Solar Cells Using
Nb-Doped TiO
2
Electrodes: Efcient Electron Injection and Transfer, Adv. Fun-
ct. Mater., 20 [3] 50915.
45
A. Bauer, K. Lee, C. Song, Y. Xie, J. Zhang, and R. Hui, Pt nanoparticles
Deposited on TiO
2
Based Nanobers: Electrochemical Stability and Oxygen
Reduction Activity, J. Power Sources, 195 [10] 310510 (2010).
46
A. Kumar, R. Jose, K. Fujihara, J. Wang, and S. Ramakrishna, Structural
and Optical Properties of Electrospun TiO
2
Nanobers, Chem. Mater., 19,
653642 (2007).
47
W. Kraus and G. Nolze, POWDER CELL-A Program for the Representa-
tion and Manipulation of Crystal Structure and Calculation of the Resulting
X-Ray Powder Pattern, J. Appl. Crystallogr., 29 [3] 3013 (1996).
48
P. S. Archana, R. Jose, C. Vijila, and S. Ramakrishna, Improved Electron
Diffusion Coefcient in Electrospun TiO
2
Nanowires, J. Phys. Chem. C, 113 [52]
2153842 (2009).
49
Q. Wang, S. Ito, M. Gra tzel, F. Fabregat-Santiago, I. Mora-Sero, J. Bisquert,
T. Bessho, and H. Imai, Characteristics of High Efciency Dye-Sensitized Solar
Cells, J. Phys. Chem. B, 110 [50] 2521021 (2006). &
4102 Journal of the American Ceramic SocietyArchana et al. Vol. 93, No. 12

Das könnte Ihnen auch gefallen