You are on page 1of 7

Article

Cite This: J. Phys. Chem. C 2017, 121, 24809-24815 pubs.acs.org/JPCC

Assessing Oxygen Vacancies in Bismuth Oxide through EELS


Measurements and DFT Simulations
Pau Torruella,*,†,‡ Catalina Coll,†,‡ Gemma Martín,†,‡ Lluís López-Conesa,†,‡,§ María Vila,∥
Carlos Díaz-Guerra,∥ María Varela,∥,⊥ María Luisa Ruiz-González,# Javier Piqueras,∥ Francesca Peiró,†,‡
and Sònia Estradé†,‡

LENS-MIND, Departament d’Enginyeries: Electrònica, ‡Institute of Nanoscience and Nanotechnology (IN2UB), and §TEM-MAT,
CCiT, Universitat de Barcelona, Barcelona 08028, Spain

Departamento de Física de Materiales, Facultad de Ciencias Físicas, ⊥Instituto de Magnetismo Aplicado, Facultad de Ciencias Físicas,
and #Departamento de Química Inorgánica, Facultad de Químicas, Universidad Complutense de Madrid, Madrid 28040, Spain
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

*
S Supporting Information
Downloaded via UNIV OF PESHAWAR on June 29, 2022 at 14:46:05 (UTC).

ABSTRACT: Pioneering electron energy loss spectroscopy


(EELS) measurements of α-Bi2O3 are performed on three
samples obtained through different synthesis methods.
Experimental low-loss and core-loss EELS spectra are acquired.
By combining them with detailed structural characterization
and Density Functional Theory (DFT) simulations, we are
able to detect and evaluate the presence of oxygen vacancies in
the samples. This type of information has not been accessed
previously from EELS data in bismuth oxide, because high-
resolution EELS spectra or how vacancies reflect in Bi2O3
spectra were unreported. This novel measurement is further
validated through comparison with photoluminescence data.
Therefore, the technique has the ability to probe oxygen
vacancies in Bi2O3 at an unprecedented resolution, which might allow solving material science and technological issues related to
this material.

■ INTRODUCTION
Bismuth oxide is a polymorph with four main crystallographic
spectra, while information on the oxidation state, coordination,
or vacancy presence of a given element in the sample can be
phases: the monoclinic α-phase, which is stable at low obtained from core-loss spectra by analyzing its energy loss
temperature; the tetragonal β and face-centered cubic γ phases, near edge structure (ELNES) features.17 The latter properties
which are metastable; and the δ-phase, which is stable at high are of particular importance when dealing with bismuth oxide,
temperatures.1,2 It is a widely used material in the state-of-the- because its high ionic conductivity is discussed in terms of
art of many technological fields such as gas sensors,3,4 optical oxygen vacancy ordering, generation, and transport.9,10,18,19
devices,5,6 and solid oxide fuel cells.7,8 However, the functional However, obtaining this information through the ELNES fine
properties of Bi2O3 are strongly dependent on its crystalline structure is not straightforward and requires previous knowl-
phase. As an example, the conduction mechanism of α-Bi2O3 is edge on how microscopic properties of the given material alter
mainly electronic, whereas for δ-Bi2O3 it is mainly ionic.9,10 the spectra. This can only be obtained from reference samples
Moreover, with the advent of nanotechnology, a huge or simulations.20,21 Surprisingly, reported EELS measurements
miscellany of Bi2O3 nanostructures is being reported, from of Bi2O3 only use low-loss spectra to determine band gap
nanocoatings to nanowires, nanotubes, or hollow nanoparticles, energy,19 and none, to the best of our knowledge, presents any
each with its own applications and virtues.11−15 Therefore, spectrum from the core-loss region.
when tailoring these complex nanostructures, a way to assess The present work aims to bridge this gap of knowledge and
Bi2O3 phase, defects, and other structural properties with high provide Bi2O3 core-loss EELS data, while linking experimental
resolution is mandatory. In this context, the use of a high- and simulation EELS results to microscopic properties of three
resolution transmission electron microscope (TEM) provides a different Bi2O3 samples, bismuth oxide nanowires synthesized
great tool to evaluate crystalline properties at the nanoscale.16 by thermal evaporation,22 commercial Bi2O3 powder, and a
Additionally, when coupled with electron energy loss spectros-
copy (EELS), it can also probe chemical and electronic Received: June 27, 2017
properties of a given material at high resolution. In particular, Revised: September 29, 2017
the band structure of a material is related to low-loss EELS Published: October 24, 2017

© 2017 American Chemical Society 24809 DOI: 10.1021/acs.jpcc.7b06310


J. Phys. Chem. C 2017, 121, 24809−24815
The Journal of Physical Chemistry C Article

Figure 1. Optical characterization. (A) Normalized Raman spectra representative of the three kinds of Bi2O3 samples investigated. (B) Normalized
PL spectra from the three Bi2O3 investigated samples.

Bi2O3 ceramic23 obtained in the form of a sintered pellet from low-loss and core-loss EELS spectra simultaneously. The latter
this powder. The crystalline structure of the samples has been were focused in the 520−560 eV energy loss range, were the O
fully characterized through scanning-TEM (STEM) and Raman K edge can be observed. The dual acquisition enables aligning
spectroscopy. Careful EELS measurements have been per- core-loss spectra with the zero-loss peak, therefore making the
formed in low-loss and core-loss regimes. To understand the absolute energy-loss position of different features of the peaks
EELS features from the obtained spectra, the measurements are reliable. Single spectra for each sample were obtained by
supported with density functional theory (DFT) simulations summing larger spectrum images after zero-loss peak alignment.
that can calculate the band structure and the ELNES features of These spectrum images were acquired with atomic resolution,
a given crystalline structure.24 Finally, photoluminescence (PL) and the corresponding coacquired HAADF images (Figure S1)
measurements are performed, and the results are compared to allow identification of the zone axis of the crystal during the
the conclusions obtained from the novel EELS measurements. acquisition as well as confirm the lack of beam damage. They


were also acquired so that there were no carbon contributions
METHODS either from the TEM grid or from contamination.
Theoretical calculations of the electronic structure of α-Bi2O3
Commercial Bi2O3 powders (99.999%) were purchased from were carried out using the linearized augmented plane wave
Sigma-Aldrich. To obtain the Bi2O3 ceramics, these powders (LAPW) method based on DFT, implemented on the
were compacted under compressive load, forming disk-shaped WIEN2k26,27 ab initio simulation package. The DFT calculation
samples of about 1 mm thickness and 7 mm diameter. These was performed considering the local density approximation28,29
samples were then annealed in air at 750 °C for 10 h.23 Bi2O3 (LDA). The atomistic input model was built according to the
nanowires were grown mixing 80 wt % Bi powder (Goodfellow, crystallographic information on the unit cell (UC) from α-
99.997%) and 20 wt % Er2O3 powder (STREM, 99.9%). This Bi2O3 structure as reported in ref 30, which has a P21c space
mixture was compressed to form disk-shaped pellets, which group (5.8486, 8.1661, and 7.5097 Å lattice parameters and
were then annealed at 800 °C for 4 h in a horizontal tube beta angle of 113°, giving five nonequivalent atomic positions).
furnace under Ar flow. This treatment led to the growth of a The cutoff energy of the simulation was set at −11.9 Ry,
high density of α-Bi2O3 nanowires.22 It must be pointed out defining the following valence and semicore electrons: 2s2 and
that Er was not incorporated into the nanowires.25 Its role is to 2p4 for oxygen and 4f14, 5d10, 6p3, and 6s2 for bismuth. DFT
alter the oxidation kinetics, favoring the formation of a high self-consistency cycles were performed until the total energy
density of nanostructures. was converged to 10−5 eV and the residual forces on atoms
Micro-Raman and micro-PL measurements were carried out were below 0.01 eV/Å with 1000 k-points. The plane-wave
at room temperature in a Horiba Jovin-Ybon LabRAM HR800 basis set cutoff was set as RKmax = 7.


system. To collect Raman spectra, the samples were excited by
a 633 nm He−Ne laser on an Olympus BX 41 confocal
microscope with a 100× objective. A 325 nm He−Cd laser and RESULTS
a 40× objective were used for PL measurements. Normalized Raman spectra representative of the Bi 2O 3
Scanning TEM (STEM) high angle annular dark field powders, ceramics, and nanowires are shown in Figure 1A.
(HAADF) and EELS measurements from the samples were No significant differences, in terms of peak positions and
acquired in a JEOL ARM200cF TEM operated at 200 kV and bandwidths, were found in Raman spectra of the different
equipped with a cold field emission fun (c-FEG) and a GATAN samples. Raman bands appear peaked at 84, 119, 139, 152, 184,
Quantum GIF EEL spectrometer. A Gatan double tilt cryo- 210, 280, 313, 411, 448, and 526 cm−1, which are all
holder (model 636) was used for the ceramic sample to characteristic of the α-Bi2O3 phase.31−33
minimize electron beam damage, with a working temperature of PL measurements, Figure 1B, show three narrow, well-
97 K. The collection and convergence semiangles were 27.78 resolved emission bands, peaked at 1.86, 1.95, and 2.05 eV.
and 35 mrad, respectively, throughout all of the EELS Besides those, two weaker and broader emissions, respectively
measurements. These measurements were performed using centered near 2.10 and 2.27 eV, can be also appreciated. The
the dual-EELS acquisition mode, which allows acquiring both PL intensity varies strongly among the specimens. The Bi2O3
24810 DOI: 10.1021/acs.jpcc.7b06310
J. Phys. Chem. C 2017, 121, 24809−24815
The Journal of Physical Chemistry C Article

Figure 2. STEM-HAADF imaging. (A) Low magnification STEM-HAADF image of a group of Bi2O3 nanocrystals from the reference powder. (B)
Atomic resolution STEM-HAADF image of a Bi2O3 nanocrystal seen along the [001]α-Bi2O3 zone axis. (C) Low magnification STEM-HAADF
image of the FIB-prepared Bi2O3 pellet. The top bright layer seen in the image is due to the preparation of the sample. (D) Atomic resolution
STEM-HAADF image from the pellet sample seen along the [001] zone axis of the α-phase. (E) Low magnification TEM image of the nanowire
sample. (F) STEM-HAADF image of an α-phase nanowire along the [010] zone axis of the α-phase.

ceramic shows the maximum PL intensity, which halves for the sizes varying from 50 to 500 nm, depicted in Figure 2A, which
nanowires, and is 3 times weaker in the powder. In addition, the consisted of Bi2O3 in the α-phase, seen in the [001] zone axis in
relative intensity of the 2.10 and 2.27 eV bands varies for each Figure 2B. As for the pellet, the sample was prepared by the
sample, with the ceramic showing the highest 2.27 eV relative standard FIB lift-out technique.34 TEM characterization
emission and the commercial powder the lowest. Moreover, the evidenced that it consisted of large (several micrometers) α-
peak positions of the above-mentioned three narrow emissions phase Bi2O3 crystals (Figure 2C,D). The Bi2O3 nanowires were
observed centered between 1.86 and 2.05 eV in PL spectra of
prepared by solution and sonication in hexane, followed by
powders and nanowires appear shifted 0.3 eV toward higher
pouring a drop of the solution onto a TEM holey carbon grid.
energies in the ceramic. Additionally, PL spectra from the latter
sample show a weak and broad new emission at about 3.0 eV as As can be seen in the low-magnification TEM image (Figure
well as an intense band centered near 1.72 eV. 2E), they were found to form hierarchical structures and also
The commercial powder sample was prepared for TEM consisted mainly of bismuth oxide in the α-phase (Figure 2F).
observation by the standard dilution in hexane and drop Sparse, smaller nanowires of the metastable β-Bi2O3 nanowires
pouring onto a TEM holey carbon grid. The TEM structural were also observed, but as Raman measurements show, they are
analysis showed that it was comprised of crystals with lateral not statistically relevant.
24811 DOI: 10.1021/acs.jpcc.7b06310
J. Phys. Chem. C 2017, 121, 24809−24815
The Journal of Physical Chemistry C Article

The EELS measurements at 200 kV for each sample are the spectra. To discuss these differences, the main features of
shown in Figure 3. Clear differences between samples appear in the spectra have been labeled. The core-loss spectra show a
main peak at 534.6 eV (b1), which is at the same position for all
samples. The onset of this peak shows a shoulder at 532.5 eV
(a1) that varies in height when comparing the commercial
powder sample with the other two, which were thermally
treated. The a1 intensity goes from being almost 25% lower
than the b1 peak in the commercial powder to being equal in
the nanowire and the sintered pellet. Moreover, the powder
shows some additional peaks in the continuum region of the
spectrum at 540.2 eV (c1) and 545.7 eV (d1).
To understand information from the EEL spectra, an initial
DFT simulation for the standard unit cell was carried out.
Importantly, the LDA approximation used in the calculations is
known to not reliably calculate absolute onset energies for
EELS edges.35 Because of this, a shift of 1.2 eV will be applied
to all simulations. This value has been chosen so that the b1
peak energy of the experimental spectrum of the commercial
powder matches the maximum of the simulated edge from the
unit cell, de facto taking the commercial powder as a reference
sample. The simulated O K edge from the unit cell is shown in
Figure 4A, where the energies of the experimental features
(a1,b1,c1,d1) are also labeled. An excellent match with the
experimental spectrum from the reference Bi2O3 powder can be
appreciated: the simulation shows all of the features of the
experimental spectra, with only small differences in the position
of the c1 peak (at a lower energy in the simulation) and the d1
peak (at a higher energy). Remarkably, this is the first reported
experimental spectrum of α-Bi2O3, with additional confirmation
Figure 3. Experimental EELS. Core-loss and low-loss from the three from the simulation, and will be made available at the EELS
investigated samples. Note that each spectrum has been divided by the DataBase service.36 In contrast, the a1 shoulder is much more
intensity at its maximum. The main peaks of the spectra have been intense in the pellet and nanowire spectra, as well as both
marked and labeled: a1 = 532.5 eV, b1 = 534.6 eV, c1 = 540.2 eV, d1 = showing a steeper onset than the reference sample. The good
545.7 eV, a2 = 11.0 eV, b2 = 18.7 eV, c2 = 29.5 eV. match between the powder sample and the UC simulation,
while small discrepancies show up for nanowire and pellet,
points to significant changes between these samples.

Figure 4. DFT simulations of EEL spectra. (A) Simulated O K edge spectrum for α-Bi2O3 (plotted in black). The contributions from each
nonequivalent oxygen in the unit cell are also plotted. An inset with the α-Bi2O3 unit cell seen along the [010] zone axis shows the position of each
of these oxygens in blue, red, and green (the Bi atoms are shown in purple). The energy positions of the spectral features identified in the
experimental spectra are also shown as vertical lines. (B) Simulated core-loss and (C) low-loss spectra for α-Bi2O3 and for α-Bi2O3 with one oxygen
less for every two unit cells.

24812 DOI: 10.1021/acs.jpcc.7b06310


J. Phys. Chem. C 2017, 121, 24809−24815
The Journal of Physical Chemistry C Article

An important fact to consider is that in the UC of α-Bi2O3


there are 6 oxygen atoms located at three different non-
equivalent crystallographic positions (two in each one), each
one seeing a different electronic environment in the structure.
These positions will be referred as OI for x = 0.2337a, y =
0.4533b, z = 0.1266c, and equivalent positions; OII for x =
0.265a, y = 0.0294b, and z = 0.0115c; and OIII for x = 0.7783a, y
= 0.3037b, and z = 0.2080c, where a, b, and c are the unit cell
parameters. The contributions of oxygen atoms in OI, OII, and
OIII can also be extracted from the simulated UC, and are
shown in Figure 4A. The total spectra can be evaluated as the
sum of these contributions weighted by their multiplicity (the
number of atoms in each position and equivalents, two for all in
this case). In Figure 4A, it is clear that these contributions are
different. In fact, the most notable difference is also the height
of the a1 shoulder, OII having the lowest, and the b1 peak
distance respect to the edge onset, OII having the highest. This
fact hints that the difference in the experimental spectra might
be due to an imbalance in the occupation of the three oxygen Figure 5. DFT simulations of oxygen vacancies. O K energy shift for
each of the DFT simulations.
sites. To further investigate this, simulations of unit cells with
oxygen vacancies were undertaken.


A few considerations must be taken into account when
removing an atom from a unit cell in a DFT simulation. Even DISCUSSION
though a unit cell with one missing oxygen atom might seem a
rather small defect in a big crystal, the simulations are As far as the cationic structure is concerned, the samples show
performed with periodic boundary conditions, which means no significant differences, all of them being mainly α-Bi2O3.
actually an infinite crystal with one less oxygen in each unit cell However, the experimental EEL spectra show clear differences,
is being simulated. Because of this, if the vacancy proportion and their origin must be elucidated.
wants to be kept low, a large “supercell” consisting of multiple At first, one could think that the origin of these small
unit cells in which only one oxygen vacancy is being introduced differences could be found in the orientation dependence of
must be simulated. Three additional DFT simulations were EEL spectra41 because the samples were not in the same zone
performed to evaluate the oxygen vacancy concentration impact axis during the acquisition. To explore this possibility, UC
on EELS spectra: a 2 × 1 × 1 “supercell”, which will be referred simulations for the relevant zone axis were carried out. As can
to as V simulation, with one oxygen atom missing, a 3 × 1 × 1 be seen in Figure S2, they show no significant differences in our
experimental conditions, so this factor can be ruled out.
“supercell” with one vacancy, and a 3 × 1 × 1 “supercell” with
Additionally, because of the known dependence on oxygen
two vacancies. Considering that the unit cell of Bi2O3 has 12
vacancies of bismuth oxide properties, it is reasonable to
oxygen atoms in total, vacancy concentrations are 1/24 =
suspect that they might have a role in these measurements. To
4.17%, 1/36 = 2.78%, and 2/36 = 5.56%, respectively. The
perform a rigorous analysis of the core-loss spectra from the
crystallographic position in which the vacancy is formed must
nanowire and pellet sample in terms of oxygen vacancies, each
be carefully considered. The vacancy formation energy for each
configuration of missing oxygen atoms for increasingly large
position was calculated yielding similar values for each one
supercells should be simulated. This approach is not feasible in
within the reliability of the simulation. Because of this,
terms of computational time. In this regard, it must be kept in
vacancies in the OII site were chosen, as their contribution
mind that bismuth is one of the heaviest stable elements on the
can account for the observed discrepancies between the
periodic table, with 83 electrons, 29 of which have been treated
experimental spectra of the powder and the nanowire/pellet
rigorously within the DFT formalism (orbitals 4f to 6s as stated
sample. The same offset of 1.2 eV applied to the UC simulation
in the Methods), and therefore costly to simulate. However, on
is applied to the vacancy simulations calculated spectra.
the basis of the knowledge that the previous simulations have
UC and V simulations were compared both in the core-loss
yielded, that vacancies shift the O K edge, and that b1, a1
(Figure 4B) and in the low-loss regime (Figure 4C).
features vary for different symmetry sites, the following simple
Remarkably, no major differences appear in the low-loss
model can be inferred: Iexp(ΔE) = k1IOI(ΔE − s) + k2IOII(ΔE −
spectra, apart from a small peak at 1.9 eV. Similarly, the
experimental low-loss spectra barely change between each s) + k3IOIII(ΔE − s). Properly normalized, the ki parameters
sample, and unfortunately at 1.9 eV the zero-loss peak overlaps represent occupancies of the different oxygen sites, and the
with the loss function of the samples too much to see any peak parameter s accounts for the energy shift that the V simulation
of this kind in experimental data. Yet the most interesting demonstrates. This model can now be used to fit the
feature in this comparison is the appearance of a 1.3 eV shift to experimental data and give some grounds for discussion on
lower energies of the O K edge in the V simulation. This seems the presence and nature of oxygen vacancies on the material.
very reasonable if one bears in mind that this shift has also been Importantly, it must be kept in mind that the positions of the
seen in EELS spectra of other transition metals, such as Fe, Co, simulated oxygen spectra have been calibrated by using the
or Mn, when they are reduced.37−40 By comparing the O K commercial powder sample, and therefore any property
EELS edge position of all of the simulations (Figure 5), it is calculated by this method will be relative to the commercial
clear that the energy shift is proportional to the vacancy powder. However, this seems reasonable because it is the purest
concentration. form in which we can obtain this material.
24813 DOI: 10.1021/acs.jpcc.7b06310
J. Phys. Chem. C 2017, 121, 24809−24815
The Journal of Physical Chemistry C Article

The model was implemented in Python using a Scipy curve


fitting routine,42 and the results of the fit with the experimental
■ CONCLUSIONS
Structural characterization and PL measurements have been
are summarized in Figure 6. The obtained ki and s parameters carried out. Core-loss EELS spectra of Bi2O3 have been
reported for the first time for three different α-Bi2O3 samples.
These measurements have been interpreted with the help of
DFT simulations of the α-Bi2O3 unit cell and of super cells
containing oxygen vacancies. The DFT simulations establish
the relation between oxygen vacancy concentration and O K
edge onset. The EELS spectra analysis confirms an increase in
oxygen vacancies in the nanowires and the pellet sample,
occupying preferably the OII crystallographic site. Remarkably,
the PL results are in good agreement with these findings.
Therefore, the current work allows ELNES analysis to be an
efficient tool to assess oxygen vacancies in Bi2O3, extracting a
type of information that was not possible to access previously
from EEL spectra.


*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.jpcc.7b06310.
HAADF images coacquired from the spectra in Figure 3;
Bi2O3 O K edge orientation dependence simulations;
vacancy formation energies obtained from the DFT
calculations; and fitting description and obtained
parameters from Figures 5 and 6 (PDF)

Figure 6. EELS fitting. Fitting of the core-loss experimental spectra


from each α-Bi2O3 sample with the weighted contributions of each
nonequivalent oxygen atom and an energy shift.
■ AUTHOR INFORMATION
Corresponding Author
*E-mail: ptorruella@el.ub.edu.
ORCID
(detailed in the Supporting Information) show that the Pau Torruella: 0000-0002-6864-4000
occupancy of OI and OIII barely changes in the three samples.
Nonetheless, a small shift of approximately 0.45 eV for the Catalina Coll: 0000-0002-3035-2555
nanowire sample and 0.58 eV for the pellet is introduced, and Lluís López-Conesa: 0000-0003-1456-079X
the occupation OII drops to zero for these samples. Notes
Furthermore, from Figure 5, the linear relation between O K The authors declare no competing financial interest.
edge energy shift and vacancy concentration can be calculated
by a linear regression, yielding an r-value of 0.97 (details in the
Supporting Information). From the regression values, the O K
■ ACKNOWLEDGMENTS
This work has been carried out in the frame of the Spanish
shift value of the nanowire and pellet sample can be translated research projects MAT2013-41506-P, MAT2016-79455-P,
into approximately 1.2% of vacant oxygen sites in the nanowires MAT2013-48628-R, MAT2015-65274-R, MAT2015-066888-
and 1.6% in the pellet sample, relative to the powder. These C3-3-R, CSD2009-0013, and FIS2013-46159-C3-3-P. It was
results would indicate that, although the commercial powder also supported by the Spanish Agencia Estatal de Investigación
has a certain density of vacancies, they are uniformly distributed (AEI) and the European Regional Development Fund (ERDF).
in all crystallographic positions, while the ceramics and Catalan Government support from the SGR2014-672 and
nanowires have a higher concentration of vacancies preferably 2014-SGR-1015 projects is also acknowledged. Financial
located at the OII site. support from the ERC PoC MAGTOOLS is acknowledged.
The findings are in good agreement with PL results. Previous Measurements were performed in the Centro Nacional de
luminescence studies of α-Bi2O3 ceramic samples annealed in Microscopiá Electrónica (CNME) at the Universidad Complu-
different atmospheres had shown that a 2.10 eV band is related tense de Madrid (UCM) and also at the Centres Cientifico- ́
to oxygen vacancies.23 Other works reported that the emission Tècnics of the Universitat de Barcelona (CCIT-UB). The DFT
in the low-energy spectral range (≈below 2.1 eV) can also be simulations were carried out in the LUSITANIA II super-
attributed to oxygen vacancies that form defect donor states,43 computer located in Cáceres, a node of the Spanish
and that the higher is the emission intensity in this spectral supercomputing net (Red Española de Supercomputación-
range, the greater is the vacancy density.44 Hence, the higher RES).
PL intensity observed in ceramics and nanowires, as compared
to that measured in powder samples, as well as the higher
relative intensity of the PL emissions in the low-energy spectral
■ REFERENCES
(1) Harwig, H. A.; Gerards, A. G. The Polymorphism of Bismuth
range, support the fact that our EELS results reveal a higher Sesquioxide. Thermochim. Acta 1979, 28, 121−131.
concentration of oxygen vacancies in the studied α-Bi2O3 (2) Harwig, H. A. On Structure of Bismuth Sesquioxide: The A, B, Γ
nanowires and ceramics. and Δ-Phase. Z. Anorg. Allg. Chem. 1978, 444, 151−166.

24814 DOI: 10.1021/acs.jpcc.7b06310


J. Phys. Chem. C 2017, 121, 24809−24815
The Journal of Physical Chemistry C Article

(3) Gou, X.; Li, R.; Wang, G.; Chen, Z.; Wexler, D. Room- (25) Vila, M.; Díaz-Guerra, C.; Piqueras, J. Laser Irradiation-Induced
Temperature Solution Synthesis of Bi2O3 Nanowires for Gas Sensing A to Δ Phase Transformation in Bi2O3 Ceramics and Nanowires. Appl.
Application. Nanotechnology 2009, 20 (49), 495501. Phys. Lett. 2012, 101 (7), 071905.
(4) Cabot, A.; Marsal, A.; Arbiol, J.; Morante, J. R. Bi2O3 as a (26) Blaha, P.; Schwarz, K.; Madsen, G.; Kvasicka, D.; Luitz, J.
Selective Sensing Material for NO Detection. Sens. Actuators, B 2004, WIEN2K, an Augmented Plane Wave + Local Orbital Program for
99 (1), 74−89. Calculating Crystal Properties; Technische Universität Wien: Germany,
(5) Vila, M.; Díaz-guerra, C.; Lorenz, K.; Piqueras, J.; Alves, E.; 2001.
Nappini, S.; Magnano, E. Structural and Luminescence Properties of (27) Blaha, P.; Schwarz, K.; Sorantin, P. Full-Potential, Linearized
Eu and Er Implanted Bi 2 O 3 Nanowires for Optoelectronic Augmented Plane Wave Programs for Crystalline Systems. Comput.
Applications. J. Mater. Chem. C 2013, 1, 7920−7929. Phys. Commun. 1990, 59 (2), 399−415.
(6) Eberl, J.; Kisch, H. Visible Light Photo-Oxidations in the (28) Hohenberg, P.; Kohn, W. Inhomogenous Electron Gas. Phys.
Presence of A-Bi2O3. Photochem. Photobiol. Sci. 2008, 7, 1400. Rev. 1964, 136 (3), 864−871.
(7) Wachsman, E. D.; Lee, K. T. Lowering the Temperature of Solid (29) Kohn, W.; Sham, L. J. Self-Consistent Equations Including
Oxide Fuel Cells. Science (Washington, DC, U. S.) 2011, 334, 935. Exchange and Correlation Effects. Phys. Rev. 1965, 140 (4), 1133−
(8) Zhong, G. H.; Wang, J. L.; Zeng, Z. Ionic Transport Properties in 1139.
Doped Δ-Bi2O3. J. Phys.: Conf. Ser. 2006, 29, 1−5. (30) Kalinchenko, F. V.; Borzenkova, M. P.; Novoselova, A. V. The
(9) Takahashi, T.; Iwahara, H.; Arao, T. High Oxide Ion Conduction BiF3-Bi2O3 System. Russ. J. Inorg. Chem. 1981, 26, 118−120.
in Sintered Oxides of the System Bi2O3-Y2O3. J. Appl. Electrochem. (31) Betsch, R. J.; White, W. B. Vibrational Spectra of Bismuth Oxide
1975, 5, 187−195. and the Sillenite-Structure Bismuth Oxide Derivatives. Spectrochim.
(10) Shuk, P.; Wiemhöfer, H. D.; Guth, U.; Göpel, W.; Greenblatt, Acta Part A Mol. Spectrosc. 1978, 34 (5), 505−514.
(32) Denisov, V. N.; Ivlev, A. N.; Lipin, A. S.; Mavrin, B. N.; Orlov,
M. Oxide Ion Conducting Solid Electrolytes Based on Bi2O3. Solid
V. G. Raman Spectra and Lattice Dynamics of Single-Crystal A-Bi2O3.
State Ionics 1996, 89 (1996), 179−196.
J. Phys.: Condens. Matter 1997, 9 (23), 4967−4978.
(11) Niu, K. Y.; Park, J.; Zheng, H.; Alivisatos, a. P. Revealing
(33) Narang, S. N.; Patel, N. D.; Kartha, V. B. Infrared and Raman
Bismuth Oxide Hollow Nanoparticle Formation by the Kirkendall
Spectral Studies and Normal Modes of A-Bi2O3. J. Mol. Struct. 1994,
Effect. Nano Lett. 2013, 13 (11), 5715−5719. 327 (2−3), 221−235.
(12) Kumari, L.; Lin, J.; Ma, Y. One-Dimensional Bi2O3 Nanohooks: (34) Giannuzzi, L. A.; Kempshall, B. W.; Schwarz, S. M.; Lomness, J.
Synthesis, Characterization and Optical Properties. J. Phys.: Condens. K.; Prenitzer, B. I.; Stevie, F. A. FIB Lift-Out Specimen Preparation
Matter 2007, 19, 406204. Techniques. Introduction to Focused Ion Beams; Kluwer Academic
(13) Qiu, Y.; Liu, D.; Yang, J.; Yang, S. Controlled Synthesis of Publishers: Boston, MA, 2005; pp 201−228.
Bismuth Oxide Nanowires by an Oxidative Metal Vapor Transport (35) Hamann, D. R.; Muller, D. A. Absolute and Approximate
Deposition Technique. Adv. Mater. 2006, 18 (19), 2604−2608. Calculations of Electron-Energy-Loss Spectroscopy Edge Thresholds.
(14) Sarma, B.; Jurovitzki, A. L.; Smith, Y. R.; Mohanty, S. K.; Misra, Phys. Rev. Lett. 2002, 89 (12), 126404−4.
M. Redox-Induced Enhancement in Interfacial Capacitance of the (36) Ewels, P.; Sikora, T.; Serin, V.; Ewels, C. P.; Lajaunie, L. A
Titania Nanotube/bismuth Oxide Composite Electrode. ACS Appl. Complete Overhaul of the Electron Energy-Loss Spectroscopy and X-
Mater. Interfaces 2013, 5 (5), 1688−1697. Ray Absorption Spectroscopy Database: Eelsdb.eu. Microsc. Microanal.
(15) Mahmoud, W. E.; Al-Ghamdi, A. A. Synthesis and Properties of 2016, 22, 717−724.
Bismuth Oxide Nanoshell Coated Polyaniline Nanoparticles for (37) Schmid, H. K.; Mader, W. Oxidation States of Mn and Fe in
Promising Photovoltaic Properties. Polym. Adv. Technol. 2011, 22 Various Compound Oxide Systems. Micron 2006, 37 (5), 426−432.
(6), 877−881. (38) López-Ortega, A.; Roca, A. G.; Torruella, P.; Petrecca, M.;
(16) Williams, D. B.; Carter, C. B. Transmission Electron Microscopy: Estradé, S.; Peiró, F.; Puntes, V.; Nogués, J. Galvanic Replacement
A Textbook for Materials Science; Springer: New York, 2009; Vols. V1− onto Complex Metal-Oxide Nanoparticles: Impact of Water or Other
V4. Oxidizers in the Formation of Either Fully Dense Onion-like or
(17) Egerton, R. F. Electron Energy-Loss Spectroscopy in the Electron Multicomponent Hollow MnOx/FeOx Structures. Chem. Mater. 2016,
Microscope, 3rd ed.; Springer: New York, 2011. 28 (21), 8025−8031.
(18) Boyapati, S.; Wachsman, E. D.; Jiang, N. Effect of Oxygen (39) Wang, Z.; Bentley, J.; Evans, N. Valence State Mapping of
Sublattice Ordering on Interstitial Transport Mechanism and Cobalt and Manganese Using near-Edge Fine Structures. Micron 2000,
Conductivity Activation Energies in Phase-Stabilized Cubic Bismuth 31 (4), 355−362.
Oxides. Solid State Ionics 2001, 140, 149−160. (40) Torruella, P.; Arenal, R.; de la Peña, F.; Saghi, Z.; Yedra, L.;
(19) Shuk, P.; Wiemhöfer, H. D.; Göpel, W. Electronic Properties of Eljarrat, A.; López-Conesa, L.; Estrader, M.; López-Ortega, A.; Salazar-
Bi2O3 Based Solid Electrolytes. Z. Anorg. Allg. Chem. 1997, 623, 892− Alvarez, G.; et al. 3D Visualization of the Iron Oxidation State in FeO/
896. Fe3O4 Core−Shell Nanocubes from Electron Energy Loss Tomog-
(20) Yedra, L.; Xuriguera, E.; Estrader, M.; López-ortega, A.; Baró, raphy. Nano Lett. 2016, 16, 5068−5073.
M. D.; Nogués, J.; Roldan, M.; Varela, M.; Estradé, S.; Peiró, F. Oxide (41) Schattschneider, P.; Jouffrey, B.; Hebert, C. Orientation
Wizard: An EELS Application to Characterize the White Lines of Dependence of Ionization Edges in EELS. Ultramicroscopy 2001, 86,
Transition Metal Edges. Microsc. Microanal. 2014, 20, 698−705. 343−353.
(21) Varela, M.; Oxley, M. P.; Luo, W.; Tao, J.; Watanabe, M.; (42) Jones, E.; Oliphant, T.; Peterson, P.; et al. {SciPy}: Open Source
Lupini, a. R.; Pantelides, S. T.; Pennycook, S. J. Atomic-Resolution Scientific Tools for {Python}, 2001.
Imaging of Oxidation States in Manganites. Phys. Rev. B: Condens. (43) Schmidt, S.; Kubaski, E. T.; Volanti, D. P.; Sequinel, T.; Bezzon,
Matter Mater. Phys. 2009, 79 (8), 1−14. V. D. N.; Beltrán, A.; Tebcherani, S. M.; Varela, J. A. Effect of
(22) Vila, M.; Diaz-Guerra, C.; Piqueras, J.; Lopez-Conesa, L.; Pressure-Assisted Heat Treatment on Photoluminescence Emission of
Estrade, S.; Peiro, F. Growth, Structure, Luminescence and Mechanical A-Bi2O3 Needles. Inorg. Chem. 2015, 54 (21), 10184−10191.
(44) Wu, Y.; Lu, G. The Roles of Density-Tunable Surface Oxygen
Resonance of Bi2O3 Nano- and Microwires. CrystEngComm 2015, 17
Vacancy over Bouquet-like Bi2O3 in Enhancing Photocatalytic Activity.
(1), 132−139.
Phys. Chem. Chem. Phys. 2014, 16 (9), 4165−4175.
(23) Vila, M.; Díaz-Guerra, C.; Piqueras, J. Luminescence and Raman
Study of A-Bi2O3 Ceramics. Mater. Chem. Phys. 2012, 133 (1), 559−
564.
(24) Hébert, C. Practical Aspects of Running the WIEN2k Code for
Electron Spectroscopy. Micron 2007, 38, 12−28.

24815 DOI: 10.1021/acs.jpcc.7b06310


J. Phys. Chem. C 2017, 121, 24809−24815

You might also like