You are on page 1of 37

Progress in Materials Science 80 (2016) 1–37

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/pmatsci

Niobium oxides and niobates physical properties:


Review and prospects
C. Nico ⇑, T. Monteiro, M.P.F. Graça
Department of Physics & I3N, University of Aveiro, Campus Universitário de Santiago, 3810-193 Aveiro, Portugal

a r t i c l e i n f o

Article history: Abstract: For the last 75 years several studies have been reporting
Available online 16 February 2016 on the physical properties of niobium oxides, but there is still
many contradictory, inconsistent and insufficient information on
Keywords: these metal oxides. This review will begin by describing the nio-
Niobium oxide
bium oxygen system and the different stoichiometric and non-
Niobate
stoichiometric phases, specifically Nb, NbO, NbO2, Nb2O5 and
Electron correlated material
Polymorph
Nb2O5d. The crystalline phases and polymorphs of these materials
Transparent conductive oxide are often inconsistently identified in different works and thus, a
Resistive switching clarification of the nomenclature of the several niobium oxides
Oxygen vacancies polymorph and their crystalline structure is also presented. Due
to their interesting physical properties, many applications of these
materials have been suggested such as solid electrolytic capacitors,
catalysis, photochromic devices, transparent conductive oxides or
memristors, becoming obvious that a good understanding of nio-
bium oxides physical properties and their control is essential and
urgent. Additionally, a short review on different types of niobates,
namely alkali niobates, columbite niobates and rare earth niobates
and the relation of the properties of these materials with niobium
oxides will be presented.
Ó 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

⇑ Corresponding author. Tel.: +351 936876475.


E-mail address: claudionico@ua.pt (C. Nico).

http://dx.doi.org/10.1016/j.pmatsci.2016.02.001
0079-6425/Ó 2016 Elsevier Ltd. All rights reserved.
2 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

2. Niobium–oxygen system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1. Nb. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2. NbO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3. NbO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4. Nb2O5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.5. Non-stoichiometric niobium oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3. Niobates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.1. Alkali niobates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2. Columbite niobates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3. Rare-earths orthoniobates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4. Summary and prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

1. Introduction

Niobium oxides can lead to many different and interesting properties, making it a very versatile
group of materials. Specifically, niobium oxides have been showing great potentiality in many techno-
logical applications such as solid electrolytic capacitors, transparent conductive oxides, photochromic
devices, memristors, dye-sensitized solar cells and others.
Despite the known potentialities of niobium oxides, and many types of niobates, in several techno-
logical applications, the understanding of these oxide systems is still noticeably insufficient. The avail-
able literature reveals that niobium oxides are a complex system, with many phases and polymorphs,
and many works reporting contradictory or inconsistent information. More specifically, stoichiometry
problems are one of the most common complications in niobium oxides and in many different nio-
bates, where NbO6 octahedra are typically the basic structural units.
Hence, a careful bibliographic review is shown, which evidences the complexity of these materials,
the difficulty in identifying their different phases and polymorphs, as well as in the interpretation of
their properties.
This work intends to be an important and solid groundwork for future research on niobium oxides
and niobates systems, to enable a faster and more sustained scientific research and to evince the wide
range of properties and applications of these materials.

2. Niobium–oxygen system

In niobium–oxygen system, the niobium element can be found in four different charge states: 0, 2+,
4+ and 5+. Generally, these charge states are related to the phases of metallic Nb and to the NbO, NbO2
and Nb2O5 respectively, as Fig. 1 synthetically illustrates.
The special complexity of the niobium oxides system is related to the existence of several stoichio-
metric and non-stoichiometric phases, some of which have several polymorphs, some of which are

Fig. 1. Schematic illustration of the different oxidation states of niobium.


C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 3

Fig. 2. Niobium–oxygen phase diagram [19].

metastable, and the difficulty in synthesizing a single phase, i.e. without mixture of phases or stoi-
chiometry. Controlling, identifying and determining small variations of the stoichiometry is rather dif-
ficult in niobium oxides as the structure of these phases may be extremely similar, and also because
the precise quantification of oxygen is, technically, a great challenge. This is probably the major draw-
back of this system which can present a wide range of interesting physical properties, which are highly
dependent on the phase, polymorph and stoichiometry.
Although the first studies in niobium oxides are from Brauer in 1940 [1], and from other authors
later in the decade of 1960 [2–11], there still has been no consensus in literature concerning the phys-
ical properties and the nomenclature of the different niobium oxides phases. The fact that many of the
first scientific works on niobium oxides are written in German may partially justify the difficulty of the
international scientific community having access to very important information regarding these mate-
rials and thus contributing for a better understanding of these materials.
David Bach makes an ample collection and review of both early and recent literature of the niobium
oxides system in his PhD thesis (2009), defended at the University of Karlsruhe [12]. However, it is
stressed that many doubts about the synthesis, structure and physical properties of niobium oxides
still remain [12]. Bach et al. [12–17] used electron energy loss spectroscopy (EELS) as a high spatial
resolution technique for quantitative analysis of the niobium oxygen ratio and their oxidation state,

Fig. 3. Cubic crystalline structure of metallic Nb [21].


4 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

which is an extremely useful type of information, hardly obtained by any other technique, especially
because the properties of niobium oxides highly depend on their stoichiometry.
One of the first works on niobium oxides, by Elliot in 1960 [18], reports the phase diagram of the
niobium–oxygen system. Since then, other phase diagrams have been reported, and one of the most
recent, from 1990 by Massalski [19], is illustrated in Fig. 2. Massalski shows the existence of four ther-
modynamically stable phases of niobium–oxygen system (Nb, NbO, NbO2 and Nb2O5) with very nar-
row single-phase fields and negligible deviations from the exact stoichiometry [12,19]. However, one
should consider that this phase diagram is incomplete, as it fails to describe the formation of stable
non-stoichiometric phases at room temperature and the formation of different polymorphs, as it will
be further mentioned.

2.1. Nb

Niobium is a metal which crystallizes in a body-centred cubic (bcc) lattice (O9h spatial group) with a
density of 8.57 g/cm3. It is a refractory metal, good thermal conductor, with a melting and boiling
points at 2477 °C and 4744 °C, respectively [20]. It has an electrical resistivity of c.a. 15.2 lX cm at
273 K and it is a superconductor below the critical temperature Tc ’ 9.3 K [20] (see Fig. 3).
Niobium can form very stable carbides, nitrides, borides and silicides, possessing high interatomic
bonding energies [22]. Hence, niobium is used in the stabilization of stainless steels to prevent inter-
granular corrosion since it helps lowering the content of carbon in the steel. A small niobium amount,
from 100 to 5000 ppm, is typically necessary for the prevention and retardation of austenitic grain
growth which leads to higher strength and toughness [23]. These high-strength microalloyed steels
are commonly used in pipelines for oil and gas transportation, in high rise buildings construction,
in tools steels and heat resisting cast steel auto parts [23–25]. Niobium is also commonly added to
most nickel-based superalloys for application in aircraft engines, land-based turbines for power gen-
eration, chemical processing industries and other applications where a special resistance to abrasion
and corrosion is required [24,25].
Pure niobium and niobium-based alloys have a high resistance to corrosion in several environ-
ments, which is attributed to a passivating niobium oxide film at the surface [24–26]. These materials
also present high melting points and refractory properties, particularly the niobium-based alloys
which can surpass the maximum application temperatures of nickel-based superalloys, thus being
well suited for space, nuclear and aircraft applications [24,25]. Furthermore, niobium-based alloys
are typically characterized by low densities and relatively high ductility at room temperature which
is an advantage for cold-working and fabrication of complex structures [23,24]. However, these alloys
typically require a protective coating, such as a silicide, to prevent them from easily oxidizing at high
temperatures [24]. Other applications of niobium comprise the fabrication of metal insulator metal
(MIM) tunnel diodes and hard tissue replacements (given its good biocompatibility) [27,28].

Fig. 4. Cubic structure of NbO [21].


C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 5

Niobium also presents type II superconductivity properties. In fact, from all the chemical elements,
metallic niobium is the one with the highest critical temperature (Tc ’ 9.3 K) [20,26]. Other niobium-
based compounds, such as Nb3Ga or Nb0.75(GeAl)0.25, are among the materials with the highest critical
temperatures for superconductivity (Tc = 20.0 K and Tc = 20.7 K respectively) [29]. Hence, niobium is
frequently used in the fabrication of radio-frequency superconducting cavities (for particle accelera-
tors), production of Josephson tunnel junctions, single photon detectors or superconducting quantum
interference devices (SQUIDs) used, for instance, to cover the surface of nearly perfect spheres for the
Gravity Probe B mission, which purpose was to measure the space–time curvature near the Earth
[30,31]. However, pure metallic niobium is not suitable for superconducting magnets where high
upper critical magnetic fields (Bc2) are required. Still, while niobium has a Bc2 ’ 420 mT (at 0 K), an
alloy of NbTi can have a Bc2 ’ 11 T (at 4.2 K), and the Nb3Sn compound a Bc2 ’ 25 T (at 4.2 K) [32].
Therefore, it is no surprise that these niobium-based materials are fundamental and used, almost
exclusively, in the production of high magnetic fields, required in a series of industrial and scientific
applications. One flagrant example is the LHC project, in CERN, that required hundreds of tons of Nb
and NbTi alloy.
Niobium is also characterized by its high affinity and binding energy to oxygen [26]. Solubility of
oxygen in the niobium matrix, where it occupies octahedral interstitial sites, increases with temper-
ature (ranging from 0.8% at 500 °C to 9.0 at.% at 1915 °C) [19]. Oxygen in solid solution leads to an
increase of the lattice parameter of metallic niobium, acts as an hardener and decreases the ductility
of the metal [33]. Oxygen does also affect the electrical properties increasing the resistivity of niobium
c.a. 4.1 lX cm per at.% of oxygen, and decreasing the superconductivity transition temperature
Tc  0.93 K per at.% of oxygen [12,26]. Therefore, and also because defects at the surface may produce
heat, high-purity niobium is required for superconductivity applications. It is commonly mentioned in
literature the existence of a native oxide layer at the surface of niobium exposed to oxygen [26,34]. In
a superconductor, the electromagnetic fields have a penetration depth of c.a. 60 nm, and therefore the
existence of an oxide layer at the surface is critical. To obtain highly pure Nb surfaces, i.e. without an
oxide layer, the metal is typically heated above 2000 K in ultra-high vacuum conditions [26]. The com-
position and thickness of such native oxide layer is not yet well clarified, but it is accepted that it
depends on the processing parameters and techniques of the metal, crystallographic orientation,
and ambient conditions such as oxygen partial pressure, temperature and exposure time [26,29,35–
40]. Some authors [26,35,38] suggest that this layer is around 5–6 nm thickness with a composition
of nearly stoichiometric Nb2O5, while other authors [12,41] report thicknesses as high as 25 nm
and/or with different layers of niobium suboxides [26,29,35,42,43]. More recently (2008), Delheusy
et al. [43] suggested a three-layer model of NbO, NbO2 and Nb2O5 between the metallic niobium
and the oxide surface.

Fig. 5. Structure of (a) tetragonal and (b) rutile NbO2 [21].


6 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

2.2. NbO

NbO crystallizes in a face-centred cubic structure (similar to NaCl), O1h space group, with 3 Nb
atoms in the lattice sites 3(c) (0 ½ ½; ½ 0 ½; ½ ½ 0), and 3 O atoms in the lattice sites 3(d) (½ 0 0;
0 ½ 0; 0 0 ½;) [7], where each Nb atom is coordinated to four O atoms in a square planar array, as illus-
trated in Fig. 4. Moreover, NbO crystalline structure is a very particular case since it shows 25%
ordered vacancies in both sublattices of Nb and O [44]. This is the highest number of point defects
among all transition metal monoxides.
The density of NbO is c.a. 7.3 g/cm3 and has a melting point at Tm ’ 1940 °C [12,18]. Niobium
monoxide presents a typical metallic behaviour, and is widely regarded as a metal [11,44,45], with
a resistivity of about 21 lX cm at 25 °C [11,45] that decreases with temperature down to 1.8 lX cm
at 4.2 K [45]. Kurmaev et al. [44] performed X-ray fluorescence measurements for different niobium
oxides and compared the results of NbO to the band structure calculations of Nb1.0O1.0, finding that
there were significant differences. Therefore, Kurmaev et al. tried to mimic the NbO structure by per-
forming calculations of the band structure for Nb0.75O0.75 in order to make account for the 25% of
vacancies. Given the good correspondence between these calculations and the experimental results,
the authors assigned the electrical properties of NbO to the aforementioned vacancies, characteristic
of this structure: ‘‘The electronic DOS of Nb0.75O0.75 is characterized by the formation of additional sub-
bands connected with vacancy states.” [44].
Below the critical temperature, Tc ’ 1.38 K, NbO becomes superconductive [45]. Hulm et al. [45]
performed electrical measurements and studied the superconductivity of different transition metal
monoxides, and found that, in NbO, there is no evidence of negative temperature-coefficient behaviour
of the type that was observed in TiO and VO. Furthermore, the authors verified that a small increase of
the oxygen ratio (towards NbO2) would induce a sharp increase of the resistivity. Oppositely, an
increase of the niobium ratio (towards metallic Nb) would promote an increase of the critical temper-
ature for superconductivity [45]. The authors attribute this behaviour to a mixed-phase sample with
contributions of both NbO and NbO2 (when the O% is increased), or of both Nb and NbO (when the Nb%
is increased).
Niobium monoxide is not used massively in any major technological application. However, the fact
that NbO has improved properties, regarding the oxygen diffusion, in comparison with Nb, makes it a
suitable candidate to niobium-based solid electrolytic capacitors [46–50].

2.3. NbO2

The NbO2 has the Nb element in a 4+ charge state, presents a stoichiometry between NbO and
Nb2O5, and has a melting point of 1901 °C [20]. NbO2 is typically characterized by a blue colour
(associated to the Nb4+ ions [8]) and can be obtained by controlled oxidation of Nb or NbO, or reduc-
tion of Nb2O5 [51].
At room temperature NbO2 crystallizes in what can be described as a distorted tetragonal super-
structure with a rutile-type sublattice, with a space group C64h, and a density c.a. 5.9 g/cm3 [52,53]. This
structure, illustrated in Fig. 5a is characterized by chains of edge-sharing NbO6 octahedrons which are
cross-linked by sharing corners [52] where the Nb–O distances in the octahedral are practically the
same, c.a. 2.05 Å [52]. Similarly to the VO2 phase at room temperature, in NbO2 the metal ions tend
to pair along the fourfold [0 0 1] axis, along the edge-sharing chains, with alternated Nb–Nb distances
of 2.80 and 3.20 Å [52,53]. Inherent instabilities of the NbO6 octahedra and the structure of NbO2 have
been reported [53,54].
It is also known that between 797 and 808 °C, the NbO2 suffers a reversible second-order phase
transition, together with a change of the crystal structure into a regular rutile lattice, illustrated in
Fig. 5b [52,53,55,56]. Electrically, this is characterized as being a semiconductor–metal transition,
where this high temperature NbO2 phase shows a typical metallic conductivity (c.a. 103 S/cm) [56].
It was also reported that above 1100 °C a small, but measurable, change of the stoichiometry towards
NbO2.006 occurs, which can strongly influence the electrical conductivity of the room temperature
phase [56,57].
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 7

Several metal dioxides, such as VO2, TiO2, GeO2 and SnO2, have a rutile-type structure, but their
physical properties can be significantly different. In fact, GeO2 with a reported band gap between
5.4 and 5.9 eV [58] is considered an insulator, while SnO2 can be considered as a wide band gap semi-
conductor (3.6 eV) [59] as also the rutile TiO2 (3.05 eV) [60,61]. TiO2, GeO2 and SnO2 crystallize in a
rutile structure at room temperature and, despite their different band gap energies, they all exhibit a
fundamental indirect absorption edge with dipole forbidden direct band gap [61,62]. However, NbO2
is more frequently compared to VO2, not only due to their structural similarities, but particularly
because they both show a semiconductor–metal transition. VO2 crystallizes in a monoclinic structure
at room temperature with band gap between 0.6 and 0.7 eV, and undergoes a sharp transition to the
rutile phase (metallic behaviour) near room temperature (340 K) at ultrafast timescales [63]. This
makes VO2 an extremely promising material for a broad range of electronic applications (e.g. sensor,
switching and memory elements) [63]. However, as aforementioned, NbO2 undergoes such metal–in-
sulator transition at a much higher temperature (c.a. 1081 K), which can be an advantage to isolate the
electric properties from temperature effects. In VO2 this type of behaviour is typically associated to a
Mott–Peierls transition [63–66], a combination of the effect of the structural change in electronic
structure, and strong electronic correlation commonly observed in oxides with 3d transition metals.
However, Nb is a 4d transition metal and therefore it would be expectable that electron correlation
would be less significant [64–66]. In fact, O’Hara and Demkov have recently shown that the metal–in-
sulator transition in NbO2 is explained by a second-order structural (Peierls) transition via dimeriza-
tion of chains of Nb atoms, without the need to account for electron correlation effects, unlike VO2
[64]. Wong et al. were able to show that in NbO2 the djj  epg orbital splitting is 0.3 eV larger than
the observed in VO2, which may explain its higher metal–insulator transition temperature [66].
Some authors [67,68] have mentioned the existence of a monoclinic crystalline structure of NbO2,
citing the work of Terao [69] as the reference for such phase. In addition, the Joint Committee on Pow-
der Diffraction Standards – International Centre for Diffraction Data (JCPDS–ICDD) card No. 19-0859
does in fact report the X-ray powder diffraction (XRD) data of a monoclinic phase of NbO2, with lattice
parameters a = 12.03 Å, b = 14.37 Å, c = 10.36 Å and b = 121.17°, also referring to the work of Terao
[69]. However, in the abstract of the Terao’s work, it was only mentioned a tetragonal phase of
NbO2, among other phases of niobium oxides. It is therefore reasonable to doubt about the true exis-
tence of such monoclinic structure, since the work of Terao is the only work reporting, allegedly, a fun-
damental structural study on an NbO2 monoclinic phase.
The physical properties of NbO2, namely the optical and electrical ones, have not been yet exten-
sively explored. Still, the tetragonal phase of NbO2 is usually classified as being an n-type semiconduc-
tor with a small band gap (between 0.5 and 1.2 eV) [65] and an electrical resistivity in the order of
104 X cm [44,53,55,56]. Weibin et al. [70] theoretically calculated the electronic structure of a tetrag-
onal phase of NbO2, with an indirect band gap of 0.25 eV, but experimental results on thin films have
shown a band gap of 0.5 eV, and a vacuum level (ionization energy) of 5 eV [70]. Recently, in an
attempt to put an end to the inconsistent reports on NbO2 electronic structure, and in line with other
rutile based oxide structures, theoretical and experimental studies reported an indirect band gap of at
least 1.0 eV [65,71]. Also very recently, Wang et al. were able to show that the electrical conduction
mechanism in tetragonal NbO2 is described by the Efros–Shklovskii variable range hopping model,
rather than Mott’s model which is associated to VO2 [72].
In 2004, Zhao et al. [55] reported the optical and dielectric properties of nanostructured NbO2 (in
the form of nanoslices), synthesized by thermal deposition, using a metallic niobium wire deposited in
low vacuum as a thin film on a Si substrate. In this work, the authors report for the first time the
Raman spectrum of NbO2, showing at least 12 active modes with energies of 139, 170, 247, 333,
343, 405, 436, 463, 580, 631, 699 and 815 cm1 [55]. Furthermore, the photoluminescence (PL) anal-
ysis of the NbO2 nanoslices, under 512 nm laser excitation and at room temperature, revealed two
emission bands centred at 700 nm (1.77 eV) and 825 nm (1.50 eV). The origin and corresponding
mechanisms for the observed PL were not clarified, but it was suggested that, in analogy with other
metal dioxides, they may be attribute to intraionic Nb4+ dxy–dyz band transitions [55]. The dielectric
characterization performed by Zhao et al., between 1 kHz and 10 MHz, revealed interesting properties
of the NbO2 nanostructures. It was observed that the relative dielectric constant of NbO2 remains
8 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

almost constant, c.a. 10, in the measured frequency range, and with low dielectric losses (<1%) for the
same frequency range [55]. This type of dielectric characteristics on nanostructures, allied to the sim-
plicity of fabrication, may be quite attractive for the production of MOS devices, as suggested by the
authors [55]. In the last couple years, many works have been exploring NbO2 due to its resistive
switching properties for the development of memristive devices [73–78]. In fact, Slesazeck et al.
[78] have recently proposed a model that describes the behaviour of NbO2 based filamentary switch-
ing devices, by a temperature activated Frenkel–Poole conduction mechanism, showing that the such
switching effect is independent on the metal insulator transition.
NbO2 has also been studied regarding other technological applications, such as impedimetric
biosensors using biofunctionalized Nb/NbO2 electrodes [79,80]. Moreover, Sasaki et al. [81] took
advantage of the good chemical stability of NbO2 to produce highly efficient electrocatalysts of
Pt/NbO2/C nanostructures for oxygen-reduction reactions for application in proton exchange mem-
brane fuel cells. In 2012 Lee et al. [80] reported the synthesis of NbO2 nanowires by chemical vapour
transport method which can further enhance the potentiality of this material for such technological
applications.

2.4. Nb2O5

Niobium pentoxide (Nb2O5) is the most thermodynamically stable state of the niobium–oxygen
system. With an charge state of 5+ in Nb2O5, the electronic structure of the Nb atom is [Kr]4d0, which
means that all the 4d electrons are bonded to the O 2p-band, thus justifying the fact that Nb2O5 has a
much lower electrical conductivity than the other niobium oxides [50]. Nb2O5 can occur in the amor-
phous state or in one of many different crystalline polymorphs. Generally, all the Nb2O5 polymorphs
have a white colour (in the form of powders) or transparent (in single-crystals). However, most of the
physical properties of Nb2O5 depend on its polymorph and on the used synthesis parameters and tech-
nique [8,49,50,82–85].
The complexity, confusion and contradictions regarding the niobium oxides system are old and still
prevail up to the present. In 1966, Schäfer et al. published a work entitled ‘‘The modifications of Nio-
bium Pentoxide” [8], where it was stated that ‘‘The abundance of observations makes it desirable to sub-
ject the available material [niobium pentoxide] to a critical review and classification, and thus
simultaneously to create a basis for further investigations”. Already then, there were many different
phases of niobium oxides with different nomenclatures and there was, therefore, the need of clarify-
ing, organizing and compiling all the existent information. Some of the Nb2O5 polymorphs were clas-
sified with a sequence of Greek letters as it is common in well-known systems. However, since there
was not yet a well-established knowledge of all polymorphs, Schäfer et al. decided to use neutral sym-
bols following, and extending, the same type of classification as Brauer’s [1]. Thus, some polymorphs
were classified based on the temperature they were obtained: TT, T, M and H (from the German Tief–
Tief, Tief, Medium and Hoch, meaning low–low, low, medium and high), while other polymorphs were
named after the shape of the particles B, N and R (from the German Blätter, Nadeln and Prismen,
meaning leaves/plates, needles and prisms). While Schäfer et al. [8] were able to match some of the
same polymorphs reported by different authors, there were other niobium pentoxide modifications,
such as the e, I-high and II, that they were not able to reproduce and study, suggesting the possibility
of these phases to be metastable and/or non-stoichiometric.
From what it is possible to assess from current literature, the opinion shared in this work is that not
all Nb2O5 polymorphs are yet well known and studied, what justifies the choice of following the
nomenclature suggested and reported by Schäfer et al. [8], also followed by some authors. There
are however other authors that misuse this nomenclature by choosing the letter according to the crys-
talline structure, e.g. T-, H-, O- and M-Nb2O5, respectively standing for Tetragonal, Hexagonal,
Orthorhombic and Monoclinic structures of niobium pentoxide [86,87], which can be confusing and
misleading.
The Nb2O5 phase can exist in an amorphous state, but it may crystallize in several kinds of poly-
morphs with different physical properties: T (D92h, orthorhombic), B (C62h, monoclinic), H (C12h, mono-
clinic), N (C32h, monoclinic), Z (C12, monoclinic), R (C32h, monoclinic), M (D17 10
4h, tetragonal), P (D4 ,
tetragonal) and also TT (pseudohexagonal or monoclinic) [8,12,83,88]. The TT-Nb2O5 polymorph can
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 9

be interpreted as a less developed phase of the T-polymorph, which is only stabilized by impurities,
such as OH or Cl, or vacancies [13,83,89]. Among all the Nb2O5 polymorphs, and niobium oxides
in general, the H-phase is thermodynamically the most stable. The H-Nb2O5 is therefore one of the
most common, and probably the most studied, niobium pentoxide polymorph. A lot more of meta-
stable Nb2O5 modifications were reported, but not extensively studied [8,90–92].
Generally, the structures of niobium pentoxides are based on NbO6 octahedra which may be more
or less distorted depending on the type of linkage between the octahedra. Most of Nb2O5 polymorphs
structures are described by the combination of one or both of these types of links between octahedra
that can happen by corner-sharing or by edge-sharing. Particularly, the structures characterized by
perpendicular edge-sharing octahedra are arranged in such a way that a zigzag chain is produced. This
type of chains, linked in parallel by corner-sharing, are very common as an additional constructional
element in many forms of niobium pentoxide but also in related compounds [8]. Actually, the fact that
there are several possible combinations of octahedral linkages that can produce an O/Nb ratio of 2.5, is
pointed as the reason for the multiplicity of Nb2O5 structures [8]. Furthermore, the principle of crys-
tallographic shear formed by the existence of different types of linkage regions (but preserving the
cation coordination), but also the formation of point defects, can explain the possibility of variations
of stoichiometry in relation to Nb2O5, giving rise to the formation of non-stoichiometric niobium oxide
phases [8,93–97]. In a very recent paper, Valencia-Balvín et al. [83] report first-principle studies on the
energetics and phase stability of the several Nb2O5 polymorphs under pressure, and also perform a
comprehensive and detailed description of their crystalline structures.
Additionally, the structural characteristics of the different niobium oxides can vary significantly
with the synthesis technique [8,34,82,83,85]. This is clearly evidenced by Reznichenko et al. [82]
who make a very complete list of numerous works in the literature with different production methods
of several stable and metastable Nb2O5 polymorphs, their structural parameters and different desig-
nations or nomenclatures. The most reported techniques used to prepare different Nb2O5 polymorphs
involve the oxidation of niobium oxides of lower stoichiometry, promoted by heating in air, or by heat
treating other Nb2O5 phases. Other methods, such as heating sulphate or chloride niobic acid, anodiza-
tion, chemical transport, sol–gel, special hydrothermal conditions, or high temperatures and/or pres-
sure, are also commonly used in the production of different Nb2O5 polymorphs [8,49,50,82,84,85,89].
Generally, the temperature and starting material used in the synthesis method will be the most deter-
minant parameters.
One interesting behaviour that is observed in the synthesis of the niobium oxides, is the ‘‘memory of
solids”, as it was referred by Schäfer et al. [8]. This behaviour is characterized by how the same Nb2O5
phase will behave differently under the same heat treatment, depending on the original preparation

Fig. 6. Monoclinic structure of the H-Nb2O5 phase [21].


10 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

Fig. 7. Monoclinic structure of the B-Nb2O5 phase [21].

Fig. 8. Monoclinic structure of the N-Nb2O5 phase [21].

method, i.e. similar crystalline samples undistinguishable by XRD measurements will ‘‘remember”
their synthesis method by changing to different phases under the same heat treatment [8]. While this
behaviour is not yet well clarified, it was suggested that it might be related to the incorporation of dif-
ferent impurities or specific structural defects, that are dependent on the original synthesis route [8].
The H-Nb2O5 phase is rather easy to be obtained. Starting from any Nb2O5 polymorph, or a lower
stoichiometric oxide (such as NbO2, NbO or even the metallic Nb), the H-Nb2O5 is obtained by a heat
treatment in air at high temperatures (>1000 °C) [8,12]. Heating any type of niobic acid precipitates
(sulphate, chloride, bromide, iodide or fluoride) at high temperatures will also produce this phase
[8]. The production of the H phase can also be achieved under hydrothermal conditions and from a
melt (if seeded with the same phase) [8]. The monoclinic structure of the H-Nb2O5, which contains
14 formula units per unit cell, is illustrated by Fig. 6. The Nb atoms occupy the Wyckoff positions
1g, 2i (with half occupation), six 2n and seven 2m, while the O atoms occupy the 1f, 1h, sixteen 2n
and eighteen 2m positions. This structure is described by layered 3  4 and 3  5 blocks of corner-
sharing octahedra. These blocks are connected by edge-sharing octahedra, with a partly but system-
atic tetrahedral site, occupied by one Nb atom per unit cell [83].
The B-Nb2O5 phase can be produced by chemical transport of Nb2O5 or NbOCl3 at temperatures
between 750 and 850 °C. However, it was observed that using this method other polymorphs are pro-
duced in addition, e.g. N and P-Nb2O5 [8]. It is possible to obtain a single B phase sample from molten
Nb2O5 using a selected crystalline seed of the same phase [8]. Alternatively, it was reported that by
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 11

Fig. 9. Tetragonal structure of the M-Nb2O5 phase [21].

Fig. 10. Tetragonal structure of the P-Nb2O5 phase [21].

heating, in air, between 500 and 800 °C metallic Nb or a lower stoichiometry niobium oxide (NbO or
NbO2 for instance), as well as the TT or the T-Nb2O5 polymorph, is a good method for producing the B
phase [8]. The monoclinic structure of this polymorph, described by edge-sharing distorted NbO6 octa-
hedra blocks (each block linked in zigzag by corner-sharing) is illustrated by Fig. 7. This structure con-
tains 4 formula units per unit cell, with Nb atoms located at the Wyckoff position 8f and the O atoms at
4e and two 8f positions [83].
The N-Nb2O5 phase can be found, as mentioned, together with the M phase when synthesized from
niobic acids. It can also be prepared by chemical transport of Nb2O5 at 840 °C, but only in the pres-
ence of small amounts of fluoride, producing additional phases like the H and B polymorphs [8]. The N
polymorph can also be produced under hydrothermal conditions where the OH groups may play the
same role as the F [8]. It is also possible to prepare this polymorph by thermal decomposition of
NbO2F, at c.a. 1000 °C in high vacuum, though it is a very sensible method given that the formation
of Nb3O7F or H-Nb2O5 can easily occur if the time and temperature are not optimal [8]. The monoclinic
12 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

Fig. 11. Monoclinic structure of the R-Nb2O5 phase [21].

Fig. 12. Monoclinic structure of the Z-Nb2O5 phase [21].

Fig. 13. Orthorhombic structure of the T-Nb2O5 phase [21].

structure of the N-Nb2O5, illustrated by Fig. 8, is described by 4  4 blocks of corner-sharing octahedra,


where the blocks are interlinked by edge-sharing. The Nb and O atoms occupy, respectively, eight and
twenty 4i Wyckoff positions [83].
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 13

The M-Nb2O5 phase is usually formed by heating sulphate or chloride niobic acid between 900 and
950 °C for some hours (or even at higher temperatures but for shorter times). This phase can also be
detected on oxidized Nb and Nb alloys when heat treated at 1000 °C in air [8]. Nevertheless, these
methods also produce additional polymorphs, especially the N phase, hampering the preparation of
a single phase sample of M-Nb2O5 [8]. Very recently, Dhawan et al. [98] were able to prepare spherical
quantum dots of M-Nb2O5, by physical vapour deposition, with controllable sizes between 1 and
20 nm. For quantum dots with diameters below 5 nm, quantum confinement effects were observed
[98]. The tetragonal structure of the M phase is illustrated in Fig. 9. The unit cell contains 16 formula
units, where the Nb atoms occupy two 8i, two 8h and three 16l Wyckoff positions, and the O atoms at
two 8j, two 8h and three 16l. The structure can be described by 4  4 blocks of corner-sharing octa-
hedra, with adjacent blocks linked by octahedra edges [83].
The P-Nb2O5 phase can be obtained by chemical transport, either in the chloride, bromide or iodide
systems, at a temperature c.a. 750 °C [8]. It was found that the presence of small amounts of water was
fundamental to promote the growth of the P phase [8]. Alternatively, the P-Nb2O5 can be formed by
very slow thermal decomposition of NbO2F [8]. The idealized structure of the P polymorph, illustrated
in Fig. 10 (courtesy of Professor Osorio-Guillén from the Physics Institute of the University of Antio-
quia, Colombia), contains 4 formula units per unit cell where the Nb atoms occupy the Wyckoff posi-
tion 8c while the O atoms occupy a 4i and two 8c positions [83]. The structure is composed by
distorted octahedra organized in blocks of two edges-sharing, where the blocks are linked by
corner-sharing [83].
The R-Nb2O5 phase was first reported by Gruehn in 1966 [90]. It is not easy to find works which
have identified this phase, though this polymorph is commonly referred in literature [83,99,100]. Orig-
inally, Gruehn obtained the R phase by chemical transport of Nb2O5 at temperatures between 600 and
800 °C. Using the same method with chloride niobic acid also produced this phase. Additionally,
Gruehn said that the R-Nb2O5 was also found in the product of NbOCl3 hydrolysis heated at 275 °C
[90]. These methods, however, did not produce a single phase material, but constantly a mixture of
R-Nb2O5 with other polymorphs (e.g. P or TT) [90]. The R phase has one of most simple structures
among niobium pentoxides, illustrated in Fig. 11. This monoclinic structure has two formula units
per unit cell with the Nb atoms occupying the Wyckoff positions 4i and the O atoms the 2a and 4i
[83]. This structure is described by distorted octahedra linked by edge-sharing, forming zigzag chains
along the b direction which are interlinked by corner-sharing [90].
The Z-Nb2O5 phase was first reported by Zibrov et al. in 1998 [101]. This modification was identi-
fied, together with the B polymorph, after heat treating the H-Nb2O5 phase, between 800 and 1100 °C,
in a high-pressure chamber at 8.0 GPa during 1–10 min [101]. The structure was determined and
refined by the Rietveld method from the XRD measurements, and is illustrated is Fig. 12. It was found
that this monoclinic structure is one of the few niobium pentoxides without sixfold coordinated nio-
bium, presenting mono-capped trigonal prisms with sevenfold coordination Nb atoms occupying the
Wyckoff position 4c, while the O atoms occupy a 2b two 4c positions [83,101].
The T-Nb2O5 phase is one of the most commonly studied niobium pentoxide phases and one of the
first to be reported (originally by Brauer in 1941 [1]). This phase can be obtained by heating sulphate
of chloride niobic acid between 600 and 800 °C. Heating the TT-Nb2O5 polymorph, a lower niobium
oxide (NbO2 or NbO), or metallic Nb, to the same range of temperatures (600–800 °C) does also pro-
duce the T phase [8]. Additionally, this polymorph can be obtained under hydrothermal conditions
(starting from amorphous niobic acid), by quenching Nb2O5 into supercooled melts, or even by chem-
ical transport of Nb2O5 [8]. The orthorhombic structure of the T phase, illustrated in Fig. 13, is not sim-
ple and contains 8.4 unit formulas per unit cell. Nb atoms with six and sevenfold coordination produce
distorted octahedra and pentagonal bipyramids, respectively. These polyhedra are linked both corner
and edge-sharing along the [0 0 1] direction but are only connected by corner-sharing along the c axis
[83]. There are 16 Nb atoms, distributed in parallel with the (0 0 1) plane and placed at four 8i Wyckoff
positions with half-occupancy, while 0.8 Nb atoms are distributed in three 4g positions with 0.08, 0.08
and 0.04 occupancies [83]. The occupancy of the Nb atoms is represented in Fig. 13 by the percentage
of blue that each atom is coloured. The O atoms occupy one 2b, four 4g and six 4h Wyckoff positions
[83].
14 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

Table 1
Niobium oxides crystalline phases.

Phase Crystal system Lattice parameters Space group References


NbO Cubic a = 4.210 Å O1h [7,13]
NbO2 Tetragonal a = 13.695 Å C64h [52,55]
c = 5.981 Å
Rutile a = 4.55 Å/a = 4.841 Å D14
4h [54,179]
c = 2.86 Å/c = 2.992 Å
Monoclinic a = 12.03 Å C2h [53,67,68]
b = 14.37 Å
c = 10.36 Å
b = 121.17°
Nb12O29 Monoclinic a = 15.686 Å C32h [173,175,178]
b = 3.831 Å
c = 20.71 Å
b = 121.17°
Orthorhombic a = 3.832 Å D17
2h [173,175,178]
b = 2.740 Å
c = 28.890 Å
Nb22O54 Monoclinic a = 15.749 Å C12h [173]
b = 3.824 Å
c = 17.852 Å
b = 102.03°
T-Nb2O5 Orthorhombic a = 6.175 Å/a = 6.144 Å D92h [99,100,102,180]
b = 29.175 Å/b = 29.194 Å
c = 3.930 Å/c = 3.940 Å
B-Nb2O5 Monoclinic a = 12.73 Å C62h [99,100,181,182]
b = 4.88 Å
c = 5.56 Å
b = 105.1°
H-Nb2O5 Monoclinic a = 21.153 Å/a = 21.163 Å C12h [99,102,180–183]
b = 3.8233 Å/b = 3.824 Å
c = 19.356 Å/c = 19.355 Å
b = 119.80°
3
N-Nb2O5 Monoclinic a = 28.51 Å C2h [99,100,181,182]
b = 3.830 Å
c = 17.48 Å
b = 120.8°
Z-Nb2O5 Monoclinic a = 5.219 Å C12 [101]
b = 4.699 Å
c = 5.928 Å
b = 108.56°
R-Nb2O5 Monoclinic a = 12.79 Å C32h [99,100]
b = 3.826 Å
c = 3.983 Å
b = 90.75°
M-Nb2O5 Tetragonal a = 20.44 Å D17
4h [99,100]
c = 3.832 Å
P-Nb2O5 Tetragonal a = 3.876 Å D410 [99,100]
c = 25.43 Å
TT-Nb2O5 Pseudohexagonal a = 3.607 Å/a = 3.600 Å – [86,99,100,102,180]
c = 3.925 Å/c = 3.919 Å
Monoclinic a = 7.23 Å
b = 15.7 Å
c = 7.18 Å
b = 119.08°

Finally, the TT-Nb2O5 phase is also frequently reported in literature, and is commonly referred as a
less crystalline form of the T phase, stabilized by vacancies or impurities, such as OH or Cl
[83,85,89]. It is reported that the TT-Nb2O5 can be obtained by heating sulphate or chloride niobic acid
to temperatures c.a. 500 °C, or by promoting the oxidation of lower niobium oxides by heating
between 320 and 350 °C in air. In both methods, amorphous Nb2O5 is observed as an intermediate.
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 15

This phase is also observed in oxidized Nb alloys at higher temperatures (800 °C), and as one of the
products of the reaction between NbO2 and Cl2 carried out between 270 and 320 °C. It is reported that
the TT-Nb2O5 can crystallize in either monoclinic or pseudo-hexagonal structure [83,85,99,100,102],
but the unit cell was never fully refined. It is affirmed, nevertheless, that the structure of the T phase
is characterized by the presence of distorted octahedra, pentagonal and hexagonal bipyramids, i.e.
NbO6, NbO7 and NbO8 polyhedra [83,102], which are reported to be the same structural units as
the amorphous phase of Nb2O5 [13,103].
A list of these Nb2O5 polymorphs (and other niobium oxides), with the respective crystal lattice
parameters and space groups, is presented in Table 1.
Regarding the physical properties of niobium pentoxides, one should expect that, depending on the
Nb2O5 polymorph, some of the properties may be different. In fact, not only the physical properties
depend on the polymorph, but also on the synthesis method [8,13,50,82,104]. Because of this, it is
common to find a wide range of values for some properties reported for Nb2O5. Nonetheless, indepen-
dently on the polymorph, Nb2O5 is generally considered a wide band gap semiconductor (or insulator,
depending on the classification criteria) [44,50].
Filho et al. performed photoelectrochemical studies and measured the band gap of thin films of
amorphous niobium oxides, which was consistently around 3.4 eV, in agreement with results of pre-
vious works [105]. The authors also measured a high quantum yield on certain niobium oxides films
and discussed the charge transfer model between this material and an electrolyte. Schultze and
Lohrengel [106] report niobium pentoxide passive films with band gap energies between 3.4 and
5.3 eV. Brayner and Bozon-Verduraz [89] report a nearly constant band gap of 3.4 eV for niobium pen-
toxide, in the amorphous, TT, T and H phases. However, the authors also report that, using a soft route
method they were able to produce nanoparticles, with sizes between 40 nm and 4.5 nm, which band
gap varied between 3.4 and 4.2 eV, respectively [89]. This variation was assigned to quantum confine-
ment effects. It was also suggested that the local coordination of the Nb atoms can be a determinant
factor in the band gap of niobium pentoxides and other niobates [89]. Additionally, it was reported
that Nb2O5 particles heat treated at low temperatures (around 400 °C) presented a brownish-white
colour, turning into white after being heated, in air, to temperatures higher than 600 °C (the charac-
teristic colour of niobium pentoxide), which was assigned to oxygen vacancies originated during the
synthesis process and respective vacancy filling with increasing temperature [89]. Soares et al. [50]
reported about the effect of the processing method on the optical and electrical properties of niobium
pentoxides. A band gap of c.a. 3.0 eV for Nb2O5 single crystals (H phase) at room temperature, increas-
ing 100 meV when cooled down to 14 K, was observed. Additionally, photoconductivity measure-
ments revealed identical band gap values for polycrystalline samples of H-Nb2O5, while for
polycrystalline T-Nb2O5 two photoconductivity peaks at 3.4 and 4.7 eV were respectively assigned
to the fundamental absorption edge and to a transition from the conduction band to a higher energy
band structure of this phase [50]. In 2011, Kovendhan et al. [107] reported a band gap of 3.75 eV in M-
Nb2O5 crystalline thin films. Very recently, Dhawan et al. [98] observed, through optical absorption
measurements, a variation of the band gap from 4.15 to 3.36 eV in M-Nb2O5 spherical quantum dots
between 1 and 20 nm in diameter, respectively. While such variation was assigned to quantum con-
finement effects, it was only detected in quantum dots with less than 5 nm in diameter. For bigger
sizes the band gap was nearly constant, therefore indicating that the exciton Bohr radius in the M-
Nb2O5 phase should be c.a. 5 nm. Additionally, a small effective mass of the exciton l ffi 0.532 was
reported [98]. Weibin et al. [70] calculated the electronic structure of B-Nb2O5 with density functional
theory, which results have indicated an indirect band gap of 2.55 eV. Additionally, the authors per-
formed X-Ray and Ultraviolet Photoelectron Spectroscopy (XPS and UPS) measurements on Nb2O5
films deposited by pulsed laser deposition on a Si(1 0 0) substrate, not specifying however the crys-
talline structure (or amorphous nature) of the samples, and the results pointed to a band gap of
3.7 eV and a vacuum level (ionization energy) of approximately 4.2 eV [70].
It is occasionally stated that Nb2O5 presents n-type conductivity [83,104], but in fact this type of
conductivity was reported, and it is today strongly accepted, as being associated to small stoichiom-
etry deviations [108,109]. Nevertheless, reported values for conductivity of Nb2O5 are wide-ranging
and inconsistent. For amorphous Nb2O5, Fischer et al. [110] reported electrical conductivities of
1011 S/cm, while Cavigliasso et al. [111] reported 1013–9  1013 S/cm, and Macek and Orel [112]
16 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

reported conductivities in the order of 1012–1011 S/cm, all at room temperature. Regarding crys-
talline Nb2O5 samples, Soares et al. [50] report conductivity values, at room temperature, that range
between 3.7  1010 S/cm (for the T phase) and 7.6  107 S/cm (for the H phase). Graça et al. [84],
prepared polycrystalline Nb2O5 samples by the sol–gel method and observed conductivities ranging
from 106 S/cm (for the H phase) to 1013 S/cm (for the T phase). Schäfer et al. [57], by its turn,
reported a conductivity of 3  106 S/cm for the H-Nb2O5 phase. While all these results are in fact
inconsistent, it appears that the H phase is often reported to be more conductive than other stoichio-
metric niobium pentoxides. Since the electrical conductivity of these materials are essentially related
to oxygen deficiencies, a more detailed discussion on the conductivity mechanisms is presented in the
following section regarding the non-stoichiometric niobium oxides.
The reported values for the dielectric constant of the Nb2O5 are not consistent as well. For amor-
phous Nb2O5 films formed by anodization, depending on the forming electrolyte Cavigliasso et al.
[111] report dielectric constants between 49 and 120, while Schultze and Lohrengel [106] report val-
ues between 41 and 46. For the T-Nb2O5, there are reported dielectric constants between 40 and 200 at
100 kHz (at room temperature) depending on the synthesis method and orientation [50,84]. For the H
phase, values between 38 and 170 at 100 kHz (at room temperature) have been reported [50,84]. In
most works, is not clear at which frequency and temperature these dielectric constant values are refer-
ring to, but it is typically stated [13,113–115] that the static dielectric constant of niobium pentoxide
is 41 when comparing with the static dielectric constant at room temperature of other materials for
the purpose of dielectric applications. Very recently, El-Shazly et al. [116] calculated of the electronic
structure and optical properties of the B-Nb2O5 polymorph, reporting a static dielectric constant of 5
and a refractive index from 2.5 to 3 within the visible range of the electromagnetic spectrum.
In fact, besides the reported potential [13,46,110,114] for the production of solid electrolytic capac-
itors, the niobium pentoxides possess a wide range of interesting properties which makes this system
suitable for many different applications. The high dielectric constant makes Nb2O5 an interesting
material for complementary metal–oxide–semiconductor (CMOS) devices, or MIM tunnel diodes, or
MIM capacitors [117]. Nb2O5 is also commonly reported for its photo and electrochromic properties
[112,118–125] being possible to change the colour of thin films by applying a voltage. It has also been
used as photoelectrode for dye-sensitized solar-cells (DSSCs) as an alternative to, or together with,
TiO2 [87,118,126,127], offering higher open circuit voltages [128] and the possibility to achieve a
higher light absorption coefficient by inducing oxygen deficiencies [129]. Another active topic of
research regarding the niobium pentoxides is their application as a catalyst [130]. Given its high cat-
alytic activity, selectivity at low temperatures and stability, specially of amorphous hydrated
Nb2O5nH2O, it has been used to catalyse different types of reactions such as esterification, hydrolysis,
dehydration, condensation or alkylation [103,130]. The use of niobium pentoxide as a catalyst for
hydrogen storage has also been reported [131]. The application of Nb2O5 in lithium batteries
[86,118], humidity sensors [132], electrochemical biosensors [133], and as a biomaterial (given its
chemical stability and low cytotoxicity) [134–136] have been reported. The incorporation of Nb2O5
in different glass systems [137–141] has been explored essentially for their interesting non-linear
optical properties. The production of rare-earth doped Nb2O5 thin films for optical waveguides and
amplifiers have also been recently reported [142]. Niobium pentoxide also finds applications as hard
coatings for optical glasses and lenses by taking advantage of its low optical absorption coefficient,
high refractive index, chemical and thermal stability and mechanical resistance [143]. Optical trans-
parency of Nb2O5 is typically reported to be around 90% in the visible range, strongly decreasing for
wavelengths smaller than 400 nm [107,129,144,145]. However, the transparency is highly dependent
on oxygen deficiencies, allowing to achieve black Nb2O5 nanostructures for enhanced solar light
absorption purposes [129]. Amorphous Nb2O5 has a reported refraction index between 2.1 and 2.6
for visible wavelengths, depending on the synthesis method, microstructure and internal stress
[144,146,147], higher than HfO2 and Ta2O5 (which are also used in optical components). In fact,
mechanical stress is one of the major issues to solve in order to achieve high refraction index and
low extinction coefficients [146,147]. Vinnichenko et al. have reported the synthesis of Nb2O5 films
with low mechanical stress (90 MPa) by reactive pulsed magnetron sputtering in a high plasma flow
configuration, which allowed them to achieve refraction indexes as high as 2.54 at 400 nm and an
extinction coefficient as low as 6  104 at 400 nm [146].
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 17

Fig. 14. Monoclinic structure of the Nb22O54 phase [21].

A very recent and promising application of Nb2O5, based on the resistive switching behaviour
reported in niobium oxides in 1965 by Hiatt and Hickmott [4], is the application as resistive random
access memories (ReRAM). This type of application and working principle fits the concept of a passive
electronic component called memristor, which existence was first suggested in 1971 by Chua [148] and
proved to exist only in 2008 by IBM Labs using TiO2 [149]. The memristor is considered to be a fourth
basic circuit element in electronics, along with the concept of resistor, capacitor and inductor, and is
defined by the dependence of its resistance on the history (i.e. integral over time) of the current that
passed through it. This is currently an active and fast-growing area of research [150–158], with many
related works based on niobium oxides [77,159–167].
Recently, a comprehensive review paper was published by Rani et al. [85] regarding the fundamen-
tal properties, some recent applications and particularly synthesis methods of different niobium pen-
toxides micro and nanostructures, such as thin films, nanorods, nanobelts or nanospheres. Such work
corroborates what was aforementioned, particularly the inconsistencies on the identification of the
Nb2O5 polymorphs, their wide range of applications and also stoichiometry issues.

2.5. Non-stoichiometric niobium oxides

It is known that crystallographic shear and point defects are the most common types of defects that
produce/accommodate stoichiometry changes in certain metal oxides (such as TiO2) [93,94,168] but
which are especially easy to occur in niobium oxides [51,95–97,169]. However, point defects cannot
explain high variations of stoichiometry since their concentration in a crystal lattice is rather limited
(usually not exceeding 104) [95]. Crystallographic shear on the other side, leads to the formation of
planes where the metal oxide octahedra changed their linkage (e.g. from corner to edge-sharing) thus
accommodating large oxygen deficiencies keeping the metal ion coordination [93,95]. In addition, the
formation of such ordered defects can lead to the formation of different phases of the metal oxide. Cit-
ing Mrowec [94], ‘‘[. . .] the mutual interaction between point defects results in formation of complexes and
defect clusters called extended defects. These extended defects can become further ordered, which leads to
superstructure ordering and to formation of intermediate phases. In some cases point defects can become
eliminated in the process of crystallographic shear which is connected with formation of a whole series
of intermediate phases.”
It is possible to find several non-stoichiometric niobium oxides reported in literature. Essentially,
these can be divided in two groups: one with stoichiometry between Nb and NbO, and other with sto-
ichiometry between NbO2 and Nb2O5, as illustrated before by Fig. 1. There are three reported non-
stoichiometric niobium oxides between Nb and NbO, which were found to be metastable and were
18 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

Fig. 15. (a) Structure of the orthorhombic Nb12O29 phase; (b) crystallographic shear planes depicted with the light blue NbO6
octahedra [21].

classified as NbOx, NbOy and NbOz [170]. NbOx was reported to be formed at 270–500 °C with a sto-
ichiometry equivalent to Nb6O, NbOy was formed at 330–500 °C with stoichiometry equivalent to
Nb4O, and NbOz was formed at 400–700 °C with an unknown exact stoichiometry [170]. To date, there
are very few published works on this group of metastable non-stoichiometric niobium oxides.
Most of the reported non-stoichiometric niobium oxides are about those having a stoichiometry
between NbO2 and Nb2O5, and more precisely between NbO2.4 and NbO2.5. Schäfer et al. [8] referred
the existence of such phases, with close but different stoichiometries, classified as ox I to ox VI, and
suggested that all, or some of such phases might belong to the same mixed-crystal phase. Subsequent
works, Schäfer et al. [57] and Kimura [171], reported the observation of measurable single phase
ranges of non-stoichiometric niobium oxides, such as Nb12O29, Nb22O52, Nb47O116, Nb25O62 and
Nb53O132. Later studies, between 900 and 1300 °C, reported different results which were summarized
and classified as ‘‘incomplete and inconsistent” by Marucco [108].
Hence, Marucco [108] performed a critical review and a detailed study in order to clarify the ther-
modynamic existence of such phases, ‘‘[. . .] by the method of equilibration between oxides and buffer gas-
eous mixtures, which is the most reliable”. However, between 1000 and 1100 °C, Marucco only found
two stable phases within the same composition range (2.4 < x < 2.5)—Nb12O29 (x = 2.42) and
Nb25O62 (x = 2.47)—suggesting that the previously reported non-stoichiometric phases were meta-
stable or were associated with the sensitivity of the different methods that were used, or by the dif-
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 19

Fig. 16. (a) Structure of the monoclinic Nb12O29 phase; (b) crystallographic shear planes depicted with the light blue NbO6
octahedra [21].

ficulty in keeping the phases unchanged after quenching. In a later study regarding the electrical resis-
tance of niobium oxides, Marucco [172] explained that phases with very small stoichiometry devia-
tions from the stable phases of Nb2O5, Nb12O29 and Nb25O62, were in fact possible to be found and
that this deviation was not related to coherent intergrowth in the crystal, but instead to point defects,
oxygen vacancies in stable phases, and niobium interstitials in metastable phases (e.g. Nb22O54 was
found to be stable only above 1250 °C).
Overall, Marucco [172] concluded that the only stable phases of niobium oxides with a stoichiom-
etry between NbO2.4 and NbO2.5 are Nb2O5, Nb12O29 and Nb25O62 where such variations in stoichiom-
etry are possible to occur, and can be interpreted as single or doubly charged oxygen vacancies in their
structure, resulting in significant variations of the electrical resistance. Since then, apart from the
Nb12O29, there have been almost no reports on non-stoichiometric niobium oxide phases that could
clarify these doubts about the stability of phases and their physical properties. Still, there is a recent
work reporting the synthesis and structural characterization of the crystalline phase of Nb22O54 [173].
20 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

Digesting the available information to date, one can say that only three crystalline phases, stable at
room temperature, of non-stoichiometric niobium oxides were reported: the monoclinic Nb22O54, and
the monoclinic and orthorhombic Nb12O29 phases. The several other non-stoichiometric phases are
reported to be metastable, i.e. to exist only at high temperatures.
The synthesis of these non-stoichiometric phases may be achieved by a controlled oxidation of
lower niobium oxides (e.g. NbO2, NbO or even metallic Nb), or by heat treating Nb2O5 samples in a
reducing atmosphere at high temperatures (from 900 to 1300 °C) [8,57,108,172]. It was also recently
reported the synthesis of crystalline Nb12O29 (both monoclinic and orthorhombic phases), in the form
of thin films, by dc magnetron off-axis sputtering at room temperature, using metallic Nb as a target,
after what they were annealed in vacuum at 1000 °C to promote the crystallization [174]. Moreover,
McQueen et al. [173] successfully grown single crystals of orthorhombic Nb12O29 and Nb22O54 by mix-
ing the right amounts of NbO2 and Nb2O5 powders into the desired stoichiometry, and by subse-
quently melting the mixtures in a vacuum furnace back-filled with argon at 1425 °C.
The single crystals obtained by McQueen et al. [173], allowed them to refine and report, for the first
time, the crystal structure of the Nb22O54. This phase crystallizes in a monoclinic structure (space
group C12h) with two formula units per unit cell, as illustrated in Fig. 14. The structure is described
by 3  3 and 4  3 blocks of NbO6 octahedra linked by corner-sharing, and by Nb atoms in a tetrahe-
dral coordination. The Nb atoms with octahedral coordination occupy six 2m, four 2n and one 1e
Wyckoff positions, while the Nb atoms with tetrahedral coordination occupy a 2i, a 2n and a 2m posi-
tions with occupancies of 0.403, 0.068 and 0.029 respectively. The O atoms occupy thirteen 2m, thir-
teen 2n, one 1c and one 1d Wyckoff positions. It should be noticed that in the 3  3 and 4  3 blocks,
the niobium atoms are displaced from the middle of the octahedra towards the centre of the block,
forming an antiferroelectric ordering of the electric dipoles [173].
The Nb12O29 phase is the most commonly reported non-stoichiometric niobium oxide phase. This
phase can crystallize into an orthorhombic structure (space group D17 2h), or in a monoclinic structure
(space group C32h) [173,175]. The structures of Nb12O29, both orthorhombic and monoclinic poly-
morphs, are based on 4  3 blocks of corner-sharing NbO6 octahedra, where the blocks are linked
by edge-sharing. The orthorhombic structure of Nb12O29, illustrated in Fig. 15, has six formula units
per unit cell, where the Nb atoms occupy six 8f Wyckoff positions and the O atoms occupy thirteen
8f and three 4c positions [173,175]. The structure of the monoclinic polymorph of Nb12O29, illustrated
in Fig. 16, has three formula units per unit cell, with the Nb atoms occupying six 4i Wyckoff positions
and the O atoms occupying one 2d and fourteen 4i positions [175]. Similarly to Nb22O54, both Nb12O29
have the Nb atoms displaced from the centre of the NbO6 octahedra, towards the centre of each block.
These displacements are typically responsible for a spontaneous polarization and electric polarization
in ferroelectrics, but in this case they are arranged in such way that each block has no net electrical
polarization, and therefore these materials are considered to be antiferroelectric [173]. These struc-
tures are a clear example of how stoichiometry deficiencies can give rise to crystallographic shear
planes, as illustrated in Figs. 15 and 16 by the light blue octahedral connected by edge-sharing.
Even though the stoichiometry of these materials is near of the Nb2O5, which is a semiconductor,
the electrical properties are quite different. Schäfer et al. [57], Janninck and Whitmore [176] and later
Marucco [172], investigated on the electrical properties of non-stoichiometric niobium oxides, verify-
ing that even a small change of the O/Nb ratio produced a significant variation of the conductivity,
showing a typical n-type conductivity. Schäfer et al. [57] reported a variation from 3  106 to
4  101, and to 3  103 S/cm, by changing the O/Nb ratios from 2.5 to 2.495, and to 2.489, respectively
(at room temperature). One should notice that these conductivity variations for different stoichiome-
tries do not refer to metastable phases or a mixture of phases. Instead, the Nb2O5 sub stoichiometries
are achieved by having different concentrations of oxygen vacancies, which can give rise to phase
changes at very high concentrations, i.e. Nb22O54 and Nb12O29.
Janninck and Whitmore [176] studied the electrical conductivity, between 77 and 1273 K, of the H-
Nb2O5 polymorph with more than 20 different stoichiometry deviations, from Nb2O4.9992 to
Nb2O4.8568, and recorded values between 103 and 103 S/cm (depending on temperature and compo-
sition) in an inert atmosphere. The authors confirmed a dependence of the electrical conductivity with
the oxygen partial pressure to the power of 1/6, which is consistent with a defect structure model
based essentially on double ionized oxygen vacancies. Additionally, the thermal dependence of the
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 21

electrical conductivity has shown two distinct regions, one with a positive temperature coefficient at
low temperatures, and another with a negative temperature coefficient at higher temperatures. An
explanation for this behaviour was proposed by the authors [176] as being due to the thermal ioniza-
tion of the electrons from the donor centres (positive temperature coefficient), until a saturation
regime is reached, i.e. the number of charge carriers remain constant. As the temperature further
increases, the decrease of the conductivity is explained by the decrease of the mobility of the charge
carriers which was found to be proportional to T0.5, in agreement with an optical mode scattering
mechanism. By assuming that the oxygen vacancies are all double ionized at 1000 °C, the magnitude
of the carrier mobility was calculated to be 0.218 cm2/(V s), and found to be independent of the stud-
ied stoichiometries [176].
Marucco [172] performed a more detailed study on the conductivity of niobium oxides, at high
temperatures, in a broad range of stoichiometries between Nb12O29 and Nb2O5, as a function of oxygen
pressure and temperature, also observing an increase of conductivity with the decrease of O/Nb ratio.
Marucco verified that for an oxygen partial pressure above 107 atm, the electrical conductivity beha-
viour suggested that the defect structure on these materials was predominantly characterized by sin-
gle ionized oxygen vacancies. However, for lower pressures (down to 1017 atm) all the stable non-
stoichiometric phases have shown a dependency consistent with double ionized oxygen vacancies,
in agreement with what Janninck and Whitmore [176] observed. Additionally, Marucco observed pro-
nounced discontinuities of the electrical conductivity in metastable phases (compared to stable
phases), which was related to the presence of niobium interstitials. It was suggested that the forma-
tion of crystallographic shear planes was enabled by the presence of such Nb interstitials [172]. Later
studies [175,177] specifically about Nb12O29, report a typical metallic behaviour regarding the electri-
cal conductivity, where Cava et al. [177] recorded values in the order of 103 S/cm, but without showing
superconductivity at least down to 0.25 K.
The electrical conduction of amorphous Nb2O5 thin films, at low pressures (1010 atm), was inves-
tigated by Jouve [115] who was able to identify different conduction processes depending on the tem-
perature and applied field. While the author never suggests a stoichiometry deficiency in the samples,
the results are consistent with the presence of oxygen vacancies in the amorphous Nb2O5 films. Jouve
[115] emphasized the role of localized trap levels, below the conduction band, in the electrical prop-
erties of these materials. At low fields and at room temperature, typical ohmic conductivity was
observed and associated to variable-range hopping. Above 360 K and at low fields, the results were
consistent with small-polaron model of nearest-neighbour hopping, which is activated by an optical
multiphonon process. As the voltage is increased (at room temperature), space-charged limited cur-
rent density obeying the Mott–Gurney law was observed. At fields higher than 5  106 V m1, as
the space-charge effects becomes negligible, the observed current density was associated to the

Table 2
Band gap energy values reported for niobium oxides.

Phase Eg (eV) Ref.


NbO2 (tetragonal) 0.5 Ceramic pellet [184]
0.7 Thin films [70]
0.88 Thin films [185]
1.0 Thin films [65,71]
1.16 Ceramic pellet [186]
Amorphous Nb2O5 3.4 Thin films [105]
3.4–5.3 Thin films [106]
3.4 Sol–gel powders [89]
TT-Nb2O5 3.4 Sol–gel powders [89]
T-Nb2O5 3.4 Sol–gel powders [89]
3.4 Single-crystal [50]
H-Nb2O5 3.4 Sol–gel powders [89]
3.0 Single-crystal [50]
B-Nb2O5 2.55 Calculated [70]
M-Nb2O5 3.75 Thin films [107]
3.36–4.15 Quantum dots [98]
22 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

Table 3
Electrical conductivity and dielectric constants of niobium oxides.

Phase rdc (S cm1) @ 300 K Ref. e0 @ 300 K Ref.


4
Nb 6.5  10 [20]
NbO 4.8  104 [11]
NbO2 (tetragonal) 104 [56] 10 1 kHz–10 MHz [55]
NbO2 (rutile) 103 @ 1000 K [56]
Amorphous Nb2O5 1011 [110] 41 Static (dc) [113–115]
1013 [111] 41–46 Static (dc) [106]
1012–1011 [112]
T-Nb2O5 3.7  1010 [50] 39 100 kHz [50]
1013–1011 [84] 60–77 100 kHz [84]
H-Nb2O5 7.6  107 [50] 38–160 100 kHz [50]
3  106 [57] 16–130 100 kHz [84]
108–106 [84]
B-Nb2O5 5 Static (dc) [116]
NbOx (2.5 > x > 2.489) 106–103 [57]

lowering of the barriers around a trapping centre localized 0.34 eV below the conduction band,
according to Jouve’s calculations [115]. The author clearly shows that such behaviour cannot be
explained by a Schottky effect, but indeed by a Poole–Frenkel mechanism. Jouve also reported an irre-
versible switching behaviour for fields higher than 5  107 V m1 coherent with a filamentary conduc-
tion process [115], similarly to what happens with NbO2 [78].
In 1991, Cava et al. [177] were the first to report on the magnetic properties of non-stoichiometric
niobium oxides. While the Nb2O5 is diamagnetic, Cava et al. verified that the orthorhombic Nb12O29
phase shows an approximated Curie–Weiss behaviour, i.e. paramagnetic, and an antiferromagnetic
ordering at 12 K. The fact that orthorhombic Nb12O29 shows, simultaneously, a Curie–Weiss beha-
viour, antiferromagnetism and metallic conductivity (even below the ordering temperature), which
are characteristics of localized and delocalized electrons, led the author to suggest that the structure
contained some Nb4+ ions with a localized 4d1 configuration, which give rise to the magnetism, while
the other Nb5+ ions at other sites contribute with delocalized electrons, responsible for the conductiv-
ity. In 2001, Waldron et al. [178] further extend this explanation to monoclinic Nb12O29, but only in
2004 the same authors [175] experimentally verify for this phase an identical metallic conductivity,
allied to a Curie–Weiss behaviour and a antiferromagnetic transition at 12 K, the same ordering tem-
perature as the orthorhombic polymorph. However, the advanced structural study reported by McQu-
een et al. [173], in 2007, shows that based on bond valence results, no Nb4+ sites are detected in the
orthorhombic Nb12O29 phase (at least at 200 K), which suggests that the magnetic behaviour observed
in these materials is not related with the valence electrons of individual Nb atoms, but instead, as the
authors suggest, ‘‘[. . .] each magnetic electron may be delocalized across a 4  3 block [. . .]”. Up to date,
this question about the magnetic nature of the Nb12O29 phases remains unclear. Moreover, it should
be noticed that Cava et al. [177] verified that other non-stoichiometric niobium oxide phases dis-
played paramagnetic behaviour, but not any kind of ordering, at least down to 2 K.
In 2011 Ohsawa et al. [174] reported the synthesis of a Nb12O29 thin films (with a thickness of
120 nm) which have shown a high transmittance in the visible range (50% in the red and 70% in
the blue regions), with a refractive index n = 2.2 at 400 nm. The films were deposited on a glass sub-
strate with a transmittance between 90% and 95%. The high transparency, allied to the high electrical
conductivity of about 300 S/cm (even after annealing at 1000 °C in vacuum) without the need of dop-
ing, makes Nb12O29 a new class of transparent conductive oxides (TCO). Classified by the authors as an
‘‘intrinsically doped d-electron-based TCO” [174], and though there is no evidence if it has a direct or
indirect band gap, the Nb12O29 may be a promising material for use in high-performance and low-
cost optoelectronic devices.
With the description of the niobium oxides, the complexity of this system is evident. It also
becomes obvious that an in-depth knowledge about the structures and stoichiometry of these oxides
is fundamental for solid scientific research or towards any type of technological application. Table 1
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 23

Fig. 17. Typical perovskite structure of an alkali niobate [21].

enumerates all the stable niobium oxides phases reported to date, while Tables 2 and 3 summarize
reported values of band gap energy, electrical conductivity and dielectric constant for these materials.

3. Niobates

3.1. Alkali niobates

Lithium Niobate (LiNbO3), Sodium Niobate (NaNbO3) and Potassium Niobate (KNbO3) are known as
the alkali niobates. These oxides, structural classified as perovskites, have been reported by many of
their interesting properties such as piezoelectricity, pyroelectricity, electro-optic and nonlinear optical
behaviour [187,188].
More particularly, LiNbO3, possibly the most studied of the alkali niobates, is a well-known trans-
parent semiconductor (with a band gap of 3.7 eV [189]) with an extremely technological importance.
It presents large pyroelectric, piezoelectric, acousto-optic, nonlinear and electro-optic coefficients, and
it is mainly used, in solid-state based optical devices, as optical modulators, for Q-switching, for sec-
ond harmonic generation (SHG), optical filters, waveguides or acoustic-wave transducers, either in
bulk or nanostructured forms [187,190]. Due to its wide band gap, LiNbO3 has also been reported
as an interesting host for achieving optically active lanthanides via suitable doping processes
[191,192].
Similarly, KNbO3 is reported for having large nonlinear optical coefficients which makes it suitable
for SHG, frequency mixing, or optical parametric oscillators [190]. Moreover, KNbO3 and also NaNbO3
are pointed as promising alternatives to the currently used piezoelectrics ceramics, mainly based on
lead zirconate titanate (PZT). The need of finding a substitute for PZT is essentially related with envi-
ronmental and health concerns, since it presents a weight percentage of lead 600 times greater than
the maximum limit imposed by governmental regulations [188]. Thus, the alkali niobate (K,Na)NbO3,
typically abbreviated as KNN, is probably the top candidate to replace PZT, with typical reported
piezoelectric constant d33 around 80 pC/N for an optimal K/Na ratio 1 [188]. The most common
approach for property optimization of KNN involves forming solid solutions with other materials such
as LiNbO3, LiTaO3, LiSbO3, BaTiO3, CaTiO3, (Bi0.5Na0.5)TiO3, or a combination of these, resulting in d33
constants in the order of 200 and 300 pC/N [188]. The crystalline structure of KNN ceramics is not sim-
ple. While the pervoskite subcell of this material has a monoclinic symmetry (illustrated in Fig. 17),
the KNN crystalline structure is orthorhombic with a space group C14 2v . The structure of K0.5Na0.5NbO3
24 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

Fig. 18. Typical orthorhombic structure of most columbite niobates [21].

is not yet refined and does not have a standard JCPDS–ICDD file, and therefore compromises a correct
crystal indexing literature. Furthermore, the KNN ceramics have different polymorphs depending on
the temperature and on typically used dopants such as Li, Ta, Sb or others [188].
Despite the fact that these alkali niobates have been presenting extremely good piezoelectric prop-
erties, their application is still limited by difficulties in their synthesis. In fact, one of the current big-
gest challenges in the study of alkali niobates is the control and reproducibility of their properties,
which are typically associated to the synthesis methods and sintering processes. It is known that small
variations of the sintering temperature may easily result in deteriorated performance [193]. Addition-
ally, there are two types of mechanisms that were proposed to justify the piezoelectric property
enhancement – polymorphism phase transition (PPT) and morphotropic phase boundary (MPB). Sum-
marizing, it becomes obvious that the investigation and control of the phase and crystalline structures
of these materials is extremely important [188].
One of the most common precursors for the synthesis of niobates is some sort of niobium oxide,
particularly Nb2O5. However, as explained above, due to the ‘‘memory of solids” the original synthesis
method of the niobium oxide can be determinant in the properties of the final product, thus affecting
the properties of the niobate in different ways. Furthermore, it is known [8] that there is a direct struc-
tural relationship between the form of Nb2O5 and the related ternary oxides. Moreover, Kuznetsova
et al. [194] ‘‘[. . .] have demonstrated that the polymorphous phase state of Nb2O5 significantly influences
the characteristics of the alkali metal niobate solid solution powders and determines the extreme variation
of the properties of synthesized ceramics, which must be taken into account in development of the Nb-
containing ferroelectric piezoceramics”. Hreščaka et al. [195] have recently published a work on this
issue, alerting for the need of a good identification of the Nb2O5 polymorph when reporting the syn-
thesis of alkali niobates. This is why it is believed that the experience and knowledge acquired, while
exploring the complex system of niobium oxides, may constitute an important basis in the research of
different niobates.
Additionally, there have been also recent works regarding multiple hetero-nanostructures based on
alkali niobates. This new class of materials are very interesting for the possibility of effectively cou-
pling different properties. Due to their properties and multifunctional character, hetero-
nanostructured systems can be more advantageous than single component systems. Particularly,
Yan et al. [196] reported the synthesis of multiple NaNbO3 nanoplates inside hollow Nb2O5 nanotubes.
The authors classified these structures as a novel class of multiple ferroelectric/semiconductor
heterostructures as they exhibit distinct ferroelectric switching due to the presence of the strain states
at the NaNbO3/Nb2O5 interface.
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 25

3.2. Columbite niobates

The columbite niobates (which have their name from the same crystalline structure of the colum-
bite mineral), have the general formula M2+Nb2O6 where M2+ is a divalent cation of Mg or Ca or of a
transition metal such as Cu, Cd, Zn, Ni, Mn, Co or Fe, and have been reported especially as microwave
ceramics [197]. These materials crystallize in a orthorhombic structure (space group D14 2h) character-
ized by zigzag chains linked by corner sharing, each chain formed by MO6 and NbO6 edge-sharing
octahedra, as illustrated by Fig. 18 [197]. The exceptions to this type of structure, in columbite nio-
bates, are those in which the divalent ion has a ionic radius higher than 1.0 Å, such as Ba2+ and
Sr2+, promoting the crystallization in other types of orthorhombic structure [197].
Microwave ceramics are characterized as being dielectric resonators with a frequency in the micro-
wave region (500 MHz to 10 GHz), where a high dielectric constant and low dielectric losses are deter-
minant characteristics. These materials find an important application in wireless communications
which is, and has been, a highly demanding technology. Therefore, it is natural that there is a constant
search for less expensive and high performance dielectric ceramics, especially because the materials
currently used are tantalum-based complex perovskites, such as BaZn0.33Ta0.67O3 (BZT) or BaMg0.33-
Ta0.67O3 (BMT) [197].
As an alternative, there are niobium-based perovskites, such as BaZn0.33Nb0.67O3 (BZN) or Ba(Zn/
Co)0.33Nb0.67O3 (BZCN), which are less expensive, but require high sintering temperatures (c.a.
1600 °C) and times, and while the relative permittivity is superior, the overall dielectric properties
are not yet competitive enough.
The reported interest in columbite niobates (especially ZnNb2O6) as microwave ceramics is related
to their dielectric properties, similar to BZN or BZCN, and significantly better than columbite tantalates
(M2+Ta2O6). In contrast with the tantalate and niobate perovskites, these materials do not require sin-
tering at such high temperatures, typically between only 1100 and 1200 °C, and also have just a super-
ficial degree of ordering [197]. Additionally, because of the simpler chemistry of binary compounds,
these materials require a less complex processing than the perovskites. Nevertheless, it can be con-
cluded from the literature that many columbite niobates are not synthesized simply by one step. Dif-
ficulty in forming a single-phase samples or requiring two or more sintering steps are commonly
reported. Furthermore, the occurrence of point defects in the lattice, dislocations, grain boundaries
and porosity are pointed as frequent problems that compromise and justify the variation of the dielec-
tric properties of these materials. It should be noticed that, also in these materials, non-stoichiometric
phases are commonly produced as well. It was for instance reported that a slightly reduced man-
ganese niobate (MnNb2O6x) presents a much higher electrical conductivity, assigned to the produc-
tion of delocalized electrons upon a small degree of Nb reduction [197]. Generally, it was observed
that a small stoichiometry variation for columbite niobates produce huge variations in their dielectric
properties, thus compromising their application [197]. It seems possible the cause of these type of
observations may be closely related to the analogous behaviour of niobium oxides, and also related
with the phase of niobium pentoxide used as precursor for the niobates, but further investigation is
nevertheless necessary.
Besides the application as dielectric ceramics, columbite niobates have also been reported by their
magnetic properties, which in general present antiferromagnetic ordering below temperatures 4–7 K
with special interest on CoNb2O6 that shows some unusual magnetic ordering due to geometrical frus-
tration [197]. The reported NiNb2O6 photocatalytic properties, for water hydrolysis, are quite interest-
ing but it was shown that this material was easily corroded, thus compromising its application. On the
other hand, Zn and Mn niobates were also reported to have interesting photocatalytic properties for
isobutene oxidation, presenting a good chemical stability and selectivity.
The optical properties of both pure and rare-earth-doped columbite niobates have been reported in
literature [197]. The pure columbite niobates can be considered wide band gap semiconductors with
reported values ranging from 2.2 to 3.9 eV, depending on the M2+ ion [198–201], and have shown (at
least for Ca, Zn, Cd and Mg niobates) intrinsic blue luminescence with broad bands centred c.a. at 430–
470 nm [197]. Hence, these materials were classified as self-activated phosphors. It was reported that
the luminescence efficiency decreased with decreasing size of the divalent ion, i.e. Ca niobate was
found to have the most efficient emission. Particularly for Ca niobate, it was found that the
26 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

464.5 nm emission was preferentially excited with 200–260 nm wavelength radiation [197]. Partial
substitution of Ca by other elements, such as Zn and Cd, were reported to enhance the blue lumines-
cence intensity, which was assigned to the participation of the 4d electrons in the charge transfer
within the NbO6 octahedra [197]. Ca niobate is transparent in a wide spectral range (300 nm to
5.5 lm), it has relatively low phonon cut-off energy (900 cm1) and a high refractive index is com-
monly reported (between 2.07 and 2.2) although it is not specified at which wavelengths
[197,202,203]. Additionally, single crystals of Ca niobate are rather easy to grow by laser-heated ped-
estal growth technique, which allied to these properties makes it a preferable host for rare-earth dop-
ing, aiming their application as optical devices [197]. Thus, there are works reporting laser action in Ca
niobate doped with Nd, Ho, Pr and Tm, or even co-doped with Eu and Ti which was proposed as a low
cost lamp phosphor [197]. Europium was found to be optically active in doped Cd niobate (band gap
between 3.35 and 3.5 eV) [204,205], showing higher intensity, when excited and detected under the
same experimental conditions, compared with the Ca niobate host and with strong evidences of mul-
tiple Eu3+ emitting sites [197]. Single crystals of CaNb2O6 doped with neodymium have shown stim-
ulated emission at 1064 nm, due to intraionic Nd3+ transitions, with a lifetime of only 145 ls; the
samples doped with erbium showed a pink colouration and an intense emission at 1550 nm at room
temperature (due to transition between the first excited multiplet and the ground one of the Er3+), and
similarly, those doped with praseodymium showed a sharp and intense band centred at 610 nm due to
the intraionic emission of the Pr3+ [197]. Additionally, CaNb2O6 was reported to have a strong blue
mechanoluminescence, observed during grinding [197]. On the other hand, nanoparticles of Ni niobate
doped with Dy, prepared by a sol–gel combustion method, did not show any characteristic lumines-
cence of the lanthanide ion in its trivalent charge state, but it was found that the intrinsic blue emis-
sion of the host was enhanced, which was assigned to the energy transfer from the Dy3+ ion to the
NbO6 octahedra. This behaviour was observed also for Zn niobate doped with Dy, but it was also found
that for >5 mol% Dy there was an opposite behaviour assigned to concentration quenching effect
[197].
There has also been some works reporting on another type of niobate perovskite based on rare-
earths with very interesting optical properties, particularly the CaRENb2O10 (where RE is a trivalent
rare-earth ion) [206]. These materials are classified as triple-layered perovskites, as they are built
by layers of corner-sharing octahedra. In a very recent work (2013), Qin et al. [206] report the synthe-
sis and characterization of these ‘‘novel luminescent materials” with an estimated band gap of 4.42 eV.
With 266 nm (4.7 eV) laser excitation, and so above the material band gap, the authors observed a
broad emission band which was assigned to the NbO6 octahedra. Depending on the lanthanide, the
intrinsic NbO6 broad band varied its spectral position between 400 and 500 nm, from blue to greenish
blue, which is preferentially excited at 270 nm. The characteristic intra-4f emissions of the trivalent
rare-earth ions were also observed under 266 nm excitation, namely for the Sm3+, Er3+, Dy3+ and Eu3+
ions, which shows that there was a charge transfer mechanism from host to the lanthanide. It is noted

Fig. 19. Structure of the monoclinic phase of RENbO4 [21].


C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 27

Fig. 20. Structure of the tetragonal phase of RENbO4 [21].

that these materials show a high critical point for luminescence concentration quenching effects. This,
and the high luminescence intensity of the lanthanides characteristic emissions at room temperature,
lead the authors to suggest these materials to be used as phosphors in near-ultraviolet or blue light
emitting diodes (LEDs), especially the CaEuNb3O10 [206].

3.3. Rare-earths orthoniobates

There are works reporting rare-earth orthoniobates (RENbO4), as well as rare-earth orthotantalates
(RETaO4), from decade of 1960 [207–211], but until recently these systems were not a research topic
which has received much attention from the scientific community.
The RENbO4 are known to crystallize in a fergusonite-type structure, i.e. in a monoclinic structure
(space group C62h), undergoing a pure and reversible ferroelastic phase transformation to a scheelite-
type structure, with a tetragonal system lattice (space group C64h), only stable at high temperatures
(>700 °C) [211–213]. In the monoclinic phase the Nb and RE atoms occupy the Wyckoff positions
4e while the O atoms occupy the 8f position, which structure is illustrated by Fig. 19 where it is pos-
sible to identify the NbO6 octahedra linked by edge-sharing forming chains along the c axis.
The tetragonal phase is not built on NbO6 octahedra, but instead on unlinked NbO4 tetrahedra, as
illustrated in Fig. 20, with the RE, Nb and O atoms occupying, individually and respectively, the 4b, 4a
and 16f Wyckoff positions. The lattice parameters of the RENbO4 compounds will obviously depend on
the lanthanide ion.
One of the most detailed structural studies on rare-earths orthoniobates, which covers all the lan-
thanide ions (from La to Lu, except Pm), is a work reported by Siqueira et al. [212]. These authors
report the synthesis of rare-earth niobates by solid-state reaction (by mixing niobium pentoxide
and rare-earths oxides), and their characterization by XRD and Raman spectroscopy. It was found that
only for heat treatment temperatures above 1150 °C, the formation of secondary phases could be
avoided. The Raman analysis on the RENbO4 samples revealed very complex spectra, with 18 active
modes, which was found to be in complete agreement with group theory calculations [212]. Further-
more, it was found that some of the vibrational modes follow a decrease of energy with the increase of
the lanthanide ionic radii (which is a natural consequence of the unit cell expansion), while other
modes followed the inverse trend (justified by a mass effect) [212]. Some of the Raman bands in
the range of 316–343 cm1 showed a tendency to split for lower lanthanide ionic radii, which was
assigned to a higher monoclinic distortion. From all the RENbO4 samples, the Raman spectra of the
Ce and La niobates were classified as an exception to the general trend because they presented a sig-
nificant broadening and down-shifting of their modes energy. This was linked to the proximity of the
transition temperature from the monoclinic to the tetragonal phase, since it decreases from the Lu to
28 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

the La ion. The authors additionally refer the identification of 15 active modes by Fourier-transform
infrared spectroscopy (FTIR) [212].
Rare-earth tantalates are known to have very similar properties and structures [212,214,215]. Still,
besides the fergusonite-type structure, recent studies report on the identification of two additional
different monoclinic polymorphs, of rare-earth orthotantalates stable at room temperature, depending
on the lanthanide ion and on the synthesis/processing conditions [214,216]. Such polymorphs are
mainly characterized by different arrangement and linkage of distorted TaO6 octahedra, presenting
different physical properties. The same was not observed for orthoniobates [212], but given the obvi-
ous similarities between these two systems, it would not be surprising that such polymorphs also
exist for RENbO4, and for that reason further research is necessary to clarify this point.
While the synthesis of RENbO4 crystals has been previously reported by using Czochralski or flux
growth methods [211], only in 2004 Octaviano et al. [217] report the synthesis of single crystals by a
floating zone technique. The authors have grown single crystals of La, Ce, Pr, Sm, Eu, Dy Ho, and Er
orthoniobates, and a few more doped with Er3+, and mentioned the particular difficulty in controlling
the crystal diameter during the growth. It was verified by X-ray Laue diffraction analysis that the crys-
tals tend to grow along the [1 1 0] direction. Octaviano et al. [217] mainly focused on the Er-doped
LaNbO4 material which showed very interesting optical properties, with high absorption coefficient
and a strong luminescence of Er3+ when pumped with a green 514 nm laser line at 300 K. Recently,
Graça et al. [218] report on the synthesis of EuNbO4 crystals, grown by laser floating zone, and their
optical and dielectric properties. The luminescence analysis of the EuNbO4 crystals revealed intense
red luminescence at room temperature with ultraviolet excitation, and the existence of multiple
Eu-related optical centres whose emission is observed at low temperatures some of which are
quenched at c.a. 50 K. Other authors, Xiao and Yan [219], report on the synthesis, by a modified in-
situ chemical co-precipitation method, and the luminescence of different doped rare-earth orthonio-
bates. Very recently, from the end of 2013, a patent registered by Sandia National Laboratories [220]
claims on the production of rare-earth niobates and tantalates, by an hydrothermal technique, aiming
their application as nanophosphors to lighting and display technologies, which testifies the interesting
optical properties of these materials.
Additionally, rare-earth orthoniobates have been reported to have a characteristic ferroelasticity
attributed to the formation of domain walls during phase transitions [213,217]. The microwave dielec-
tric properties and effects of such ferroelasticity on these properties were investigated by Kim et al.
[213]. The authors estimated the mean static dielectric constants of several orthoniobates using the
Clausius–Mossotti relation. The experimental dielectric constants of the RENbO4 samples, however,
were significantly lower: 20, almost half than the estimated values. This was related to a monoclinic
distortion which has its origin in the phase transformation from the tetragonal to the monoclinic
structure, resulting in ferroelastic domain structure that resembles twinning, with two orientation
states identical in structure but different in orientation [213]. Despite the low dielectric constants
measured by Kim et al. [213], the RENbO4 samples presented excellent quality factors, ranging
between 33,000 and 56,600 GHz, depending on the lanthanide, and also high temperature coefficients
of resonant frequency (sf) Particularly, LaNbO4 presented a positive sf of 9 ppm/K. The transition tem-
perature of several RENbO4, listed by the authors, was found to be tunable by mixing two different
lanthanides, thus allowing to control the strain induced by the lattice distortion, and therefore the
dielectric properties. The high-temperature stability of the resonant frequency and quality factor,
make the rare-earth orthoniobates interesting ceramics for microwave applications [213]. In other fre-
quency range, from 100 Hz to 2 MHz, Graça et al. [218] verified that, for EuNbO4 crystalline fibres, the
dielectric constant depends on the crystal growth speed, being highest (e0 ffi 40) for the slowest pulling
rate. Furthermore, it was found that from 80 to 400 K the dielectric constant of the EuNbO4 samples
remained constant in a remarkable way, which may be interesting for dielectric applications which
require stable operation in a wide temperature range.
Probably, the major interest in the rare-earth orthoniobates is related to the reported proton con-
ductivity, thus having the potential application in sensors and fuel cells electrolytes [215]. The first
work reporting on the protonic conduction on these materials is from 2006 by Haugsrud and Norby
[215], and though the authors have studied both Ca and Sr-doped rare-earth orthoniobates (and also
orthotantalates), the main focus was on proton conductivity and solubility of protonic defects in wet
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 29

H2 atmospheres of Ca-doped RENbO4 samples. The authors clearly identified a pure protonic conduc-
tivity for these samples, increasing with the temperature between 400 and 1200 °C. Generally,
between 400 and 600 °C it was possible to notice a break in the slope of the conductivity, which
was associated with the phase transformation from monoclinic to tetragonal. The Ca-doped LaNbO4
material presented the highest conductivity, reaching 103 S/cm. The analogous orthotantalates
were found to behave similarly to the niobates but with a protonic conductivity almost half an order
of magnitude lower. Additionally, since their phase transition is much higher, the tantalates remained
with the monoclinic structure up to 1200 °C. The authors refer that ‘‘the tetragonal polymorph of the
niobates, in general, shows more exothermic (more negative) hydration enthalpies and is thus more easily
hydrate than the monoclinic polymorph of the tantalates”. Still, since the phase transformation may lead
to thermal expansion, Haugsrud and Norby [215] suggest a partial substitution of Nb by Ta in order to
minimize this effect. The fact that these materials are the oxides (without Ba or Sr as the main com-
ponents) with the highest reported protonic conductivities, make them particularly interesting for
fuel-cell electrolytes and humidity sensors, as aforementioned [215]. It is therefore no surprise that,
since this work, there have been many other reported studies on the protonic conductivity of these
materials, especially on LaNbO4 [221–229].
It is also known that some RENbO4 present antiferromagnetic ordering at temperatures below 4 K,
while most of them follow a Curie–Weiss law [210]. Additionally, in 1998, Baughman et al. [230]
report a large negative Poisson’s ratio (in at least one direction) for LaNbO4, thus being classified as
an auxetic material. This type of property makes these materials interesting for high pressure applica-
tions as, for instance, optical line systems installed at deep sea level.

4. Summary and prospects

The available literature on niobium oxides reveals that this is a complex metal oxide system, with
many phases and polymorphs, and many works reporting contradictory or inconsistent information.
Therefore, it should be emphasized that the review of the physical properties and state of art of nio-
bium oxides, is an important and solid groundwork for future research on these materials. Moreover,
like most electron correlated materials, it was shown that this metal oxide system can lead to many
different and exotic properties, making it a very versatile group of materials (but difficult to work
with). More specifically, stoichiometry problems are one of the most common complications in nio-
bium oxides and in many different niobates, where NbO6 octahedra are typically the basic structural
units.
Niobium oxides have been pointed as an alternative to tantalum for the production of solid elec-
trolyte capacitors, but the diffusion of oxygen easily promoted by the application of a voltage has been
their main limitation [49]. Conversely, these oxides have been showing great potentiality in many
other technological applications, most of them based on one important matter that can lead to differ-
ent properties: oxygen vacancies.
This is the case of the reported photochromic properties of niobium oxides [231] and the synthesis
of Nb12O29 thin films to be used as a transparent conductive oxide (TCO) [174]. Also, it should be
emphasized one of the most interesting and hot topics to which niobium oxides are candidates: the
memristors. In the case of niobium oxides, what may be a problem for the application in the solid elec-
trolytic capacitors, i.e. the oxygen diffusion and variation of conductivity, can be the key factor to
achieve a fully operational memristor. There are some, but still few, works that have been reported
regarding resistive switching in niobium oxides (and other metal oxides) [4,77,155–167,232–235],
and it seems clear that following this research topic will lead the near future reports on these mate-
rials. The synthesis of niobium oxides thin films and nanostructures, and their electrical and optical
characterization is encouraged for future work. While the most common experimental techniques
used to characterize memristive properties are based on cyclic voltammetry [235–238], i.e. recording
hysteretic current-voltage profiles, Messerschmitt et al. [235] have done an interesting work showing
that chronoamperometry and also bias-dependent resistive measurements can be powerful methods
to understand the defect kinetics that give rise to the resistive switching behaviour.
30 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

The fact that Nb12O29 can be a TCO, can enable its application as an interface layer in organic light
emitting diodes (OLEDs) to facilitate the charge injection and extraction to and from the device
[239,240]. Also in this context and inspired in reported works [126] regarding dye-sensitized solar
cells using mesoporous Nb2O5 tubular structures as an alternative to TiO2, it would be interesting
to try using Nb12O29 instead (transparent and conductive) by performing a thermal treatment in a
reducing atmosphere to the Nb2O5 mesoporous structure.
Therefore, future studies on niobium oxides focused in understanding and controlling the phase
transformations, oxidations and oxygen vacancies are crucial to take advantage of the broad range
of properties. Not only fundamental characterization is necessary, but also the synthesis and fabrica-
tion of micro and nanostructures are worth to be further explored.
Additionally, the research and development of different types of niobates, particularly alkali nio-
bates, is often limited by the difficulties on the synthesis and phase stability. Many of such problems
are related to niobium oxygen system and therefore it is believed that a good understanding of the
fundamental properties of the niobium oxides, presented and discussed in this work, are an important
contribution to better understand such niobates.

Acknowledgements

This work is funded by FEDER funds through the COMPETE 2020 Programme and National Funds
through FCT – Portuguese Foundation for Science and Technology under the project UID/
CTM/50025/2013 and RECI/FIS-NAN/0183/2012 (FCOMP-01-0124-FEDER-027494). C. Nico thanks to
FCT for his PhD grant SFRH/BD/68295/2010.

References

[1] Brauer G. Die oxyde des niobs. Z Für Anorg Allg Chem 1941;248:1–31.
[2] Schwartz N, Gresh M, Karlik S. Niobium solid electrolytic capacitors. J Electrochem Soc 1961;108:750–8. http://dx.doi.org/
10.1149/1.2428210.
[3] Ling HW, Kolski TL. Niobium solid electrolyte capacitors. J Electrochem Soc 1962;109:69–70. http://dx.doi.org/10.1149/
1.2425333.
[4] Hiatt WR, Hickmott TW. Bistable switching in niobium oxide diodes. Appl Phys Lett 1965;6:106. http://dx.doi.org/
10.1063/1.1754187.
[5] Cox B, Johnston T. The oxidation and corrosion of niobium (columbium). Trans AIME 1963;227.
[6] Gatehouse BM, Wadsley AD. The crystal structure of the high temperature form of niobium pentoxide. Acta Crystallogr
1964;17:1545–54. http://dx.doi.org/10.1107/S0365110X6400384X.
[7] Bowman AL, Wallace TC, Yarnell JL, Wenzel RG. The crystal structure of niobium monoxide. Acta Crystallogr 1966;21:843.
http://dx.doi.org/10.1107/S0365110X66004043.
[8] Schäfer H, Gruehn R, Schulte F. The modifications of niobium pentoxide. Angew Chem Int Ed Engl 1966;5:40–52. http://
dx.doi.org/10.1002/anie.196600401.
[9] Fromm E. Thermodynamische beschreibung der festen lösung von kohlenstoff, stickstoff und sauerstoff in niob und tantal.
J Common Met 1968;14:113–25.
[10] Allpress JG, Sanders JV, Wadsley AD. Electron microscopy of high-temperature Nb2O5 and related phases. Phys Status
Solidi B 1968;25:541–50. http://dx.doi.org/10.1002/pssb.19680250206.
[11] Pollard Jr ER. Electronic properties of niobium monoxide. Massachusetts Institute of Technology; 1968.
[12] Bach D. EELS investigations of stoichiometric niobium oxides and niobium-based capacitors. Universität Karlsruhe,
Fakultät für Physik; 2009.
[13] Bach D, Störmer H, Schneider R, Gerthsen D, Verbeeck J. EELS investigations of different niobium oxide phases. Microsc
Microanal 2006;12:416–23. http://dx.doi.org/10.1017/S1431927606060521.
[14] Bach D, Störmer H, Schneider R, Gerthsen D. Quantitative EELS analysis of niobium oxides. Microsc Microanal 2007;13.
http://dx.doi.org/10.1017/S143192760708195.
[15] Bach D, Störmer H, Schneider R, Gerthsen D, Sigle W. EELS investigations of reference niobium oxides and anodically
grown niobium oxide layers. Microsc Microanal 2007;13:1274–5. http://dx.doi.org/10.1017/S143192760707359X.
[16] Bach D, Schneider R, Gerthsen D, Verbeeck J, Sigle W. EELS of niobium and stoichiometric niobium-oxide phases—part I:
plasmon and near-edges fine structure. Microsc Microanal 2009;15:505. http://dx.doi.org/10.1017/S143192760999105X.
[17] Bach D, Schneider R, Gerthsen D. EELS of niobium and stoichiometric niobium-oxide phases—part II: quantification.
Microsc Microanal 2009;15:524. http://dx.doi.org/10.1017/S1431927609991061.
[18] Elliott RP. Columbium–oxygen system. Trans Am Soc Met 1960;52:990–1014.
[19] Massalski TB, International ASM. Binary alloy phase diagrams. 2nd ed. Materials Park (OH): ASM Intl; 1990.
[20] Lide DR. CRC handbook of chemistry and physics. 85th ed. CRC Press; 2004.
[21] Momma K, Izumi F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J Appl
Crystallogr 2011;44:1272–6.
[22] Williams WS. Transition metal carbides, nitrides, and borides for electronic applications. JOM 1997;49:38–42. http://dx.
doi.org/10.1007/BF02914655.
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 31

[23] DeArdo AJ. Niobium in modern steels. Int Mater Rev 2003;48:371–402. http://dx.doi.org/10.1179/095066003225008833.
[24] Heisterkamp F, Carneiro T. Niobium: future possibilities – technology and the market place. In: Niobium sci technol proc
int symp, niobium, Orlando, Florida, USA; 2001.
[25] Kommel L, Kimmari E, Saarna M, Viljus M. Processing and properties of bulk ultrafine-grained pure niobium. J Mater Sci
2013;48:4723–9. http://dx.doi.org/10.1007/s10853-013-7210-3.
[26] Halbritter J. On the oxidation and on the superconductivity of niobium. Appl Phys A 1987;43:1–28. http://dx.doi.org/
10.1007/BF00615201.
[27] Periasamy P, Guthrey HL, Abdulagatov AI, Ndione PF, Berry JJ, Ginley DS, et al. Metal–insulator–metal diodes: role of the
insulator layer on the rectification performance. Adv Mater 2013;25:1301–8. http://dx.doi.org/10.1002/adma.201203075.
[28] Matsuno H, Yokoyama A, Watari F, Uo M, Kawasaki T. Biocompatibility and osteogenesis of refractory metal implants,
titanium, hafnium, niobium, tantalum and rhenium. Biomaterials 2001;22:1253–62. http://dx.doi.org/10.1016/S0142-
9612(00)00275-1.
[29] Lindau I, Spicer WE. Oxidation of Nb as studied by the UV-photoemission technique. J Appl Phys 1974;45:3720–5. http://
dx.doi.org/10.1063/1.1663849.
[30] Bouchiat V, Faucher M, Thirion C, Wernsdorfer W, Fournier T, Pannetier B. Josephson junctions and superconducting
quantum interference devices made by local oxidation of niobium ultrathin films. Appl Phys Lett 2001;79:123–5. http://
dx.doi.org/10.1063/1.1382626.
[31] Everitt CWF, DeBra DB, Parkinson BW, Turneaure JP, Conklin JW, Heifetz MI, et al. Gravity probe B: final results of a space
experiment to test general relativity. Phys Rev Lett 2011;106:221101. http://dx.doi.org/10.1103/
PhysRevLett.106.221101.
[32] Seeber B, editor. Handbook of applied superconductivity, vol. 2. Bristol: CRC Press; 1998.
[33] Gupta CK, Suri AK. Extractive metallurgy of niobium. CRC Press; 1993.
[34] Shamrai VF, Blagoveshchenski YV, Gordeev AS, Mitin AV, Drobinova IA. Structural states and electrical conductivity of
oxidized niobium nanopowders. Russ Metall Met 2007;2007:322–6. http://dx.doi.org/10.1134/S0036029507040106.
[35] Grundner M, Halbritter J. XPS and AES studies on oxide growth and oxide coatings on niobium. J Appl Phys
1980;51:397–405.
[36] Bai XD, Zhu DH, Liu BX. Anodic and air oxidation of niobium studied by ion beam analysis with implanted Xe marker. Nucl
Instrum Methods Phys Res Sect B Beam Interact Mater Atoms 1996;114:64–9.
[37] d’Acapito F, Mobilio S, Cikmacs P, Merlo V, Davoli I. Temperature modification of the Nb oxidation at the Nb/Al interface
studied by reflEXAFS. Surf Sci 2000;468:77–84.
[38] Wu AT. Investigation of oxide layer structure on niobium surface using a secondary ion mass spectrometry. Phys C
Supercond 2006;441:79–82.
[39] Batchelor AD, Leonard DN, Russell PE, Stevie FA, Griffis DP, Myneni GR. TEM and SIMS analysis of (1 0 0), (1 1 0), and (1 1 1)
single crystal niobium. Single cryst – large grain niobium technol int niobium workshop, vol. 927. AIP Publishing; 2007. p.
72–83.
[40] Yoon KE, Seidman DN, Bauer P, Boffo C, Antoine C. Atomic-scale chemical-analyses of niobium for superconducting radio-
frequency cavities. IEEE Trans Appl Supercond 2007;17:1314–7. http://dx.doi.org/10.1109/TASC.2007.898059.
[41] Kovacs K, Kiss G, Stenzel M, Zillgen H. Anodic oxidation of niobium sheets and porous bodies. J Electrochem Soc
2003;150:B361–6. http://dx.doi.org/10.1149/1.1585052.
[42] Franchy R, Bartke TU, Gassmann P. The interaction of oxygen with Nb(1 1 0) at 300, 80 and 20 K. Surf Sci 1996;366:60–70.
http://dx.doi.org/10.1016/0039-6028(96)00781-9.
[43] Delheusy M, Stierle A, Kasper N, Kurta RP, Vlad A, Dosch H, et al. X-ray investigation of subsurface interstitial oxygen at
Nb/oxide interfaces. Appl Phys Lett 2008;92:101911. http://dx.doi.org/10.1063/1.2889474.
[44] Kurmaev EZ, Moewes A, Bureev OG, Nekrasov IA, Cherkashenko VM, Korotin MA, et al. Electronic structure of niobium
oxides. J Alloys Compd 2002;347:213–8. http://dx.doi.org/10.1016/S0925-8388(02)00765-X.
[45] Hulm JK, Jones CK, Hein RA, Gibson JW. Superconductivity in the TiO and NbO systems. J Low Temp Phys 1972;7:291–307.
http://dx.doi.org/10.1007/BF00660068.
[46] Qiu Y, Smyth D, Kimmel J. The stabilization of niobium-based solid electrolyte capacitors. Act Passive Electron Compon
2002;25:201–9.
[47] Karnik T. Ceramic powder for use in forming an anode of an electrolytic capacitor. GB 2454049 (A); 2009.
[48] Karnik T. Doped ceramic powder for use in forming capacitor anodes. Patent US 7760487 B2; 2010.
[49] Nico C, Soares MRN, Rodrigues J, Matos M, Monteiro R, Graça MPF, et al. Sintered NbO powders for electronic device
applications. J Phys Chem C 2011;115:4879–86. http://dx.doi.org/10.1021/jp110672u.
[50] Soares MRN, Leite S, Nico C, Peres M, Fernandes AJS, Graça MPF, et al. Effect of processing method on physical properties
of Nb2O5. J Eur Ceram Soc 2011;31:501–6. http://dx.doi.org/10.1016/j.jeurceramsoc.2010.10.024.
[51] Forghany SKE, Anderson JS. Reduction and polymorphic transformation of B-Nb2O5. J Chem Soc, Dalton Trans
1981:255–61. http://dx.doi.org/10.1039/DT9810000255.
[52] Cheetham AK, Rao CNR. A neutron diffraction study of niobium dioxide. Acta Crystallogr B 1976;32:1579–80. http://dx.
doi.org/10.1107/S0567740876005876.
[53] Rimai DS, Sladek RJ. Pressure dependences of the elastic constants of semiconducting NbO2 at 296 K. Phys Rev B
1978;18:2807. http://dx.doi.org/10.1103/PhysRevB.18.2807.
[54] Gervais F, Kress W. Lattice dynamics of oxides with rutile structure and instabilities at the metal–semiconductor phase
transitions of NbO2 and VO2. Phys Rev B 1985;31:4809. http://dx.doi.org/10.1103/PhysRevB.31.4809.
[55] Zhao Y, Zhang Z, Lin Y. Optical and dielectric properties of a nanostructured NbO2 thin film prepared by thermal
oxidation. J Phys Appl Phys 2004;37:3392–5. http://dx.doi.org/10.1088/0022-3727/37/24/006.
[56] Janninck RF, Whitmore DH. Electrical conductivity and thermoelectric power of niobium dioxide. J Phys Chem Solids
1966;27:1183–7. http://dx.doi.org/10.1016/0022-3697(66)90094-1.
[57] Schäfer H, Bergner D, Gruehn R. Beiträge zur Chemie der Elemente Niob und Tantal. LXXI. Die thermodynamische
Stabilität der sieben zwischen 2,00 und 2,50 O/Nb existierenden Phasen. Z Für Anorg Allg Chem 1969;365:31–50. http://
dx.doi.org/10.1002/zaac.1969365010.
32 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

[58] Afanas’ev VV, Stesmans A, Delabie A, Bellenger F, Houssa M, Meuris M. Electronic structure of GeO2-passivated interfaces
of (1 0 0)Ge with Al2O3 and HfO2. Appl Phys Lett 2008;92:022109. http://dx.doi.org/10.1063/1.283166.
[59] Batzill M, Diebold U. The surface and materials science of tin oxide. Prog Surf Sci 2005;79:47–154. http://dx.doi.org/
10.1016/j.progsurf.2005.09.002.
[60] Pascual J, Camassel J, Mathieu H. Resolved quadrupolar transition in TiO2. Phys Rev Lett 1977;39:1490–3. http://dx.doi.
org/10.1103/PhysRevLett.39.1490.
[61] Pascual J, Camassel J, Mathieu H. Fine structure in the intrinsic absorption edge of TiO2. Phys Rev B 1978;18:5606–14.
http://dx.doi.org/10.1103/PhysRevB.18.5606.
[62] Robertson J. Electronic structure of SnO2, GeO2, PbO2, TeO2 and MgF2. J Phys C: Solid State Phys 1979;12:4767. http://dx.
doi.org/10.1088/0022-3719/12/22/018.
[63] Yang Z, Ko C, Ramanathan S. Oxide electronics utilizing ultrafast metal-insulator transitions. Annu Rev Mater Res
2011;41:337–67. http://dx.doi.org/10.1146/annurev-matsci-062910-100347.
[64] O’Hara A, Demkov AA. Nature of the metal-insulator transition in NbO2. Phys Rev B 2015;91:094305.
[65] O’Hara A, Nunley TN, Posadas AB, Zollner S, Demkov AA. Electronic and optical properties of NbO2. J Appl Phys
2014;116:213705.
[66] Wong FJ, Hong N, Ramanathan S. Orbital splitting and optical conductivity of the insulating state of NbO2. Phys Rev B
2014;90:115135.
[67] Vezzoli GC. Electrical properties of NbO2 and Nb2O5 at elevated temperature in air and flowing argon. Phys Rev B
1982;26:3954. http://dx.doi.org/10.1103/PhysRevB.26.3954.
[68] Passier F, Wouters Y, Galerie A, Caillet M. Thermal oxidation of metallic niobium by water vapor. Oxid Met
2001;55:153–63.
[69] Terao N. Structures des oxydes de niobium. Jpn J Appl Phys 1963;2:156–74. http://dx.doi.org/10.1143/JJAP.2.156.
[70] Weibin Z, Weidong W, Xueming W, Xinlu C, Dawei Y, Changle S, et al. The investigation of NbO2 and Nb2O5 electronic
structure by XPS, UPS and first principles methods. Surf Interface Anal 2013;45:1206–10. http://dx.doi.org/10.1002/
sia.5253.
[71] Posadas AB, O’Hara A, Rangan S, Bartynski RA, Demkov AA. Band gap of epitaxial in-plane-dimerized single-phase NbO2
films. Appl Phys Lett 2014;104:092901. http://dx.doi.org/10.1063/1.4867085.
[72] Wang Y, Comes RB, Kittiwatanakul S, Wolf SA, Lu J. Epitaxial niobium dioxide thin films by reactive-biased target ion
beam deposition. J Vac Sci Technol A 2015;33:021516. http://dx.doi.org/10.1116/1.4906143.
[73] Kim S, Park J, Woo J, Cho C, Lee W, Shin J, et al. Threshold-switching characteristics of a nanothin-NbO2-layer-based Pt/
NbO2/Pt stack for use in cross-point-type resistive memories. Microelectron Eng 2013;107:33–6. http://dx.doi.org/
10.1016/j.mee.2013.02.084.
[74] Liu X, Nandi SK, Venkatachalam DK, Belay K, Song S, Elliman RG. Reduced threshold current in NbO2 selector by
engineering device structure. IEEE Electron Device Lett 2014;35:1055–7. http://dx.doi.org/10.1109/LED.2014.2344105.
[75] Joshi T, Senty TR, Borisov P, Bristow AD, Lederman D. Preparation, characterization, and electrical properties of epitaxial
NbO2 thin film lateral devices. J Phys Appl Phys 2015;48:335308. http://dx.doi.org/10.1088/0022-3727/48/33/335308.
[76] Zhang J, Norris K, Samuels K, Ge N, Zhang M, Park J, et al. Electron energy-loss spectroscopy (EELS) study of NbOx film for
resistive memory applications. Microsc Microanal 2015;21:285–6. http://dx.doi.org/10.1017/S1431927615002226.
[77] Nandi SK, Liu X, Venkatachalam DK, Elliman RG. Threshold current reduction for the metal–insulator transition in
NbO2x-selector devices: the effect of ReRAM integration. J Phys Appl Phys 2015;48:195105. http://dx.doi.org/10.1088/
0022-3727/48/19/195105.
[78] Slesazeck S, Mähne H, Wylezich H, Wachowiak A, Radhakrishnan J, Ascoli A, et al. Physical model of threshold switching
in NbO2 based memristors. RSC Adv 2015;5:102318–22. http://dx.doi.org/10.1039/C5RA19300A.
[79] Helali S, Abdelghani A, Hafaiedh I, Martelet C, Prodromidis MI, Albanis T, et al. Functionalization of niobium electrodes for
the construction of impedimetric biosensors. Mater Sci Eng, C 2008;28:826–30. http://dx.doi.org/10.1016/j.
msec.2007.10.078.
[80] Lee S, Yoon H, Yoon I, Kim B. Single crystalline NbO2 nanowire synthesis by chemical vapor transport method. Bull Korean
Chem Soc 2012;33:839.
[81] Sasaki K, Maier J. Re-analysis of defect equilibria and transport parameters in Y2O3-stabilized ZrO2 using EPR and optical
relaxation. Solid State Ionics 2000;134:303–21. http://dx.doi.org/10.1016/S0167-2738(00)00766-9.
[82] Reznichenko LA, Akhnazarova VV, Shilkina LA, Razumovskaya ON, Dudkina SI. Invar effect in n-Nb2O5, aht-Nb2O5, and L-
Nb2O5. Crystallogr Rep 2009;54:483–91. http://dx.doi.org/10.1134/S1063774509030183.
[83] Valencia-Balvín C, Pérez-Walton S, Dalpian GM, Osorio-Guillén JM. First-principles equation of state and phase stability of
niobium pentoxide. Comput Mater Sci 2014;81:133–40. http://dx.doi.org/10.1016/j.commatsci.2013.07.032.
[84] Graça MPF, Meireles A, Nico C, Valente MA. Nb2O5 nanosize powders prepared by sol–gel – structure, morphology and
dielectric properties. J Alloys Compd 2013;553:177–82. http://dx.doi.org/10.1016/j.jallcom.2012.11.128.
[85] Rani RA, Zoolfakar AS, O’Mullane AP, Austin MW, Kalantar-Zadeh K. Thin films and nanostructures of niobium pentoxide:
fundamental properties, synthesis methods and applications. J Mater Chem A 2014;2:15683. http://dx.doi.org/10.1039/
C4TA02561J.
[86] Viet AL, Reddy MV, Jose R, Chowdari BVR, Ramakrishna S. Nanostructured Nb2O5 polymorphs by electrospinning for
rechargeable lithium batteries. J Phys Chem C 2010;114:664–71. http://dx.doi.org/10.1021/jp9088589.
[87] Le Viet A, Jose R, Reddy MV, Chowdari BVR, Ramakrishna S. Nb2O5 photoelectrodes for dye-sensitized solar cells: choice of
the polymorph. J Phys Chem C 2010;114:21795–800. http://dx.doi.org/10.1021/jp106515k.
[88] Kumagai N, Koishikawa Y, Komaba S, Koshiba N. Thermodynamics and kinetics of lithium intercalation into Nb2O5
electrodes for a 2 V rechargeable lithium battery. J Electrochem Soc 1999;146:3203–10. http://dx.doi.org/10.1149/
1.1392455.
[89] Brayner R, Bozon-Verduraz F. Niobium pentoxide prepared by soft chemical routes: morphology, structure, defects and
quantum size effect. Phys Chem Chem Phys 2003;5:1457–66.
[90] Gruehn R. Eine weitere neue Modifikation des niobpentoxids. J Common Met 1966;11:119–26. http://dx.doi.org/10.1016/
0022-5088(66)90076-2.
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 33

[91] Hibst H, Gruehn R. Weitere neue Nb2O5-Modifikationen, metastabile Oxidationsprodukte von NbOx-Phasen (2,4 < x < 2,5).
Z Für Anorg Allg Chem 1978;442:49–78.
[92] Meyer B, Gruehn R. Investigation of metastable niobium oxides by transmission electron microscopy. Comments Inorg
Chem 1982;1:361–77. http://dx.doi.org/10.1080/02603598208078104.
[93] Stoneham AM, Durham PJ. The ordering of crystallographic shear planes: theory of regular arrays. J Phys Chem Solids
1973;34:2127–35. http://dx.doi.org/10.1016/S0022-3697(73)80060-5.
[94] Mrowec S. On the defect structure in nonstoichiometric metal oxides. Ceramurg Int 1978;4:47–58. http://dx.doi.org/
10.1016/0390-5519(78)90119-9.
[95] Anderson JS, Browne JM, Cheetham AK, Dreele RV, Hutchison JL, Lincoln FJ, et al. Point defects and extended defects in
niobium oxides. Nature 1973;243:81–3. http://dx.doi.org/10.1038/243081a0.
[96] Iijima S, Kimura S, Goto M. High-resolution microscopy of nonstoichiometric Nb22O54 crystals: point defects and
structural defects. Acta Crystallogr Sect A 1974;30:251–7. http://dx.doi.org/10.1107/S0567739474000519.
[97] Iijima S, Kimura S, Goto M. Direct observation of point defects in Nb12O29 by high-resolution electron microscopy. Acta
Crystallogr Sect A 1973;29:632–6. http://dx.doi.org/10.1107/S0567739473001610.
[98] Dhawan S, Dhawan T, Vedeshwar AG. Growth of Nb2O5 quantum dots by physical vapor deposition. Mater Lett
2014;126:32–5. http://dx.doi.org/10.1016/j.matlet.2014.03.107.
[99] Tamura S. High-pressure phase research on Nb2O5. J Mater Sci 1972;7:298–302. http://dx.doi.org/10.1007/BF00555631.
[100] Madelung O, Rössler U, Schulz M, editors. Non-tetrahedrally bonded binary compounds II, vol. 41D. Berlin/
Heidelberg: Springer-Verlag; 2000.
[101] Zibrov IP, Filonenko VP, Werner P-E, Marinder B-O, Sundberg M. A new high-pressure modification of Nb2O5. J Solid State
Chem 1998;141:205–11. http://dx.doi.org/10.1006/jssc.1998.7954.
[102] Jehng JM, Wachs IE. Structural chemistry and Raman spectra of niobium oxides. Chem Mater 1991;3:100–7. http://dx.doi.
org/10.1021/cm00013a025.
[103] Nowak I, Ziolek M. Niobium compounds: preparation, characterization, and application in heterogeneous catalysis. Chem
Rev 1999;99:3603–24. http://dx.doi.org/10.1021/cr9800208.
[104] de Sá AI, Rangel CM, Skeldon P. Semiconductive properties of anodic niobium oxides. Port Electrochim Acta
2006;24:305–11.
[105] de Filho DAB, Franco DW, Filho PPA, Alves OL. Niobia films: surface morphology, surface analysis, photoelectrochemical
properties and crystallization process. J Mater Sci 1998;33:2607–16. http://dx.doi.org/10.1023/A:1004361420740.
[106] Schultze JW, Lohrengel MM. Stability, reactivity and breakdown of passive films. Problems of recent and future research.
Electrochim Acta 2000;45:2499–513. http://dx.doi.org/10.1016/S0013-4686(00)00347-9.
[107] Kovendhan M, Joseph DP, Manimuthu P, Ganesan S, Sambasivam S, Maruthamuthu P, et al. Spray deposited Nb2O5 thin
film electrodes for fabrication of dye sensitized solar cells. Trans Indian Inst Met 2011;64:185–8. http://dx.doi.org/
10.1007/s12666-011-0036-2.
[108] Marucco JF. Thermodynamic study of the system NbO2–Nb2O5 at high temperature. J Solid State Chem 1974;10:211–8.
http://dx.doi.org/10.1016/0022-4596(74)90028-0.
[109] Naito K, Kamegashira N, Sasaki N. Phase equilibria in the system between NbO2 and Nb2O5 at high temperatures. J Solid
State Chem 1980;35:305–11. http://dx.doi.org/10.1016/0022-4596(80)90526-5.
[110] Fischer V, Stormer H, Gerthsen D, Stenzel M, Zillgen H, Ivers-Tiffee E. Niobium as new material for electrolyte capacitors
with nanoscale dielectric oxide layers. In: Proc 7th int conf prop appl dielectr mater, vol. 3, ICPADM 2003, Nagoya, Japan;
2003. p. 1134–7.
[111] Cavigliasso GE, Esplandiu MJ, Macagno VA. Influence of the forming electrolyte on the electrical properties of tantalum
and niobium oxide films: an EIS comparative study. J Appl Electrochem 1998;28:1213–9.
[112] Macek M, Orel B. Electrochromism of sol–gel derived niobium oxide films. Sol Energy Mater Sol Cells 1998;54:121–30.
http://dx.doi.org/10.1016/S0927-0248(98)00062-2.
[113] Störmer H, Weber A, Fischer V, Ivers-Tiffée E, Gerthsen D. Anodically formed oxide films on niobium: microstructural and
electrical properties. J Eur Ceram Soc 2009;29:1743–53. http://dx.doi.org/10.1016/j.jeurceramsoc.2008.10.019.
[114] Zednicek T, Vrana B. Tantalum and niobium technology roadmap. AVX Corp; n.d.
[115] Jouve G. Electrical conduction mechanisms in electrochemically formed amorphous films of Nb2O5. Philos Mag Part B
1991;64:207–18. http://dx.doi.org/10.1080/13642819108207614.
[116] El-Shazly TS, Hassan WMI, Abdel Rahim ST, Allam NK. Unravelling the interplay of dopant concentration and band
structure engineering of monoclinic niobium pentoxide: a model photoanode for water splitting. Int J Hydrogen Energy
2015;40:13867–75. http://dx.doi.org/10.1016/j.ijhydene.2015.08.05.
[117] Kukli K, Ritala M, Leskelä M. Development of dielectric properties of niobium oxide, tantalum oxide, and aluminum oxide
based nanolayered materials. J Electrochem Soc 2001;148:F35–41. http://dx.doi.org/10.1149/1.1343106.
[118] Aegerter MA. Sol–gel niobium pentoxide: a promising material for electrochromic coatings, batteries, nanocrystalline
solar cells and catalysis. Sol Energy Mater Sol Cells 2001;68:401–22.
[119] Maruyama T, Arai S. Electrochromic properties of niobium oxide thin films prepared by radio-frequency magnetron
sputtering method. Appl Phys Lett 1993;63:869–70.
[120] Özer N, Rubin MD, Lampert CM. Optical and electrochemical characteristics of niobium oxide films prepared by sol–gel
process and magnetron sputtering: a comparison. Sol Energy Mater Sol Cells 1996;40:285–96. http://dx.doi.org/10.1016/
0927-0248(95)00147-6.
[121] Pawlicka A, Atik M, Aegerter MA. Synthesis of multicolor Nb2O5 coatings for electrochromic devices. Thin Solid Films
1997;301:236–41. http://dx.doi.org/10.1016/S0040-6090(96)09583-1.
[122] Costa ED, Avellaneda CO, Pawlicka A. Alternative Nb2O5–TiO2 thin films for electrochromic devices. J Mater Sci
2001;36:1407–10. http://dx.doi.org/10.1023/A:1017576125092.
[123] Huang Y, Zhang Y, Hu X. Structural, morphological and electrochromic properties of Nb2O5 films deposited by reactive
sputtering. Sol Energy Mater Sol Cells 2003;77:155–62. http://dx.doi.org/10.1016/S0927-0248(02)00318-5.
[124] Pehlivan E, Tepehan FZ, Tepehan GG. Comparison of optical, structural and electrochromic properties of undoped and
WO3-doped Nb2O5 thin films. Solid State Ionics 2003;165:105–10. http://dx.doi.org/10.1016/j.ssi.2003.08.021.
34 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

[125] Romero R, Dalchiele EA, Martín F, Leinen D, Ramos-Barrado JR. Electrochromic behaviour of Nb2O5 thin films with
different morphologies obtained by spray pyrolysis. Sol Energy Mater Sol Cells 2009;93:222–9. http://dx.doi.org/10.1016/
j.solmat.2008.10.012.
[126] Abdul Rani R, Zoolfakar AS, Subbiah J, Ou JZ, Kalantar-zadeh K. Highly ordered anodized Nb2O5 nanochannels for dye-
sensitized solar cells. Electrochem Commun; n.d. http://dx.doi.org/10.1016/j.elecom.2013.12.011.
[127] Barea E, Xu X, González-Pedro V, Ripollés-Sanchis T, Fabregat-Santiago F, Bisquert J. Origin of efficiency enhancement in
Nb2O5 coated titanium dioxide nanorod based dye sensitized solar cells. Energy Environ Sci 2011;4:3414–9. http://dx.doi.
org/10.1039/C1EE01193F.
[128] Jose R, Thavasi V, Ramakrishna S. Metal oxides for dye-sensitized solar cells. J Am Ceram Soc 2009;92:289–301. http://dx.
doi.org/10.1111/j.1551-2916.2008.02870.x.
[129] Cui H, Zhu G, Xie Y, Zhao W, Yang C, Lin T, et al. Black nanostructured Nb2O5 with improved solar absorption and
enhanced photoelectrochemical water splitting. J Mater Chem A 2015;3:11830–7. http://dx.doi.org/10.1039/
C5TA01544H.
[130] Tanabe K. Catalytic application of niobium compounds. Catal Today 2003;78:65–77.
[131] Fan M-Q, Sun L-X, Zhang Y, Xu F, Zhang J, Chu H. The catalytic effect of additive Nb2O5 on the reversible hydrogen storage
performances of LiBH4–MgH2 composite. Int J Hydrogen Energy 2008;33:74–80.
[132] Kurioka N, Watanabe D, Haneda M, Shimanouchi T, Mizushima T, Kakuta N, et al. Preparation of niobium oxide films as a
humidity sensor. Catal Today 1993;16:495–501. http://dx.doi.org/10.1016/0920-5861(93)80090-N.
[133] Choi J, Lim JH, Rho S, Jahng D, Lee J, Kim KJ. Nanoporous niobium oxide for label-free detection of DNA hybridization
events. Talanta 2008;74:1056–9. http://dx.doi.org/10.1016/j.talanta.2007.07.007.
[134] Eisenbarth E, Velten D, Müller M, Thull R, Breme J. Nanostructured niobium oxide coatings influence osteoblast adhesion.
J Biomed Mater Res A 2006;79:166–75.
[135] Velten D, Eisenbarth E, Schanne N, Breme J. Biocompatible Nb2O5 thin films prepared by means of the sol–gel process. J
Mater Sci – Mater Med 2004;15:457–61.
[136] Ramírez G, Rodil SE, Arzate H, Muhl S, Olaya JJ. Niobium based coatings for dental implants. Appl Surf Sci
2011;257:2555–9.
[137] Cardinal T, Fargin E, Le Flem G, Leboiteux S. Correlations between structural properties of Nb2O5–NaPO3–Na2B4O7 glasses
and non-linear optical activities. J Non-Cryst Solids 1997;222:228–34.
[138] Praveena R, Venkatramu V, Babu P, Jayasankar CK. Fluorescence spectroscopy of Sm3+ ions in P2O5–PbO–Nb2O5 glasses.
Phys B Condens Matter 2008;403:3527–34.
[139] Sanghi S, Rani S, Agarwal A, Bhatnagar V. Influence of Nb2O5 on the structure, optical and electrical properties of alkaline
borate glasses. Mater Chem Phys 2010;120:381–6. http://dx.doi.org/10.1016/j.matchemphys.2009.11.016.
[140] Graça MPF, Valente MA, Ferreira Da Silva MG. The electric behavior of a lithium-niobate-phosphate glass and glass-
ceramics. J Mater Sci 2006;41:1137–44. http://dx.doi.org/10.1007/s10853-005-3652-6.
[141] Lin H, Jiang S, Wu J, Song F, Peyghambarian N, Pun EYB. Er3+ doped Na2O–Nb2O5–TeO2 glasses for optical waveguide laser
and amplifier. J Phys Appl Phys 2003;36:812.
[142] Aquino FT, Ferrari JL, Ribeiro SJL, Ferrier A, Goldner P, Goncalves RR. Broadband NIR emission in novel sol–gel Er3+-doped
SiO2–Nb2O5 glass ceramic planar waveguides for photonic applications. Opt Mater 2013;35:387–96.
[143] Eckert J, Reichert K, Schnitter C, Seyeda H. Processing of columbite-tantalite ores and concentrates for niobium and
niobium compounds in electronic applications. In: Proc int symp niobium 2001, Niobium 2001 Ltd, Bridgeville, USA; n.d.
p. 67–87.
[144] Michael Scherer HH. Innovative production of thin film laser components 2005; 5963. http://dx.doi.org/10.1117/12.
625226.
[145] Dhar A, Alford TL. Optimization of Nb2O5/Ag/Nb2O5 multilayers as transparent composite electrode on flexible substrate
with high figure of merit. J Appl Phys 2012;112:103113. http://dx.doi.org/10.1063/1.4767662.
[146] Vinnichenko M, Rogozin A, Grambole D, Munnik F, Kolitsch A, Möller W, et al. Highly dense amorphous Nb2O5 films with
closed nanosized pores. Appl Phys Lett 2009;95:081904. http://dx.doi.org/10.1063/1.3212731.
[147] Hunsche B, Vergöhl M, Neuhäuser H, Klose F, Szyszka B, Matthée T. Effect of deposition parameters on optical and
mechanical properties of MF- and DC-sputtered Nb2O5 films. Thin Solid Films 2001;392:184–90. http://dx.doi.org/
10.1016/S0040-6090(01)01025-2.
[148] Chua LO. Memristor – the missing circuit element. IEEE Trans Circuit Theory 1971;18:507–19. http://dx.doi.org/10.1109/
TCT.1971.1083337.
[149] Strukov DB, Snider GS, Stewart DR, Williams RS. The missing memristor found. Nature 2008;453:80–3. http://dx.doi.org/
10.1038/nature06932.
[150] Yang JJ, Strukov DB, Stewart DR. Memristive devices for computing. Nat Nanotechnol 2013;8:13–24. http://dx.doi.org/
10.1038/nnano.2012.240.
[151] Jo SH, Chang T, Ebong I, Bhadviya BB, Mazumder P, Lu W. Nanoscale memristor device as synapse in neuromorphic
systems. Nano Lett 2010;10:1297–301. http://dx.doi.org/10.1021/nl904092h.
[152] Liu X, Sadaf SM, Son M, Park J, Shin J, Lee W, et al. Co-occurrence of threshold switching and memory switching in cells for
crosspoint memory applications. Electron Device Lett IEEE 2012;33:236–8.
[153] Linn E, Rosezin R, Kügeler C, Waser R. Complementary resistive switches for passive nanocrossbar memories. Nat Mater
2010;9:403–6.
[154] Rosezin R, Linn E, Nielen L, Kugeler C, Bruchhaus R, Waser R. Integrated complementary resistive switches for passive
high-density nanocrossbar arrays. Electron Device Lett IEEE 2011;32:191–3.
[155] Sawa A. Resistive switching in transition metal oxides. Mater Today 2008;11:28–36. http://dx.doi.org/10.1016/S1369-
7021(08)70119-6.
[156] Yang JJ, Zhang M-X, Strachan JP, Miao F, Pickett MD, Kelley RD, et al. High switching endurance in TaOx memristive
devices. Appl Phys Lett 2010;97:232102.
[157] Lee M-J, Lee CB, Lee D, Lee SR, Chang M, Hur JH, et al. A fast, high-endurance and scalable non-volatile memory device
made from asymmetric Ta2O5x/TaO2x bilayer structures. Nat Mater 2011;10:625–30.
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 35

[158] Chen Y-S, Lee H-Y, Chen P-S, Chen W-S, Tsai K-H, Gu P-Y, et al. Novel defects-trapping TaOX/HfOX RRAM with reliable
self-compliance, high nonlinearity, and ultra-low current; 2014.
[159] Liu X, Sadaf SM, Park S, Kim S, Cha E, Lee D, et al. Complementary resistive switching in niobium oxide-based resistive
memory devices. Electron Device Lett IEEE 2013;34:235–7.
[160] Grishin AM, Velichko AA, Jalalian A. Nb2O5 nanofiber memristor. Appl Phys Lett 2013;103:053111. http://dx.doi.org/
10.1063/1.4817302.
[161] Sadaf SM, Liu X, Son M, Park S, Choudhury SH, Cha E, et al. Highly uniform and reliable resistance switching properties in
bilayer WOx/NbOx RRAM devices. Phys Status Solidi A 2012;209:1179–83.
[162] Kundozerova TV, Grishin AM, Stefanovich GB, Velichko AA. Anodic Nb2O5 nonvolatile RRAM. IEEE Trans Electron Devices
2012;59:1144–8. http://dx.doi.org/10.1109/TED.2011.2182515.
[163] Liu X, Sadaf SM, Son M, Shin J, Park J, Lee J, et al. Diode-less bilayer oxide (WOx–NbOx) device for cross-point resistive
memory applications. Nanotechnology 2011;22:475702.
[164] Jo Y, Sim H, Hwang H, Ahn K, Jung M-H. Reverse resistance switching in polycrystalline Nb2O5 films. Thin Solid Films
2010;518:5676–8. http://dx.doi.org/10.1016/j.tsf.2009.10.026.
[165] Sim Hyunjun, Choi Dooho, Lee Dongsoo, Seo Sunae, Lee Myong-Jae, Yoo In-Kyeong, et al. Resistance-switching
characteristics of polycrystalline Nb2O5 for nonvolatile memory application. IEEE Electron Device Lett 2005;26:292–4.
http://dx.doi.org/10.1109/LED.2005.846592.
[166] Mitin A, Shamray V, Gordeev A. Effects of resistance switching in niobium oxides: metastable nanochannels. Bull Russ
Acad Sci Phys 2009;73:125–8. http://dx.doi.org/10.3103/S106287380901033X.
[167] Baek H, Lee C, Choi J, Cho J. Nonvolatile memory devices prepared from sol–gel derived niobium pentoxide films.
Langmuir 2013;29:380–6. http://dx.doi.org/10.1021/la303857b.
[168] Roberts MW, Thomas JM, Anderson JS, Tilley RJD. Crystallographic shear and non-stoichiometry. In: Surf defect prop
solids. vol. 3; 2007.
[169] Iijima S. Direct observation of lattice defects in H-Nb2O5 by high resolution electron microscopy. Acta Crystallogr Sect A
1973;29:18–24. http://dx.doi.org/10.1107/S0567739473000045.
[170] Niebuhr J. Die niederen Oxide des Niobs. J Common Met 1966;11:191–203.
[171] Kimura S. Phase equilibria in the system NbO2–Nb2O5: phase relations at 1300 and 1400 °C and related thermodynamic
treatment. J Solid State Chem 1973;6:438–49.
[172] Marucco JF. Electrical resistance and defect structure of stable and metastable phases of the system Nb12O29–Nb2O5
between 800 and 1100 °C. J Chem Phys 1979;70:649–54. http://dx.doi.org/10.1063/1.437545.
[173] McQueen T, Xu Q, Andersen EN, Zandbergen HW, Cava RJ. Structures of the reduced niobium oxides Nb12O29 and Nb22O54.
J Solid State Chem 2007;180:2864–70.
[174] Ohsawa T, Okubo J, Suzuki T, Kumigashira H, Oshima M, Hitosugi T. An n-type transparent conducting oxide: Nb12O29. J
Phys Chem C 2011;115:16625–9.
[175] Waldron JEL, Green MA, Neumann DA. Structure and electronic properties of monoclinic Nb12O29. J Phys Chem Solids
2004;65:79–86. http://dx.doi.org/10.1016/j.jpcs.2003.09.001.
[176] Janninck RF, Whitmore DH. Electrical conduction in nonstoichiometric a-Nb2O5. J Chem Phys 1962;37:2750–4. http://dx.
doi.org/10.1063/1.1733100.
[177] Cava RJ, Batlogg B, Krajewski JJ, Gammel P, Poulsen HF, Peck WF, et al. Antiferromagnetism and metallic conductivity in
Nb12O29. Nature 1991;350:598–600. http://dx.doi.org/10.1038/350598a0.
[178] Waldron JEL, Green MA, Neumann DA. Charge and spin ordering in monoclinic Nb12O29. J Am Chem Soc
2001;123:5833–4. http://dx.doi.org/10.1021/ja015506b.
[179] Xu J, Jarlborg T, Freeman A. Self-consistent band structure of the rutile dioxides NbO2, RuO2, and IrO2. Phys Rev B:
Condens Matter 1989;40:7939–47.
[180] Ikeya T, Senna M. Change in the structure of niobium pentoxide due to mechanical and thermal treatments. J Non-Cryst
Solids 1988;105:243–50. http://dx.doi.org/10.1016/0022-3093(88)90313-4.
[181] McConnell AA, Aderson JS, Rao CNR. Raman spectra of niobium oxides. Spectrochim Acta Part Mol Spectrosc
1976;32:1067–76. http://dx.doi.org/10.1016/0584-8539(76)80291-7.
[182] Huang BX, Wang K, Church JS, Li Y-S. Characterization of oxides on niobium by Raman and infrared spectroscopy.
Electrochim Acta 1999;44:2571–7. http://dx.doi.org/10.1016/S0013-4686(98)00385-5.
[183] Balachandran U, Eror NG. Raman spectrum of the high temperature form of Nb2O5. J Mater Sci Lett 1982;1:374–6. http://
dx.doi.org/10.1007/BF00724842.
[184] Adler D. Mechanisms for metal–nonmental transitions in transition-metal oxides and sulfides. Rev Mod Phys
1968;40:714–36. http://dx.doi.org/10.1103/RevModPhys.40.714.
[185] Lee JC, Durand WW. Electrically stimulated optical switching of NbO2 thin films. J Appl Phys 1984;56:3350–2. http://dx.
doi.org/10.1063/1.333863.
[186] Sakai Y, Tsuda N, Sakata T. Electrical properties of semiconducting NbO2. J Phys Soc Jpn 1985;54:1514–8. http://dx.doi.
org/10.1143/JPSJ.54.1514.
[187] Manzo M, Laurell F, Pasiskevicius V, Gallo K. Lithium niobate: the silicon of photonics! In: Bartolo BD, Collins J, editors.
Nano-opt enhancing light-matter interact mol scale. Netherlands: Springer; 2013. p. 421–2.
[188] Wang K, Li J-F. (K, Na)NbO3-based lead-free piezoceramics: phase transition, sintering and property enhancement. J Adv
Ceram 2012;1:24–37. http://dx.doi.org/10.1007/s40145-012-0003-3.
[189] Thierfelder C, Sanna S, Schindlmayr A, Schmidt WG. Do we know the band gap of lithium niobate? Phys Status Solidi C
2010;7:362–5. http://dx.doi.org/10.1002/pssc.200982473.
[190] Rachel, Dutto F, Radenovic A. Niobates nanowires: synthesis, characterization and applications. In: Hashim A, editor.
Nanowires – implement. Appl. InTech; 2011.
[191] Graça MPF, Valente MA, Peres M, Cruz A, Soares MJ, Neves AJ, et al. Structural and optical properties of Er3+ ion in sol–gel
grown LiNbO3. J Phys: Condens Matter 2007;19:016213. http://dx.doi.org/10.1088/0953-8984/19/1/016213.
[192] Nico C, Graça MPF, Monteiro T, Valente MA. Lithium niobosilicate glass doped with samarium: dielectric and optical
analysis. Phys Status Solidi A 2011;208:2288–92. http://dx.doi.org/10.1002/pssa.201000764.
36 C. Nico et al. / Progress in Materials Science 80 (2016) 1–37

[193] Fang J, Wang X, Tian Z, Zhong C, Li L, Zuo R. Two-step sintering: an approach to broaden the sintering temperature range
of alkaline niobate-based lead-free piezoceramics. J Am Ceram Soc 2010;93:3552–5. http://dx.doi.org/10.1111/j.1551-
2916.2010.04085.x.
[194] Kuznetsova E, Reznichenko L, Razumovskaya O, Shilkina L. The polymorphism of niobium pentoxide and the properties of
alkali metal niobates for ferroelectric piezoceramics. Tech Phys Lett 2001;27:189–92. http://dx.doi.org/10.1134/
1.1359821.
[195] Hreščak J, Bencan A, Rojac T, Malič B. The influence of different niobium pentoxide precursors on the solid-state synthesis
of potassium sodium niobate. J Eur Ceram Soc 2013;33:3065–75. http://dx.doi.org/10.1016/j.jeurceramsoc.2013.07.006.
[196] Yan C, Nikolova L, Dadvand A, Harnagea C, Sarkissian A, Perepichka DF, et al. Multiple NaNbO3/Nb2O5 heterostructure
nanotubes: a new class of ferroelectric/semiconductor nanomaterials. Adv Mater 2010;22:1741–5. http://dx.doi.org/
10.1002/adma.200903589.
[197] Pullar RC. The synthesis, properties, and applications of columbite niobates (M2+Nb2O6): a critical review. J Am Ceram Soc
2009;92:563–77. http://dx.doi.org/10.1111/j.1551-2916.2008.02919.x.
[198] Cho I-S, Bae ST, Kim DH, Hong KS. Effects of crystal and electronic structures of ANb2O6 (A = Ca, Sr, Ba) metaniobate
compounds on their photocatalytic H2 evolution from pure water. Int J Hydrogen Energy 2010;35:12954–60. http://dx.
doi.org/10.1016/j.ijhydene.2010.04.057.
[199] Hsiao Y-J, Liu C-W, Dai B-T, Chang Y-H. Sol–gel synthesis and the luminescent properties of CaNb2O6 phosphor powders. J
Alloys Compd 2009;475:698–701. http://dx.doi.org/10.1016/j.jallcom.2008.07.142.
[200] Cho I-S, Bae ST, Yim DK, Kim DW, Hong KS. Preparation, characterization, and photocatalytic properties of CaNb2O6
nanoparticles. J Am Ceram Soc 2009;92:506–10. http://dx.doi.org/10.1111/j.1551-2916.2008.02894.x.
[201] Arroyo y de Dompablo ME, Lee Y-L, Morgan D. First principles investigation of oxygen vacancies in columbite MNb2O6 (M
= Mn, Fe Co, Ni, Cu). Chem Mater 2010;22:906–13. http://dx.doi.org/10.1021/cm901723.
[202] Thornton JR, Fountain WD, Flint GW, Crow TG. Properties of neodymium laser materials. Appl Opt 1969;8:1087–102.
http://dx.doi.org/10.1364/AO.8.001087.
[203] Ballman AA, Porto SPS, Yariv A. Calcium niobate Ca(NbO3)2—a new laser host crystal. J Appl Phys 1963;34:3155–6. http://
dx.doi.org/10.1063/1.1729154.
[204] Hsiao Y-J, Chang Y-S, Chen G-J, Chang Y-H. Synthesis and the luminescent properties of CdNb2O6 oxides by sol–gel
process. J Alloys Compd 2009;471:259–62. http://dx.doi.org/10.1016/j.jallcom.2008.03.081.
[205] Hsiao Y-J, Fang T-H, Ji L-W, Chi S-S. Surface and photoluminescence characteristics of CdNb2O6 nanocrystals. Open Surf
Sci J 2009;1:30–3. http://dx.doi.org/10.2174/1876531900901010030.
[206] Qin L, Wei D, Huang Y, Kim SI, Yu YM, Seo HJ. Triple-layered perovskite niobates CaRNb3O10 (R = La, Sm, Eu, Gd, Dy, Er, Yb,
or Y): new self-activated oxides. Inorg Chem 2013;52:10407–13. http://dx.doi.org/10.1021/ic401854r.
[207] Rooksby HP, White EAD, Langston SA. Perovskite-type rare-earth niobates and tantalates. J Am Ceram Soc
1965;48:447–9.
[208] Rooksby HP, White EAD. Rare-earth niobates and tantalates of defect fluorite-and weberite-type structures. J Am Ceram
Soc 1964;47:94–6.
[209] Rooksby HP, White EAD. The structures of 1:1 compounds of rare earth oxides with niobia and tantala. Acta Crystallogr
1963;16:888–90.
[210] Cashion JD, Cooke AH, Leask MJM, Thorp TL, Wells MR. Crystal growth and magnetic susceptibility of some rare-earth
compounds. J Mater Sci 1968;3:402–7. http://dx.doi.org/10.1007/BF00550984.
[211] Garton G, Wanklyn BM. Crystal growth and magnetic susceptibility of some rare-earth compounds. J Mater Sci
1968;3:395–401. http://dx.doi.org/10.1007/BF00550983.
[212] Siqueira KPF, Moreira RL, Dias A. Synthesis and crystal structure of lanthanide orthoniobates studied by vibrational
spectroscopy. Chem Mater 2010;22:2668–74.
[213] Kim D-W, Kwon D-K, Yoon SH, Hong KS. Microwave dielectric properties of rare-earth ortho-niobates with ferroelasticity.
J Am Ceram Soc 2006;89:3861–4. http://dx.doi.org/10.1111/j.1551-2916.2006.01302.x.
[214] Siqueira KPF, Carvalho GB, Dias A. Influence of the processing conditions and chemical environment on the crystal
structures and phonon modes of lanthanide orthotantalates. Dalton Trans 2011;40:9454–60. http://dx.doi.org/10.1039/
C1DT10783F.
[215] Haugsrud R, Norby T. Proton conduction in rare-earth ortho-niobates and ortho-tantalates. Nat Mater 2006;5:193–6.
http://dx.doi.org/10.1038/nmat1591.
[216] Nyman M, Rodriguez MA, Rohwer LES, Martin JE, Waller M, Osterloh FE. Unique LaTaO4 polymorph for multiple energy
applications. Chem Mater 2009;21:4731–7. http://dx.doi.org/10.1021/cm9020645.
[217] Octaviano ES, Reyes Ardila D, Andrade LHC, Siu Li M, Andreeta JP. Growth and evaluation of lanthanoids orthoniobates
single crystals processed by a miniature pedestal growth technique. Cryst Res Technol 2004;39:859–63.
[218] Graça MPF, Peixoto M, Ferreira N, Rodrigues J, Nico C, Costa F, et al. Optical and dielectric behaviour of EuNbO4 crystals. J
Mater Chem C 2013;1:2913–9.
[219] Xiao X, Yan B. Synthesis and luminescent properties of novel RENbO4:Ln3+ (RE = Y, Gd, Lu; Ln = Eu, Tb) micro-crystalline
phosphors. J Non-Cryst Solids 2005;351:3634–9. http://dx.doi.org/10.1016/j.jnoncrysol.2005.09.018.
[220] Nyman MD, Rohwer LES. Rare-earth tantalates and niobates suitable for use as nanophosphors. US8585928 B1; 2013.
[221] Kato K, Toyoura K, Nakamura A, Matsunaga K. First-principles analysis on proton diffusivity in La3NbO7. Solid State Ionics
2013;262:472–5.
[222] Ivanova M, Ricote S, Meulenberg WA, Haugsrud R, Ziegner M. Effects of A-and B-site (co-) acceptor doping on the
structure and proton conductivity of LaNbO4. Solid State Ionics 2012;213:45–52.
[223] Kepaptsoglou D-M, Hadidi K, Løvvik O-M, Magraso A, Norby T, Gunnaes AE, et al. Interfacial charge transfer and chemical
bonding in a Ni–LaNbO4 cermet for proton-conducting solid-oxide fuel cell anodes. Chem Mater 2012;24:4152–9.
[224] Amsif M, Marrero-López D, Ruiz-Morales JC, Savvin S, Núñez P. Low temperature sintering of LaNbO4 proton conductors
from freeze-dried precursors. J Eur Ceram Soc 2012;32:1235–44.
[225] Mather GC, Fisher CAJ, Islam MS. Defects, dopants, and protons in LaNbO4. Chem Mater 2010;22:5912–7. http://dx.doi.
org/10.1021/cm1018822.
C. Nico et al. / Progress in Materials Science 80 (2016) 1–37 37

[226] Tolchard JR, Lein HL, Grande T. Chemical compatibility of proton conducting LaNbO4 electrolyte with potential oxide
cathodes. J Eur Ceram Soc 2009;29:2823–30. http://dx.doi.org/10.1016/j.jeurceramsoc.2009.03.034.
[227] Reijers R, Haije W. Literature review on high temperature proton conducting materials. Energy Res Cent Neth; 2008.
[228] Mokkelbost T, Kaus I, Haugsrud R, Norby T, Grande T, Einarsrud M-A. High-temperature proton-conducting lanthanum
ortho-niobate-based materials. Part II: sintering properties and solubility of alkaline earth oxides. J Am Ceram Soc
2008;91:879–86. http://dx.doi.org/10.1111/j.1551-2916.2007.02232.x.
[229] Mokkelbost T, Andersen Ø, Strøm RA, Wiik K, Grande T, Einarsrud M-A. High-temperature proton-conducting LaNbO4-
based materials: powder synthesis by spray pyrolysis. J Am Ceram Soc 2007;90:3395–400. http://dx.doi.org/10.1111/
j.1551-2916.2007.01904.x.
[230] Baughman RH, Stafström S, Cui C, Dantas SO. Materials with negative compressibilities in one or more dimensions.
Science 1998;279:1522–4.
[231] Pan L, Wang Y, Wang X, Qu H, Zhao J, Li Y, et al. Hydrogen photochromism in Nb2O5 powders. Phys Chem Chem Phys
2014. http://dx.doi.org/10.1039/C4CP02834A.
[232] Joshua Yang J, Miao F, Pickett MD, Ohlberg DAA, Stewart DR, Lau CN, et al. The mechanism of electroforming of metal
oxide memristive switches. Nanotechnology 2009;20:215201. http://dx.doi.org/10.1088/0957-4484/20/21/21520.
[233] Park S-J, Lee J-P, Jang JS, Rhu H, Yu H, You BY, et al. In situ control of oxygen vacancies in TiO2 by atomic layer deposition
for resistive switching devices. Nanotechnology 2013;24:295202. http://dx.doi.org/10.1088/0957-4484/24/29/295202.
[234] Waser R, Aono M. Nanoionics-based resistive switching memories. Nat Mater 2007;6:833–40.
[235] Messerschmitt F, Kubicek M, Schweiger S, Rupp JLM. Memristor kinetics and diffusion characteristics for mixed anionic-
electronic SrTiO3-d bits: the memristor-based cottrell analysis connecting material to device performance. Adv Funct
Mater 2014;24:7448–60. http://dx.doi.org/10.1002/adfm.201402286.
[236] Valov I. Redox-based resistive switching memories (ReRAMs): electrochemical systems at the atomic scale.
ChemElectroChem 2014;1:26–36. http://dx.doi.org/10.1002/celc.201300165.
[237] Lee S, Na H, Kim J, Moon J, Sohn H. Anion-migration-induced bipolar resistance switching in electrochemically deposited
TiOx films. J Electrochem Soc 2011;158:H88–92. http://dx.doi.org/10.1149/1.3516464.
[238] Waser R, Dittmann R, Staikov G, Szot K. Redox-based resistive switching memories – nanoionic mechanisms, prospects,
and challenges. Adv Mater 2009;21:2632–63.
[239] Hamwi S, Meyer J, Kröger M, Winkler T, Witte M, Riedl T, et al. The role of transition metal oxides in charge-generation
layers for stacked organic light-emitting diodes. Adv Funct Mater 2010;20:1762–6. http://dx.doi.org/10.1002/
adfm.201000301.
[240] Liang J, Zu F-S, Ding L, Xu M-F, Shi X-B, Wang Z-K, et al. Aqueous solution-processed MoO3 thick films as hole injection
and short-circuit barrier layer in large-area organic light-emitting devices. Appl Phys Exp 2014;7:111601. http://dx.doi.
org/10.7567/APEX.7.111601.

You might also like