Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Nanoscale

View Article Online


PAPER View Journal | View Issue

Doping of wide-bandgap titanium-dioxide


Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

nanotubes: optical, electronic and magnetic


Cite this: Nanoscale, 2014, 6, 10839
properties†
Yahya Alivov,*a Vivek Singh,a Yuchen Ding,a Logan Jerome Cerkovnika
and Prashant Nagpal*abc

Doping semiconductors is an important step for their technological application. While doping bulk
semiconductors can be easily achieved, incorporating dopants in semiconductor nanostructures has
proven difficult. Here, we report a facile synthesis method for doping titanium-dioxide (TiO2) nanotubes
that was enabled by a new electrochemical cell design. A variety of optical, electronic and magnetic
dopants were incorporated into the hollow nanotubes, and from detailed studies it is shown that the
doping level can be easily tuned from low to heavily-doped semiconductors. Using desired dopants –
Received 5th May 2014
Accepted 7th July 2014
electronic (p- or n-doped), optical (ultraviolet bandgap to infrared absorption in co-doped nanotubes),
and magnetic (from paramagnetic to ferromagnetic) properties can be tailored, and these
DOI: 10.1039/c4nr02417f
technologically important nanotubes can be useful for a variety of applications in photovoltaics, display
www.rsc.org/nanoscale technologies, photocatalysis, and spintronic applications.

(carbon nanotubes, activated carbons, and graphene) which


Introduction cause a red-shi in the absorption spectrum and increase elec-
Titanium-dioxide (TiO2) is a wide-bandgap semiconductor that tron mobility.20–23 Since doping can simultaneously tune the
is a key component of devices like solar cells, photocatalytic bandgap and increase charge conductivity, it can provide an
reactors, photoelectrochemical cells, etc.1–4 While doping (elec- important alternative. Therefore, desired doping of these wide-
tronic, optical, and magnetic) allows careful tuning of desired bandgap nanostructured lms can provide important materials
properties, oxygen vacancies formed during the synthesis are for a variety of applications in renewable energy, articial
used as n-type dopants in these nominally “undoped” oxides. displays, and other optoelectronic and magnetic processes.
Since desired doping of these semiconductor nanostructures One of the synthesis methods for the growth of TiO2 nano-
has proved challenging, most applications are focused on bulk tubes is electrochemical oxidation, also termed as anodiza-
semiconductor lms.5–7 This can severely limit their applications tion.24–28 Since this synthesis provides a facile route towards
and device architectures, like fabricating p-type oxide lms or large-scale fabrication of TiO2 nanotubes, simply incorporating
changing the doping level of n-type dopants for optimized desired dopants during the growth can provide an important
conguration.1,3 Recent research in utilizing heavily-doped method for the synthesis of doped wide-bandgap semi-
semiconductors for low-loss plasmonics8,9 has also been gener- conductor nanostructures. However, simple addition of the
ating interest in doping of these TiO2 nanostructures, already desired concentration of precursors, especially cationic, does
used in thin lm solar-cells, for enhanced light harvesting. not lead to doping of nanotubes (see Experimental section).
Moreover, addition of desired dopants, like shallow or deep Therefore, we designed a new electrochemical cell that enables
donors, can also lead to the formation of heavily-doped trans- doping and co-doping of TiO2-nanotubes with desired cations
parent oxide nanostructures, or co-doped photocatalysts10,11 and anions, and characterized their resulting optical, electronic
which can absorb infrared light. Other efforts to improve the and magnetic properties, to demonstrate the feasibility of this
performance of TiO2 photocatalytic activity12–19 involves hybrid new method. Using detailed studies, we demonstrate that this
composites consisting of TiO2 and carbonaceous materials method can provide an important route for the synthesis of
doped wide-bandgap TiO2 semiconductor nanotubes.

a
Department of Chemical and Biological Engineering, University of Colorado, Boulder,
USA. E-mail: y.alivov@colorado.edu; pnagpal@colorado.edu Experimental
b
Renewable and Sustainable Energy Institute, University of Colorado, Boulder, USA
The method for TiO2 nanotube growth by anodization
c
Materials Science and Engineering, University of Colorado, Boulder, USA
† Electronic supplementary information (ESI) available: See DOI: In a regular setting of an electrochemical cell, a negative-biased
10.1039/c4nr02417f platinum (Pt) electrode and positive-biased titanium (Ti) sheet

This journal is © The Royal Society of Chemistry 2014 Nanoscale, 2014, 6, 10839–10849 | 10839
View Article Online

Nanoscale Paper

are immersed in the electrolyte 2–3 cm apart.24–28 In this currents in this conguration are comparable to the currents in
conguration negative ions in the electrolyte move toward the regular electrochemical cells with the DC power supply, ranging
positive Ti sheet and positive ions move toward the negative Pt from 0.1–4.0 mA within the voltage range of 7–200 V. For the
sheet. Addition of desired dopant metal cations into the elec- separating membrane, we used a hydrophilic polymer lter
trolyte results in the repulsion of the positively charged ions (pore size 0.1–0.4 mm). This membrane was attached in the
from the positively biased Ti sheet (away from the growing TiO2 middle of the container using epoxy. While the membrane
nanotubes). This prevents TiO2 nanotubes from doping with provides a diffusion barrier, some cations move toward the Pt
metal cations. electrode and over time the conductivity of the dopant-free
Replacing the DC power supply with AC can provide a solu- electrolyte will increase. The electrolyte should be changed if the
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

tion. During the “negative” cycle (switched polarities) of the resistivity decreases by 50%.
electrochemical growth, the desired cations are attracted by the Using an AC voltage source, rectangular shaped power pulses
Ti sheet, thus incorporating cationic dopants uniformly in the were used in this present study. The shape of the AC voltage and
growing TiO2 nanotubes (Fig. 1a). However, simple addition of its parameters are shown in schematic Fig. 1b. The period of AC
cationic precursors in the growth solution leads to other prob- voltage is characterized by the positive part V1, the negative part
lems. First, the added cations (especially in high doping 1
V2, and the frequency f, which is equal to f ¼ , where t1
concentrations) dramatically increase conductivity, making it t1 þ t2
impossible to apply high enough voltage to enable TiO2 nano- and t2 are durations for positive and negative parts of the cycle,
tube growth (voltages higher than 5 V need to be applied to respectively. These parameters – t1 and t2 control the growth
initiate nanotube growth24–28). Second, during the positive bias and doping times, respectively. From a series of detailed
on the Ti sheet (negative bias on the Pt counter electrode), experiments, it was found that the dopant level, or dopant
cations deposit on the Pt electrode, rapidly covering it with a concentration n, mostly depended on the negative bias V2,
reduced metal layer, quickly degrading its performance. frequency f, t1/t2 ratio (see Fig. 1b), and dopant precursor
To overcome these problems we modied the electro- amount b in the electrolyte. Our studies showed that the
chemical cell. The electrochemical cell was divided into two dependence of n on f and t1/t2 at xed V2 and b is convoluted,
sections, separated by a porous membrane (Fig. 1a). One part of depending on a number of factors. Therefore, we studied the
the cell contains the Pt electrode and is lled with an electrolyte dependence of doping (n) on each of these parameters sepa-
free of the dopant precursor. The second part of the cell contains rately, keeping the other parameters constant. The dependence
the Ti sheet and is lled with an electrolyte that contains the of n on V2 (when other parameters were xed) was more
dopant precursor. The porous membrane provides a barrier for predictable; however, a strong sublinear relationship between
two different parts of the electrochemical cell to prevent from these two parameters was observed. The dependence between n
mixing, while still allowing a ow of the electrolyte. The resulting and the amount of precursor b in the electrolyte was nearly
linear (see Discussion below). Therefore, variation of the dopant
concentration was used for systematically controlling the
doping level in fabricated TiO2 nanotubes.
The electrolyte used in electrochemical growth solution
consisted of a solvent (glycerol or ethylene glycol) with 1%
ammonium uoride (NH3F), and 2% of water. As the starting
material commercial Ti sheets of 99.99% purity were used. The
growth rate of nanotubes in ethylene glycol was 4 mm per hour.
All as-grown samples were amorphous and annealed at 500  C
for 1 hour in air, to convert to the anatase phase. Three main
types of cationic dopants were used in this study – niobium (Nb),
iron (Fe), and copper (Cu). Among the prominent anion dopants,
we used nitrogen (N), while several combinations of co-doped
anion and cation co-dopants were prepared. As precursors nio-
bium(V) chloride, iron(III) chloride, and copper(II) sulfate were
used. Nitrogen doping was performed by using hexamethylene-
tetramine in the electrolyte. The nitrogen doping level was
controlled by changing the amount of the precursor (hexa-
methylenetetramine) in the electrolyte. While this method can
be applied to any cation, we focused on Nb, Fe, and Cu due to
their importance for a variety of applications. For example, Nb
was shown to be a shallow donor for TiO2, and therefore, a good
Fig. 1 (a) A sketch showing the design of a new electrochemical cell candidate for the growth transparent conducting oxide thin
developed in this work. It is divided into two parts separated with a
lms.29–32 Fe is a good dopant for fabricating magnetic semi-
porous membrane to prevent from high currents and cation deposi-
tion onto the Pt electrode. (b) Schematic illustrating rectangular shape conductors.33–35 Cu and Nb co-doped TiO2 thin lms with N have
AC signals with parameters employed for doping nanotubes. been shown to increase the photocatalytic activity of TiO2.36

10840 | Nanoscale, 2014, 6, 10839–10849 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper Nanoscale

The morphology of grown TiO2 nanotubes was examined magnetometer (SQUID). Magnetization measurements were
using a eld-emission scanning electron microscope (FE-SEM) performed using a Quantum Design SQUID-VSM (Vibrating
JEOL 7401F. The crystal structures of doped and undoped TiO2 Sample Magnetometer), and magnetization (M) was measured
nanotube samples were analyzed by X-ray diffraction (XRD) as a function of the applied eld (H) and temperature (T) in the
measurements, using a Scintag XDS 2000 X-ray diffractometer. magnetic eld range from 7 to 7 Tesla and a temperature
The measurements were performed using Cu Ka radiation at 45 range of 1.8 to 300 K. Temperature dependent magnetization
kV and 40 mA. Compositional analysis was performed by energy (M–T) measurement was performed at the applied magnetic
dispersive X-ray spectroscopy (EDS) embedded in an FE-SEM eld of 1000 Oe. TiO2 nanotube materials for SQUID studies
JEOL 7401F instrument. The analysis was performed in were collected from the surface of a Ti sheet aer the growth.
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

different modes (2D mapping, point scan, etc.) to test for The eld-cooled (FC) and zero-eld-cooled (ZFC) magnetization
uniformity and possible secondary phases. The spatial resolu- measurements are performed from 5 to 300 K. TiO2 nanotube
tion during EDS was limited by the focused spot size of the samples were scraped from the Ti sheet aer the TiO2 nanotube
electron beam, which was 10 nm. growth. Therefore, the scraped sample was a mixture of
Electrical characterization. I–V and I–V–T characterizations randomly oriented nanotubes. As a result, induced magnetiza-
were performed in two different ways: (1) from single nanotubes tion of the TiO2 nanotube sample was averaged from all nano-
using CS-AFM; and (2) ensemble measurements using a Keith- tube orientations, and therefore, it was isotropic with respect to
ley source meter (Keithley 2612A, Tektronix Inc.). Temperature the magnetic eld direction. Extreme precautions were taken
dependent current–voltage characterization (I–V–T) was per- during these measurements to avoid any contamination of the
formed in the temperature range of 20–300 K with 10 K samples with possible magnetic dopants. Magnetization of the
temperature steps, using a closed loop Helium Cryostat (ARS- sample holder was also measured and subtracted from that of
202AE with an ARS-2HW Helium compressor, Advanced TiO2 nanotube samples to calculate the pure magnetization of
Research Systems Inc.). The voltage for these measurements doped and undoped TiO2 nanotubes.
was varied in the range from 20 V to +20 V. Sheet resistance of Photocatalysis – Gas-phase reduction of CO2 and H2O.
nanotubes R was calculated from I–V measurements using Twenty to thirty mg of the catalyst was deposited in a rounded
V glass vial cut in half with a 0.64 cm2 cross-sectional area. The
Ohm's law I ¼ . Then, resistivity r was calculated using the
R vial was then enclosed in a 48 mL reactor and purged for 45
L minutes with CO2 (75 cm3 min1) humidied in a bubbler lled
equation R ¼ r , where L is the nanotube length, and A – the
A with D.I. water. Aer purging, the reactor was closed and irra-
total nanotube cross-section area. The cross-section area A was
diated with 1 Sun (100 mW cm2) through a transparent
estimated using the density of nanotubes, thickness of the
window using a solar simulator (ABET Technologies). One-
nanotube wall, and inner and outer diameters of the nanotube,
milliliter samples were extracted from the reactor and injected
which were measured using SEM images.
into a gas chromatograph (GC) equipped with a thermal
Photoresponse measurements. Spectral photo I–V measure-
conductivity detector (TCD) and a ame ionization detector
ments were performed in the wavelength range of 300 –1000
(FID) for measuring concentrations. Separation of the hydro-
nm. Monochromatic light was obtained using a mono-
carbons was done with a Hayesep D column or with Silica-gel
chromator from a light source. Photocurrent was calibrated to
column. Reported data of the photocatalytic rate and quantum
the incident power of monochromatic light measured using an
yield were calculated based on the electron ux (mmol cm2 h1)
Si-based detector (NIST calibration). The intensity of mono-
used to form H2, CH4, C2H6, hydrocarbons and others.
chromatic light was measured using an optical power meter
(Newport, 1830-C).
STM and STS measurements. Scanning tunneling micro- Results and discussion
scope images were obtained using a customized Molecular SEM, XRD, STM, and absorption
Imaging PicoScan 2500 setup (with a PicoSPM II controller). An
Fig. 2 presents SEM images of doped TiO2 nanotubes fabricated
STM nosecone (N9533A series, Agilent Technologies) was used
using our new electrochemical cell. The electron micrographs
for scanning and spectroscopy, using chemically etched Pt–Ir
clearly show that the tubular structure of these TiO2 nanotubes
tips (80 : 20) purchased from Agilent Technologies, USA. The
is well preserved aer doping. Fig. S1 in the ESI† shows SEM
measurements were done at room temperature under atmo-
images of different diameter nanotubes grown at different
spheric conditions. Tunneling junction parameters were set at
anodization voltages. XRD patterns of all doped samples shown
tunneling currents ranging between 100 and 500 pA and a
in Fig. 3a present X-ray diffraction peaks corresponding to the
sample bias voltage between 5 and +5 V. Spectroscopy
anatase phase, with predominant peaks (101) and (200) at 2q ¼
measurements were obtained at a scan rate of 1 V s1. For STM
25.3 and 47.95 , respectively.37 The XRD pattern of samples
measurements, the pre-amp sensitivity was set to 1 nA V1. The
does not change even with a heavy doping level up to 12%,
tunneling current as a function of the applied bias voltage (STS)
demonstrating the uniform incorporation of dopants and the
was recorded at multiple positions on the sample at room
absence of secondary phases. At such high doping levels (>9%),
temperature and atmospheric pressure.
they can also be regarded as solid solutions. As an illustration in
SQUID measurements. The magnetic properties were
Fig. S2a,† we show XRD patterns of Nb doped TiO2 nanotubes
studied with a Superconducting Quantum Interference Device
with a doping level up to 12%, which shows only reection

This journal is © The Royal Society of Chemistry 2014 Nanoscale, 2014, 6, 10839–10849 | 10841
View Article Online

Nanoscale Paper
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

Fig. 2 Representative scanning electron micrographs (SEM) of TiO2 nanotube arrays grown by electrochemical anodization: (a) shows the top
view of densely packed nanotubes, (b) shows the bottom view after removing nanotubes from the Ti metal, and the side view of the vertical,
hollow, doped TiO2 nanotubes, and (c) and (d) show side profiles at different magnifications.

peaks corresponding to anatase TiO2 without any detectable demonstrate that dopants (Cu here) were uniformly incorpo-
secondary phase. No shi of diffraction peaks was observed rated in the TiO2 crystal lattice. Similar scans for all dopants
with the increase of Nb concentration up to 12%, as shown in (cations and anions) were performed to ensure uniform doping
Fig. S2a.† This result could be explained by close ionic radii of in TiO2 nanotubes.
Ti4+ and Nb5+, which are 0.605 Å and 0.64 Å, respectively.38 For Absorbance spectra, of undoped and different doped
such similar lattice parameters, no signicant distortion of the nanotubes are shown in Fig. 4a. The optical spectrum does
crystal lattice was expected. Energy dispersive X-ray spectros- not reveal either the nature of the dopant, or the direction of
copy (EDS) conrmed the presence of doped elements in TiO2 the shi of the Fermi-level (which was analyzed by Scanning
nanotubes as shown in Fig. 3b. EDS analysis of undoped TiO2 Tunnelling Spectroscopy (STS), as shown in Fig. 4b) towards
nanotubes showed no detectable emission peaks from impuri- the conduction or valence band. But optical spectroscopy was
ties (only Ti and O emission peaks were seen, Fig.3b, the red used to monitor the semiconductor bandgap by changing the
curve). This indicates that the impurity concentration was dopant and the concentration of the dopant in the TiO2
below the detection limit (<0.01%). The amount of incorporated nanotube. Moreover, the electronic level of the dopant was
dopants in TiO2 nanotubes follows a linear relationship with also estimated, since the Fermi-level (and the effective optical
the concentration of the precursor in the electrolyte. To bandgap) is also pinned at high dopant concentrations.
demonstrate this, we show EDS spectra of Nb doped samples at However, this analysis is complicated by the indirect
different NbCl5 wts% in the electrolyte (Fig. S2b†), from which a bandgap of the TiO2 semiconductor, large Rayleigh scat-
nearly linear relationship between the detected EDS Nb signal tering (f1/l4) observed in the nanotubes (shown as a black
in TiO2 nanotubes and the wt% NbCl5 precursor is clearly seen dashed curve in Fig. 4a), and the presence of impurity donor
(Fig. 3c). The R2 value of the linear tting was calculated to be states (below the bandgap) in undoped TiO2. The band gaps
0.995. Similar studies for other dopants (Nb, Fe, Cu, and N) also estimated from their respective optical spectra (aer
showed a linear relationship between the incorporated dopant considering scattering) were 2.7 eV (due to impurity donors),
and the dopant concentration in the electrolyte. The conduc- 2.6 eV, 2.5 eV, 2.4 eV, 2.3 eV, 2.2 eV, and 2.0 eV for undoped,
tivity of Nb-doped TiO2 nanotubes increases (resistivity Nb, Cu, Fe, N, Nb–N, and Cu–N doped and co-doped samples,
decreases) with the wt% NbCl5 precursor (Fig. 3c), conrming respectively (Fig. 4a). These bandgaps were in good agree-
the incorporation of Nb into the TiO2 crystal. Nb is a well-known ment with the bandgaps derived from STS (Fig. 4b), which
shallow donor in TiO2 used for creating transparent conducting uses the xed Fermi-level of the Pt–Ir tip (at 0 V) as a probe of
thin lms.39,40 EDS point scans were also taken from different the shi of the Fermi-level with doping. To further evaluate
points on individual doped TiO2 nanotubes. This scan was the role played by these dopants on optoelectronic proper-
performed by pointing a focused electron beam onto the surface ties, and to ensure that the dopants are properly incorporated
of a single nanotube (or mapping the elemental distribution in the TiO2 crystal (and not as recombination sites or traps on
over a large area). Fig. S3† presents results for Cu-doped the surface), we evaluated the photoresponse of these doped
samples. Highly pronounced Cu peaks seen at different points nanotubes (Fig. 4c). We illuminated these lms with a small
(we are showing two points for clarity) indicate Cu is uniformly ux of photons (to identify possible recombination from
incorporated in TiO2 nanotube crystals, rather than accumu- dopants as traps), and observed the change in current (from
lating as clusters or secondary phase oxides. These data further the dark) due to photogenerated charge carriers, normalized

10842 | Nanoscale, 2014, 6, 10839–10849 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper Nanoscale
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

Fig. 3 (a). XRD patterns for undoped, Nb, Cu, Fe, and N doped, and
Nb/N and Cu/N co-doped and undoped samples. Labels on the right
guide to the corresponding samples. Only reflection peaks corre-
sponding to the anatase TiO2 phase were detected in all samples
indicating the lack of secondary phases. (b) Energy dispersive spec-
troscopy (EDS) for undoped, Fe-doped, Cu-doped, Nb-doped, and Fig. 4 (a) STS spectrum of (a) undoped, (b) Cu-doped, (c) N-doped, (d)
N-doped, Nb and N co-doped, and Cu and N co-doped TiO2 nano- Nb-doped, (e) Nb–N co-doped, and (f) Cu–N co-doped samples. (b)
tubes. Well pronounced characteristic X-ray emission peaks of Extinction spectra of undoped, monodoped and co-doped TiO2
dopants can be seen along with the TiO2 matrix (Ti and O peaks). These samples. A clear change in the bandgap can be seen using the STS and
data prove the presence of desired dopants in TiO2 nanotube films; (c) the optical spectrum of these doped wide-bandgap semiconductor
Nb concentration (blue) and resistivity (red) of Nb doped TiO2 nano- nanotubes. The dashed-black line represents the large Rayleigh
tubes, as a function of wt% NbCl5 in the electrolyte. scattering observed in these nanotubes, which makes exact determi-
nation of bandgaps difficult in these indirect bandgap semiconductors,
using optical spectroscopy; (c) spectrally resolved photoresponse
(photocurrent normalized by incident light intensity) for undoped
Ilight  Idark
by the intensity of light. The photoresponse Iph ¼ (blue), Cu-doped (pink), N-doped (green), and Cu–N co-doped (red)
Plight TiO2 nanotubes.
for highly conductive samples (Fig. 4c) was within the error
bar for the measurement, due to the high concentration of
background electrons (photogenerated charge carriers are I–V, I–V–T, and photo-I–V
much less than equilibrium carrier concentrations).
We also conducted current–voltage spectroscopy (I–V), on all
However, the acceptor dopants (Cu, N, and Cu–N) and
undoped and doped nanotubes, using ensemble and single
undoped samples showed good transport for photogenerated
nanotube current sensing AFM (CS-AFM) measurements
charges in these devices, following exciton dissociation and
(Fig. 5a and 7). The dopant concentrations for all cations and
charge transport in these doped nanotubes. Spectrally
anions were kept constant for all samples in Fig. 5a, such that
resolved normalized photocurrent (Fig. 4c) mimics the
the differences in the charge carrier concentrations (electrons
absorption spectra of the respective nanotubes, shown in
or holes) arise from the number of ionized dopants
Fig. 4a, with similar bandgaps. This was expected since the  
Ei
photocurrent is limited by the thin lm absorption of (n ¼ n0 exp  ), which depends on the dopant electronic
kT
incident light.
level Ei and the available energy at room temperature (kBT ¼ 25

This journal is © The Royal Society of Chemistry 2014 Nanoscale, 2014, 6, 10839–10849 | 10843
View Article Online

Nanoscale Paper
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

Fig. 5 Electronic n- and p-type doping of TiO2 nanotubes: (a) room-


temperature current–voltage (I–V) spectrum of TiO2 nanotubes doped
with 1 wt% of (a) Copper (Cu), (b) Copper–Nitrogen (Cu–N) co-dopant,
(c) Nitrogen (N), (d) Niobium–Nitrogen (Nb–N) co-dopant, (e) undoped,
(f) Iron (Fe), and (g) Niobium (Nb); (b) Temperature dependent current–
voltage characteristics presented in 3D mode; (b), conductance (I/V) of
Nb-doped (curve 1), Fe-doped (curve 2), undoped (curve 3), and Cu-
doped (curve 4) samples as a function of 1/kT. The slope at the higher Fig. 6 (a) Conductivity (green), carrier concentration (red), and
temperature region (>160 K) revealed activation energies 0.018 eV, mobility (blue) of N-doped samples as a function of N concentration.
0.037 eV, 0.093 eV, and 0.502 eV for Nb-doped, undoped, Fe-doped, (b) Scanning tunneling spectroscopy (STS) data for nitrogen-doped
and Cu-doped samples, respectively; (c) CS-AFM I–V curve and the TiO2 nanotubes, showing a clear shift of the Fermi-level (at 0 eV)
corresponding ln(I)–V plot for undoped TiO2 nanotubes to illustrate the towards the valence band (p-type doping).
calculation method for carrier concentration; (d) STS spectra for Nb-
doped and N-doped TiO2 nanotube; dashed line guides to the Fermi-
level (constant for our Pt–Ir tip) which is located near the bottom of the
conduction band for the Nb-doped sample (n-type), and near the top of conduction in these doped nanostructures, and further
the valence band for the nitrogen doped sample (p-type). evidence for electronic doping of TiO2 wide-bandgap nanotubes
using cationic and anionic dopants.

meV, where kB is the Boltzmann constant and T is the temper-


ature). To further understand ionization of dopants to generate
doped charge carriers, which determines the electronic property
of doped semiconductors, temperature dependent I–V charac-
teristics (I–V–T) of samples were measured. We plotted the
conductivity s ¼ ln(I/V) as a function of temperature T (Fig. 5b),
to measure the activation barrier for charge transport (or the
energy required for ionizing charge carriers in a doped semi-
conductor). Activation energies (Ea) for undoped and Nb, Fe,
and Cu doped samples were found to be 0.027 eV, 0.018 eV,
0.093 eV, and 0.502 eV, respectively (Fig. 5b and Table 1). The
activation energy for Nb-doped TiO2 nanotubes was close to the
reported ionization energy values for Nb donors in anatase TiO2
thin lms and bulk crystals, which was in the range of 10–40
meV.39,40 These values are close to the values derived using
Richardson plots (ln(I/T2) vs. 1/kT) (Fig. S5†). The mobility of
electron doped TiO2 nanotubes – oxygen vacancies (undoped),
Fe-doped, and Nb-doped; was very close with values 4.3, 2.7, and
3.8 cm2 V1 s1, respectively (Table 1). In this case the differ-
ences in conductivity of donor doped samples can be accounted
due to the difference in ionized donors. In the case of Cu, N, and
Cu–N dopants a drastic drop in mobility was observed (Table 1) Fig. 7 (a) I–V characteristics of low conductivity Cu-doped (1 wt%)
TiO2 nanotubes are well described by the typical space charge limited
that can be explained by the contribution of holes to transport
current (SCLC) model (I–V2 at higher bias). (b) Modified Fowler–
properties, since the hole mobility in TiO2 is lower than the Nordheim plot (explained in the text) for photogenerated electrons in
electron mobility.41 These ndings provide insights into charge Cu-doped TiO2 nanotubes.

10844 | Nanoscale, 2014, 6, 10839–10849 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper Nanoscale

Table 1 Summary of electrical parameters of doped TiO2 nanotubes

Undoped Fe Nb Cu N Cu–N

Resistivity, r, U cm 6.92 2.97 0.201 906 2300 4.4  106


Carrier concentration, n, cm3 8.9  1017 2.5  1018 5  1019 1.6  1016 2.8  1016 1.3  1015
Mobility, m, cm2 V1 s1 4.3 2.7 3.8 0.3 0.097 0.002
Activation energy (eV) 0.027 0.093 0.014 0.502 0.14
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

We analyzed the transport properties of doped TiO2 nano- the available reports indicate that substitutional Fe produces
tubes, such as carrier concentration (n) and mobility (m) for n-type conductivity.43–45 Roldan et al. showed, using periodic
these nanotubes, using single nanotube CS-AFM measure- density functional calculations, that the Fe-dopant stimulates
ments. We used a conductive gold coated AFM tip (known cross- the generation of oxygen vacancies that act like shallow
section of the contact), with a single nanotube (Schottky-barrier) donors.45 Nitrogen doped nanotubes revealed increased resis-
and measured the current–voltage (I–V) curves from a number tivity (Table 1). Similar to cation dopants, doping of N anions
of nanotubes. We used the I–V curves (Fig. 5c) in the interme- was also in good linear relationship with the added N precursor
diate bias regime (where the reverse-biased Schottky barrier in the electrolyte, with an R2 tting value of 0.988 (Fig. S7a†).
dominates the total current, Fig. S6, details in the ESI†):42 Nitrogen is known as a good acceptor in TiO2, with an energy
  position of 0.14 eV above the top of the valence band.46 The
  e 1
ln I ¼ ln S þ  V þ lnJs (1) acceptor nature of the nitrogen dopant in our samples is
kT E0
conrmed from STS studies, where the spectrum shows that the
where J is the current density, S is the contact area, E0 is a Fermi level shis toward the valence band, indicating a p-type
parameter depending on carrier concentration n, e is the effect of these dopants (Fig. 5d). By contrast, the Fermi-level of
elementary charge, k – the Boltzmann constant, and Js is a Nb doped samples shis towards the conduction band
function of the applied bias (see ESI†). This logarithmic plot of (compared with the Pt–Ir tip level at zero). The p-type behavior
the current I as a function of the bias V gives approximately a of nitrogen-doped TiO2 nanotubes was also observed from
straight line, the slope of which is equal to e/(kT)  1/E0. electrical characterization, which showed a clear injection of
The electron mobility can be calculated using the relation holes, as opposed to the electron injection in n-type Nb-doped
m ¼ 1/(ner), with r being the resistivity of the nanotubes. Table 1 TiO2 nanotubes (Fig. S7b†). Since nominally undoped nano-
summarizes the extracted values of resistivity, carrier concen- tubes behave as n-doped semiconductors, the addition of an
tration, and mobility for doped, co-doped, and undoped increasing amount of p-type dopants (like nitrogen) in nano-
samples. The resistivity r of the samples was calculated using tubes rst results in a decrease in n-type conductivity ((from 0 to
sheet resistance R determined using the Ohmic law (see 1 wt%, Fig. 6a), followed by conversion to the p-type and an
Experimental section). There was a strong effect of dopants on increase in p-type conduction, as shown in Fig. 6a. The
electrical properties, and the resistivity ranged from 6  104 conductivity rst decreases and then increases, with an increase
(Fig. 9a) to 3  106 U cm (Table 1). The resistivity of the undoped of N concentration (m) in TiO2, reaching a minimum at m ¼ 2%,
sample was 9.92 U cm. As-grown undoped TiO2 nanotubes were as shown. These phenomena can be easily explained by the shi
of the n-type conductivity due to oxygen vacancies acting as in the Fermi-level of the semiconductor with increasing N
shallow donors.2,15 The n-type conductivity of as-grown TiO2 concentration, monitored by the tunneling spectra using Pt–Ir
nanotubes is seen from the STM spectrum (Fig. 5d), where the STM tips (the tip Fermi-level is xed). A clear shi of the Fermi-
density of states (DOS) is seen near the conduction band and level towards the valence band was observed with an increase in
the Fermi energy is also close to the conduction band. Nb doped N concentration, indicating an increasing p-type character for
samples revealed the lowest resistivity 0.2 U cm. The increased doped TiO2 nanotubes shown in Fig. 6a.
conductivity of Nb-doped TiO2 nanotubes can be explained by While most dopants described so far have behaved as
the replacement of Ti atoms by Nb atoms in the crystal lattice. substitutional dopants, copper incorporated in the TiO2 nano-
Nb is known as a good shallow donor in the TiO2 crystal lattice, tube lattice shows an interesting behavior. A copper dopant also
with the activation energy in the range of 10–50 meV.29–32 The behaves like an acceptor, leading to an enhancement of the
presence of shallow donor states near the conduction band (CB) initial resistivity of Cu doped TiO2 nanotubes (Fig. 4a and b).
of Nb doped TiO2 nanotubes samples was also conrmed by the The resistivity of copper (Cu) doped samples increases (Fig. 5a
scanning tunneling spectrum, which showed a high density of and Table 1), and can be explained by the creation of acceptor
states near the conduction band (Fig. 4b). The iron (Fe)-doped centers that compensate for background impurity donors in a
TiO2 sample also increased the conductivity, compared to the nominally undoped sample. This behavior is surprising since
undoped sample, which indicates donor-like substitution of the substitution of the titanium cation (Ti4+) with copper (Cu+ or
this dopant iron. This effect of Fe is less dramatic compared to Cu2+) is expected to show n-type doping.36 However, copper
Nb, which can be explained by the higher activation energy for doped nanotubes show clear p-type conductivity (conrmed
Fe (0.097 eV) compared to Nb (0.018 eV). There have been few using STS spectra shown in Fig. 4b). This result can be
reports on electrical properties of Fe-doped TiO2 nanotubes and explained by interstitial doping of copper in TiO2 nanotubes,

This journal is © The Royal Society of Chemistry 2014 Nanoscale, 2014, 6, 10839–10849 | 10845
View Article Online

Nanoscale Paper

since it has been shown that Cu atoms can act as deep acceptors
when they are in interstitial positions in the crystal lattice.36
Therefore, Cu likely occupies interstitial positions in the TiO2
crystal lattice (interstitial dopant) in doped nanotubes fabri-
cated using our electrochemical cell. I–V characteristics of Cu
doped TiO2 nanotube samples at higher voltages showed a clear
nonlinear behavior of the form I–V2 (Fig. 7a), which can be
explained by the space-charge limited current model (SCLC).47
In this SCLC region the carrier concentration can be calculated
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

3Vc
using equation n ¼ 2 ,47 where 3 is the dielectric constant, R- is
eR
the radius of the nanotubes, and Vc is the crossover voltage
presenting a transition from Ohmic to SCLC regimes. Extracting
charge carrier concentration using the SCLC model for Cu
doped TiO2 nanotubes shows a concentration of 5  1015 cm3,
which is very close to the value of 1.3  1016 cm3 that was
calculated using eqn (1). The Cu-dopant (Cu2+ precursor) exists
as the Cu2+ (thermodynamically stable) ion, as conrmed by the
electronic measurements of the number of charge carriers
added per dopant atom. To further evaluate the charge
conduction mechanism, we developed a modied Fowler–
Nordheim (F–N) plot for photogenerated charges. For identi-
fying different regimes of the charge transport and acceptor (or
donor) level below the bandgap, instead of the F–N plot of
 
Iphoto;l  Idark
ln(I/V2) vs. 1/V, we plotted V  ln 2 as a func-
V  Plight  Abs
tion of photon energy (Fig. 7b), where Plight is the incident light
intensity, and Abs is the absorption coefficient (for details, see
ESI Fig. S8†). The data for copper doped nanotubes showed a
photocurrent of 0.5–0.7 eV below the nominal TiO2 bandgap, Fig. 8 (a) Room-temperature magnetization (M) vs. magnetic field (H)
indicating the deep acceptor level for copper. These measure- hysteresis loop for Fe-doped TiO2 nanotube samples; (b) magnetiza-
ments are consistent with the measurements for activation tion–magnetic field loop for undoped TiO2 nanotube samples
energy from I–V–T and STM measurements (Fig. 4b and 5). measured using SQUID, which show clear paramagnetic behavior; (c)
presents zero-field cooled (ZFC) and field cooled (FC) temperature
Moreover, this indicates that even at low uence, photo-
dependent magnetization curves for the Fe-doped sample, which
generated charges are not trapped at the copper dopant sites demonstrates that the ferromagnetic behavior of Fe-doped nanotubes
(interstitial acceptor). These measurements clearly show that results from substitutional Fe magnetic dopants, rather than secondary
the dopant atoms are likely to be incorporated in the lattice, and phases.
contribute towards imparting the desired optical and electronic
properties to doped TiO2 nanotubes.
was parallel or perpendicular to nanotube axes49). For compar-
ison the M–H plot for undoped TiO2 nanotubes is also shown at
Magnetic properties 300 K, which shows nearly linear dependence indicating para-
Magnetic properties of Fe-doped TiO2 nanotube samples were magnetic behavior of undoped samples. It was shown earlier
studied as Fe is frequently chosen as the magnetic dopant when that the paramagnetic behavior of undoped nanotubes did not
pursuing magnetic semiconductors.34,35,48 Fig. 8 presents the change with cooling and stayed paramagnetic up to 1.8 K.50 The
magnetization vs. magnetic eld dependence (M–H) measured temperature dependence of magnetization (M–T) was
at 300 K aer subtracting the diamagnetic background. The measured, both under eld cooled (FC) and zero eld cooled
well-dened hysteresis loop is observed as typical to that of (ZFC) conditions, in order to better understand the origin of
ferromagnetic materials. The saturation magnetization (Ms), ferromagnetism in Fe-doped TiO2 nanotubes.51–53 The FC results
coercive eld (Hc), and remnant magnetization were extracted are obtained by measuring the magnetic moment of the sample
to be 0.22 emu g1, 402 Oe, and 0.028 emu g1, respectively. in a magnetic eld of 1000 Oe during cooling. The ZFC results
These values averaged along different nanotube orientations, as are obtained by rst cooling the sample to 5 K in zero elds and
our sample was a mixture of randomly oriented nanotubes (see then warming it in the same eld as that of the FC measure-
Experimental section), provide further evidence of the desired ment. The ZFC and FC curves are separated below 300 K and
magnetic doping (it was shown previously that magnetization, ZFC magnetization shows stronger temperature dependence
coercive eld, and remnant magnetization of magnetic nano- compared to the FC curve. The divergence between the FC and
tubes varied depending on whether the applied magnetic eld ZFC curves and the lack of blocking temperature peaks indicate

10846 | Nanoscale, 2014, 6, 10839–10849 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper Nanoscale

intrinsic magnetism of TiO2 nanotubes resulting from the Applications


substitution of Ti atoms with Fe atoms, not from clusters of Fe
Metallic oxide nanotubes. By heavily doping TiO2 nanotubes
atoms.51–53 As was mentioned above, XRD studies revealed no
with a shallow dopant Nb, we fabricated transparent nanotube
traces of secondary phases. Also it should be mentioned that the
lms (UV-bandgap 3.1 eV). Two samples with two different
M–H and M–T dependence of our Fe-doped TiO2 nanotubes are
concentrations of Nb were prepared – 1% and 10%. The resis-
similar to those in many other reports on Fe-doped TiO2 lms,
tivity TiO2 sample doped with 1% Nb (Ti0.99Nb0.01O2) increased
nanoparticles, and bulk crystals.34,35,48 These all together
with temperature, as a typical semiconductor (due to reduced
strongly suggest that the ferromagnetic signal from our Fe- carrier concentration), as shown in Fig. 9a, curve 2. However,
doped TiO2 nanotube samples is intrinsic resulting from Fe
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

the heavily doped Ti0.9Nb0.1O2 sample (10% Nb-doped TiO2),


atoms substituting Ti atoms. The mechanism of ferromagne-
showed a reduced resistance to electron transport with
tism in doped transition-metal oxides is still under debate.
temperature, typical for metals (due to reduced electron scat-
However, all existing models explaining ferromagnetism in
tering), Fig. 9a, curve 1. Thus, these results show that Nb-doped
oxide semiconductors (superexchange, double exchange, free-
TiO2 nanotubes can be excellent candidates for use as trans-
carrier-mediated exchange, and exchange through bound
parent conducting oxide nanotube lms. At the high tempera-
magnetic polarons) involve the substitutional nature of
ture region of 100–300 K, the r  T dependence is almost linear,
magnetic dopants replacing Ti atoms in the TiO2 crystal.54–58 In similar to metals, and can be described by r ¼ r0(1 + aDT),
addition, it was shown from density functional theory (DFT) where r0 is the residual resistivity approximately 9.15  104 U
calculations that interstitial magnetic dopants destroy spin
cm, a – is the temperature coefficient 4.7  104 U K1, and DT
polarization.59 Furthermore, using electronic measurements of
– the temperature difference. The carrier concentration n was
Fe-doped TiO2 nanotubes, the nanotubes show increased n-type
almost independent of temperature with only a slight decrease
conductivity (doping created additional ionized electrons,
by 10% when the temperature decreased from 300 K to 20 K
Fig. 5a), which can be expected by substitutional doping with
(Fig. S9†). Mobility, however, increased by a factor of 3.5 with a
Fe2+ ions (of Ti4+ ions) in the TiO2 lattice. However, interstitial
decrease of temperature within the same temperature range
doping with Fe2+ ions would have resulted in an opposite p-type (from 2.9 to 10.3 cm2 V1 s1) which is explained by the
electronic behavior. Therefore, it is reasonable to conclude that reduction of phonon scattering at lower temperatures. We
Fe2+ ions in our TiO2 nanotubes predominantly substitute Ti4+
ions in the crystal lattice.

Fig. 10 (a) Co-doped TiO2 nanotubes (Cu–N and Nb–N) show


infrared absorption of incident light, shown using small bandgaps
Fig. 9 (a) The resistivity for Nb doped nanotubes as a function of determined by the STS spectrum. These photocatalysts can absorb
temperature with two different Nb doping levels – 1% (Ti0.99Nb0.01O2), much more light than the undoped wide-bandgap nanotubes. (b)
curve 2, and 10% (Ti0.9Nb0.1O2), curve 1. Ti0.99Nb0.01O2 shows an Stability of undoped and nitrogen-doped TiO2 nanotube catalysts.
increase of resistivity with temperature as typical to semiconductors. Undoped nanotubes show a strong drop in the catalytic activity, for
By contrast Ti0.99Nb0.1O2 shows a decrease of resistivity with CO2–H2O reduction (artificial photosynthesis), with time. However,
temperature, revealing metallic behavior. (b) Transmittance of 10% Nb nitrogen doped nanotubes show a relatively stable photocatalytic yield
doped TNT samples. Transmittance above 90% was observed in the of selective high-value products, on illumination with simulated solar
visible range for the 10% Nb doped TNT. irradiation (AM1.5 spectrum).

This journal is © The Royal Society of Chemistry 2014 Nanoscale, 2014, 6, 10839–10849 | 10847
View Article Online

Nanoscale Paper

Table 2 Summary of photocatalytic reduction experiments for enabled by a new electrochemical cell design. A variety of
carbon-dioxide to demonstrate higher selectivity and higher stability optical, electronic and magnetic dopants were successfully
for the doped TiO2 nanotubes
incorporated into the hollow nanotubes leading to desired
Production Rate (e mmol cm2 h1) physical properties. Detailed investigations shown here prove
that this versatile method is applicable for tailoring the dopant
Sample Methane Ethane Acetaldehyde Selectivity (cationic, anionic, mono and co-doping) and the doping level
(from low doping to highly-doped semiconductors) over a wide
Undoped 0.014 0.023 0.088 70%
0.5 wt% nitrogen 0.003 0.006 0.060 87 range of materials and physical properties (semiconductor to
metal-like transition). These results can have important impli-
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

cations for the development of new devices and device archi-


tectures for applications in the broad eld of optoelectronics,
conrmed that the transmission obtained in these lms is catalysis, display technologies, and power electronics.
above 90% for the entire visible wavelength range (Fig. 9b), and
demonstrate metal-like conductivity (Fig. 9a). The high trans- Acknowledgements
mittance of these nanotubes (>90% above the bandgap and
10 wt% Nb dopant) allows the demonstration of TCO nano- We would like to thank Zachary Fisk and Ted Grant from the
tubes. The transmittance h was estimated from reectivity University of California–Irvine for their assistance in magneti-
measurements using the following equation: zation measurements using SQUID. Financial support for this
  work was provided by start-up funds from the University of
Rb  Rd
h¼ 1  100% (2) Colorado and NSF Career Award CBET-1351281.
Rs  Rd

where Rb, are measured signals from the blank Ti sheet, Rs – is References
the signal from the sample, and Rd – is the background signal.
Photocatalysis. Using suitable (deep and shallow) mono- 1 B. Oregan and M. Gratzel, Nature, 1991, 353, 737.
dopants and co-dopants (Cu, N, Cu–N, and Nb–N), we fabricated 2 A. L. Linsebigler, G. Q. Lu and J. T. Yates, Chem. Rev., 1995,
photocatalysts with high absorption throughout the visible 95, 735.
wavelengths (for enhanced light absorption, Fig. 10a). These 3 X. B. Chen, L. Liu, P. Y. Yu and S. S. Mao, Science, 2011, 331,
photocatalysts can absorb much more light than the undoped 746.
wide-bandgap nanotubes. Besides the higher (infrared 4 Y. C. Nah, I. Paramasivam and P. Schmuki, ChemPhysChem,
bandgap) absorption, we can easily tune the electronic levels in 2010, 11, 2698.
these doped wide-bandgap nanotubes, to match the desired 5 D. Mocatta, G. Cohen, J. Schattner, O. Millo, E. Rabani and
redox reactions for CO2/H2O for selective photocatalysis60 U. Banin, Science, 2011, 332, 77.
(Fig. 10a and S11†). Photocatalytic reduction of carbon-dioxide, 6 S. C. Erwin, L. J. Zu, M. I. Hael, A. L. Efros, T. A. Kennedy
using simulated AM1.5 solar irradiation (Fig. S10†), was used to and D. J. Norris, Nature, 2005, 436, 91.
demonstrate the higher selectivity (see Table 2) and higher 7 D. J. Norris, A. L. Efros and S. C. Erwin, Science, 2008, 319,
stability (Fig. 10b) of the doped TiO2 nanotubes, using nitrogen 1776.
doping (see ESI†). While the apparent quantum yield (AQY) of 8 A. Boltasseva and H. A. Atwater, Science, 2011, 331, 290.
these doped nanotubes decreases from 0.48% for undoped 9 J. M. Luther, P. K. Jain, T. Ewers and A. P. Alivisatos, Nat.
nanotubes to 0.3% for nitrogen-doped TiO2, the improved Mater., 2011, 10, 361.
selectivity and stability can lead to important applications in the 10 Y. Q. Gai, J. B. Li, S. S. Li, J. B. Xia and S. H. Wei, Phys. Rev.
generation of solar fuels. Moreover, measurements of the elec- Lett., 2009, 102, 036402.
tronic properties of these doped nanotubes reveal no charge 11 W. G. Zhu, X. F. Qiu, V. Iancu, X. Q. Chen, H. Pan, W. Wang,
trapping due to the addition of dopants, and high collection N. M. Dimitrijevic, T. Rajh, H. M. Meyer, M. P. Paranthaman,
efficiency for photogenerated charge carriers (discussion on G. M. Stocks, H. H. Weitering, B. H. Gu, G. Eres and
photo I–V, Fig. 7). Therefore, different light harvesting strategies Z. Y. Zhang, Phys. Rev. Lett., 2009, 103, 226401.
and solar concentrators can also be used to further improve the 12 T. H. Zhou, Y. H. Du, A. Borgna, J. D. Hong, Y. B. Wang,
quantum yield for xing CO2 using sunlight as desired solar J. Y. Han, W. Zhang and R. Xu, Energy Environ. Sci., 2013,
fuels.60 These results indicate the applicability of this new 6, 3229.
doping method for improving the absorption and selectivity of 13 W. J. Ong, L. L. Tan, S. P. Chai, S. T. Yong and
new TiO2 photocatalysts for applications in articial photosyn- A. R. Mohamed, Nanoscale, 2014, 6, 1946.
thesis, improved photocatalytic and photoelectrochemical cells, 14 X. Y. Pan, M. Q. Yang, X. Z. Fu, N. Zhang and Y. J. Xu,
and other devices for harvesting incident solar energy. Nanoscale, 2013, 5, 3601.
15 Q. J. Xiang and J. G. Yu, J. Phys. Chem. Lett., 2013, 4, 753.
Conclusions 16 W. J. Ong, L. L. Tan, S. P. Chai, S. T. Yong and
A. R. Mohamed, ChemSusChem, 2014, 7, 690.
We presented a new method for nanotube doping during elec- 17 J. D. Hong, Y. S. Wang, Y. B. Wang, W. Zhang and R. Xu,
trochemical anodization using alternating current that was ChemSusChem, 2013, 6, 2263.

10848 | Nanoscale, 2014, 6, 10839–10849 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper Nanoscale

18 J. Ran, J. Zhang, J. Yu, M. Jaroniec and S. J. Qiao, Chem. Soc. 40 J. Osorio-Guillen, S. Lany and A. Zunger, Phys. Rev. Lett.,
Rev., 2014, DOI: 10.1039/c3cs60425j. 2008, 100, 036601.
19 M. Ni, M. K. H. Leung, D. Y. C. Leung and K. Sumathy, 41 P. Zawadzki, J. Rossmeisl and K. W. Jacobsen, Phys. Rev. B:
Renewable Sustainable Energy Rev., 2007, 11, 401. Condens. Matter Mater. Phys., 2011, 84, 121203.
20 G. Lui, J. Y. Liao, A. S. Duan, Z. S. Zhang, M. Fowler and 42 Z. Y. Zhang, C. H. Jin, X. L. Liang, Q. Chen and L. M. Peng,
A. P. Yu, J. Mater. Chem. A, 2013, 1, 12255. Appl. Phys. Lett., 2006, 88, 073102.
21 W. J. Ong, M. M. Gui, S. P. Chai and A. R. Mohamed, RSC 43 D. V. Wellia, Q. C. Xu, M. A. Sk, K. H. Lim, T. M. Lim and
Adv., 2013, 3, 4505. T. T. Y. Tan, Appl. Catal., A, 2011, 401, 98.
22 Q. J. Xiang, J. G. Yu and M. Jaroniec, Nanoscale, 2011, 3, 3670. 44 A. R. Bally, E. N. Korobeinikova, P. E. Schmid, F. Levy and
Published on 12 August 2014. Downloaded by Universitätsbibliothek Bern on 05/09/2014 14:04:08.

23 Y. H. Hu, H. Wang and B. Hu, ChemSusChem, 2010, 3, 782. F. Bussy, J. Phys. D: Appl. Phys., 1998, 31, 1149.
24 G. K. Mor, O. K. Varghese, M. Paulose, K. Shankar and 45 A. Roldan, M. Boronat, A. Corma and F. Illas, J. Phys. Chem.
C. A. Grimes, Sol. Energy Mater. Sol. Cells, 2006, 90, 2011. C, 2010, 114, 6511.
25 M. Paulose, K. Shankar, S. Yoriya, H. E. Prakasam, 46 C. Di Valentin, E. Finazzi, G. Pacchioni, A. Selloni,
O. K. Varghese, G. K. Mor, T. A. Latempa, A. Fitzgerald and S. Livraghi, M. C. Paganini and E. Giamello, Chem. Phys.,
C. A. Grimes, J. Phys. Chem. B, 2006, 110, 16179. 2007, 339, 44.
26 C. A. Grimes, J. Mater. Chem., 2007, 17, 1451. 47 A. A. Talin, F. Leonard, B. S. Swartzentruber, X. Wang and
27 J. M. Macak, H. Hildebrand, U. Marten-Jahns and S. D. Hersee, Phys. Rev. Lett., 2008, 101, 076802.
P. Schmuki, J. Electroanal. Chem., 2008, 621, 254. 48 N. H. Hong, J. Sakai, W. Prellier and A. Ruyter, J. Phys. D:
28 Y. Alivov, M. Pandikunta, S. Nikishin and Z. Y. Fan, Appl. Phys., 2005, 38, 816.
Nanotechnology, 2009, 20, 225602. 49 V. M. Prida, M. Hernandez-Velez, K. R. Pirota, A. Menendez
29 H. Tang, K. Prasad, R. Sanjines, P. E. Schmid and F. Levy, J. and M. Vazquez, Nanotechnology, 2005, 16, 2696.
Appl. Phys., 1994, 75, 2042. 50 Y. Alivov, T. Grant, C. Capan, W. Iwamoto, P. G. Pagliuso and
30 Y. Furubayashi, N. Yamada, Y. Hirose, Y. Yamamoto, S. Molloi, Nanotechnology, 2013, 24, 275704.
M. Otani, T. Hitosugi, T. Shimada and T. Hasegawa, J. 51 G. J. Yang, D. Q. Gao, Z. H. Shi, Z. H. Zhang, J. Zhang,
Appl. Phys., 2007, 101, 093705. J. L. Zhang and D. S. Xue, J. Phys. Chem. C, 2010, 114, 21989.
31 T. Hitosugi, A. Ueda, S. Nakao, N. Yamada, Y. Furubayashi, 52 D. Y. Inamdar, A. D. Lad, A. K. Pathak, I. Dubenko, N. Ali and
Y. Hirose, T. Shimada and T. Hasegawa, Appl. Phys. Lett., S. Mahamuni, J. Phys. Chem. C, 2010, 114, 1451.
2007, 90, 212106. 53 A. Manivannan, P. Dutta, G. Glaspell and M. S. Seehra, J.
32 T. Hitosugi, A. Ueda, Y. Furubayashi, Y. Hirose, S. Konuma, Appl. Phys., 2006, 99, 08M110.
T. Shimada and T. Hasegawa, Jpn. J. Appl. Phys., Part 2, 2007, 54 M. H. F. Sluiter, Y. Kawazoe, P. Sharma, A. Inoue, A. R. Raju,
46, L86. C. Rout and U. V. Waghmare, Phys. Rev. Lett., 2005, 94,
33 N. H. Hong, J. Sakai, W. Prellier, A. Hassini, A. Ruyter and 187204.
F. Gervais, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 55 J. M. D. Coey, A. P. Douvalis, C. B. Fitzgerald and
70, 195204. M. Venkatesan, Appl. Phys. Lett., 2004, 84, 1332.
34 F. Lin, D. M. Jiang, Y. Lin and X. M. Ma, Phys. B, 2008, 403, 56 A. C. Tuan, J. D. Bryan, A. B. Pakhomov, V. Shutthanandan,
2193. S. Thevuthasan, D. E. McCready, D. Gaspar,
35 K. Y. S. Chan and G. K. L. Goh, Thin Solid Films, 2008, 516, M. H. Engelhard, J. W. Rogers, K. Krishnan, D. R. Gamelin
5582. and S. A. Chambers, Phys. Rev. B: Condens. Matter Mater.
36 M. Ni, M. K. H. Leung, D. Y. C. Leung and K. Sumathy, Phys., 2004, 70, 054424.
Renewable Sustainable Energy Rev., 2007, 11, 401. 57 P. V. Radovanovic and D. R. Gamelin, Phys. Rev. Lett., 2003,
37 O. K. Varghese, D. W. Gong, M. Paulose, C. A. Grimes and 91, 157202.
E. C. Dickey, J. Mater. Res., 2003, 18, 156. 58 M. S. Park, S. K. Kwon and B. I. Min, Phys. Rev. B: Condens.
38 R. D. Shannon, Acta Crystallogr., Sect. A: Found. Crystallogr., Matter Mater. Phys., 2002, 65, 161201.
1976, 32, 751. 59 W. T. Geng and K. S. Kim, Phys. Rev. B: Condens. Matter
39 Y. Furubayashi, T. Hitosugi, Y. Yamamoto, K. Inaba, Mater. Phys., 2003, 68, 125203.
G. Kinoda, Y. Hirose, T. Shimada and T. Hasegawa, Appl. 60 V. Singh, I. J. C. Beltran, J. C. Ribot and P. Nagpal, Nano Lett.,
Phys. Lett., 2005, 86, 252101. 2014, 14, 597.

This journal is © The Royal Society of Chemistry 2014 Nanoscale, 2014, 6, 10839–10849 | 10849

You might also like