Bi Doped NaNbO3

You might also like

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

Journal of

Materials Chemistry C
View Article Online
PAPER View Journal | View Issue
Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

Tunable phase transitions in NaNbO3 ceramics


through bismuth/vacancy modification†
Cite this: J. Mater. Chem. C, 2021,
9, 4289
Ludan Zhang,a Zhongna Yan,ab Tao Chen,ac Hang Luo, b Hangfeng Zhang,ad
Taslema Khanom,a Dou Zhang, b Isaac Abrahams *a and Haixue Yan *d

Sodium niobate, NaNbO3, which exhibits a perovskite structure, has recently stimulated interest in the
field of energy storage capacitors, with derived solid solutions shown to have promising energy storage
densities. Here A-site Bi/vacancy doping in NaNbO3 in the system Na13xBixV2xNbO3 (where V = vacancy
and x = 0.015, 0.05, 0.10, 0.15 and 0.20) is investigated. Phase evolution was systematically examined
through X-ray powder diffraction, Raman spectroscopy and thermal analysis, and is found to correlate
with changes in electrical behaviour. It is shown that through tuning the Bi/vacancy content, different
high temperature phases (above 550 1C) of NaNbO3, including the tetragonal P4/mbm and cubic Pm3m %
phases are stabilised at room temperature. The phase evolution from Pbcm (x = 0.015 and 0.05) to
Received 21st December 2020, P4/mbm (x = 0.10 and 0.15) to Pm3m % (x = 0.20) with increased Bi/vacancy content is accompanied by
Accepted 8th March 2021 lattice expansion, which is explained in terms of the accommodation of increased disorder resulting
DOI: 10.1039/d0tc05969b from the mixed arrangement of species (xBi3+, (1  x)Na+ and 2xVNa) on the A-site. Bismuth/vacancy
modification of NaNbO3 is seen to induce relaxor-like behaviour, significantly increasing both the
rsc.li/materials-c recoverable energy storage density and energy storage efficiency.

Introduction anti-ferroelectric phase can be stabilised at room temperature


through grain size control,19 while solid solutions based on
Ceramic dielectric capacitors are promising energy storage NaNbO3 have been shown to have promising energy storage
devices due to their advantages of broad working voltage, high densities.20,21
working temperature ranges, high power density and excellent NaNbO3 exhibits complex polymorphism with several distinct
cycling ability.1–3 Traditional lead-based ceramics such as perovskite-like phases identified. These include the ferroelectric
modified PbZrO3 and Pb(Mg1/3Nb2/3)O3 exhibit outstanding (FE) N-phase below 100 1C (R3c), antiferroelectric (AFE) phases,
power density due to their antiferroelectric or relaxor including the P-phase over the range 100 1C to 373 1C (Pbcm)
characteristics.4–7 However, environmental and toxicity concerns and the R-phase over the range 373–480 1C (Pmmn), as well as the
over the use of lead have resulted in much interest in lead-free non-polar paraelectric phases including the S-phase from 480 1C
functional ceramics for such applications.8–11 to 520 1C (Pmmn), the T1- phase (520–575 1C, Cmcm), the
NaNbO3 has potential applications in photocatalysis (for T2-phase (575–641 1C, P4/mbm) and U phase (4641 1C, Pm3% m).22–24
nano-sized NaNbO3 with appropriate band gap),12–15 hetero- The AFE P-phase is not easy to isolate at room temperature.
structure semiconductors16 and piezoelectric actuators and This is mainly because another phase, the FE Q-phase (P21ma)
nanogenerators.17,18 NaNbO3 has also stimulated interest in and the N-phase always coexist with the P-phase under
the field of energy storage, with recent work showing that the traditional calcination conditions due to their similar free
energies, resulting in electric displacement–electric field (D–E)
loops that are FE-like.24–26
a
School of Biological and Chemical Sciences, Queen Mary University of London,
The parameters and related functions to assess the energy
Mile End Road, London, E1 4NS, UK. E-mail: i.abrahams@qmul.ac.uk
b
State Key Laboratory of Powder Metallurgy, Central South University, Changsha,
storage density and efficiency of ceramic materials can be
Hunan 410083, China described by eqn (1)–(3).27,28
c
Key Laboratory of Inorganic Functional Materials and Devices, ð Ds
Shanghai Institute of Ceramics, Chinese Academy of Sciences, 1295 Dingxi Road, Wtotal ¼ EdD ðduring changeÞ (1)
Shanghai 200050, China 0
d
School of Engineering and Materials Science, Queen Mary University of London,
Mile End Road, London, E1 4NS, UK. E-mail: h.x.yan@qmul.ac.uk ð Ds
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ Wrec ¼ EdD ðduring dischargeÞ (2)
d0tc05969b Dr

This journal is © The Royal Society of Chemistry 2021 J. Mater. Chem. C, 2021, 9, 4289–4299 | 4289
View Article Online

Paper Journal of Materials Chemistry C

Wrec materials show sudden changes in strain induced by the


Z¼ (3)
Wtotal external electric field.38,39 This can result in cracking of the
material and represents a potential risk for actual storage
where Wtotal, Wrec and Z represent the total energy storage applications. While RFE materials also show field induced
density during charging, the recoverable energy storage density strain, the effect is more gradual.
during discharging and the energy storage efficiency, respectively.
E, D, Dr and Ds represent the applied electric field, field-induced
Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

dielectric displacement, dielectric displacement at zero field and Experimental


dielectric displacement at the highest applied field, respectively.
Ideally, to boost Wrec and Z simultaneously with low energy Polycrystalline Na13xBixNbO3 (x = 0.015, 0.05, 0.10, 0.15 and
storage loss (Wloss = Wtotal  Wrec), the difference between Dr 0.20) ceramics were obtained by a conventional solid-state
and Ds should be increased, along with an increase in the reaction method. All starting materials were purchased from
electrical field breakdown strength, Eb. Sigma-Aldrich. Stoichiometric amounts of Na2CO3 (99.5%,
In recent years, efforts have been made to realise the dried at 200 1C for 24 h before use), Bi2O3 (99.9%) and Nb2O5
potential of NaNbO3 in energy storage applications, for example (99.9%) were ground in anhydrous ethanol with zirconium
by trying to facilitate the reversible AFE phase transition when oxide balls for 4 h using a planetary ball mill at 300 rpm in a
removing the electric field to increase the efficiency,29 although nylon jar. After drying on a hot plate to evaporate the ethanol,
this approach is difficult to realise. Solid solution formation the blended powders were sieved using a 250 mm sieve.
with CaZrO3 was reported to stabilise antiferroelectric behaviour, The powders were then calcined at 900 1C for 3 h in an alumina
with an energy storage density of 0.55 J cm3 at 13 kV mm1 crucible. The calcined powders were re-milled for 4 h in
realised in 0.96NaNbO3–0.04CaZrO3.30 Another approach is to anhydrous ethanol. After drying and sieving, powders were
target relaxor-ferroelectric phases to achieve high energy blended with 5 wt% PVA aqueous solution and were pressed
densities.31 Yang et al.32 reported that 0.77NaNbO3–0.23BaTiO3 into pellets of 15 mm diameter and 1.0–1.5 mm thickness
relaxor ceramics can exhibit an energy storage density of under a pressure of 150 MPa. The pellets were then covered
1.5 J cm3 at 17.5 kV mm1, while Fan et al.33 reported that with powder of the same composition and heated at 800 1C for
0.91NaNbO3–0.09Bi(Zn0.5Ti0.5)O3 relaxor ceramics exhibited a 2 h to burn off the PVA. Samples were then heated to a
recoverable energy storage density of 2.2 J cm3 at 25 kV mm1 temperature in the range 1150 to 1335 1C (1150, 1165, 1250,
and Qi et al.34 reported that 0.76NaNbO3–0.24(Bi0.5Na0.5)TiO3 1275 and 1320 1C for x = 0.20, 0.15, 0.10, 0.05 and 0.015,
relaxor ceramics exhibited a recoverable energy storage density respectively) at a rate of 5 1C min1 and sintered at this
of 12.2 J cm3 at 68 kV mm1 due to the combined contributions temperature for 3 h before cooling to room temperature.
from the AFE and relaxor behaviours. In addition, even higher The density of the ceramic pellets was measured based on
energy storage densities of 13.5 J cm3 at 400 kV mm1 have been the classical Archimedes method by displacement in water.
reported in two-dimensional NaNbO3/PVDF (polyvinylidene X-ray powder diffraction (XRD) was used to characterise the
difluoride) polymer composites.35 crystal structure of the samples, the sintered pellets were
Calculations have shown that in the classical perovskite manually ground into powder to ensure that the sample was
system, PbTiO3, hybridisation between the Pb2+ 6s and O2 2p representative of the bulk material. The XRD data were
orbitals results in enlarged strain, leading to the stabilisation collected on a PANalytical X’Pert Pro diffractometer, equipped
of the tetragonal phase.36 Considering that the Bi3+ ion is with an X’Celerator detector, in y/y geometry using Ni-filtered
isoelectronic with Pb2+, a similar effect on structural stability Cu-Ka radiation (l = 1.5418 Å), over the 2y range 51 to 1201 in
might be expected. The effect of Bi3+ introduction is further steps of 0.03341 per step, with an effective count time of 200 s
demonstrated by the BiFeO3–BaTiO3–SrTiO3 solid-solution, per step. The XRD data were analysed using the Rietveld
films of which exhibit high polarisation and ultrahigh energy method through the GSAS suite of programs.40 Raman spectro-
density.37 In the case of NaNbO3, it has previously been shown scopy was used to investigate the structural changes in Na13x
that high recoverable energy density can be achieved in the BixNbO3 using a Renishaw INVIA Raman Microscope with a
composition Na0.7Bi0.1NbO3,20 suggesting that Bi/vacancy He/Ne (633 nm) laser. Differential scanning calorimetry (DSC)
doping can lead to changes in the relative stability of phases was employed to monitor the phase transition behaviour on
in the NaNbO3 system, although no systematic investigation of heating and cooling, at a rate of 10 1C min1, on a TA
this has been reported to date. instruments DSC25. The microstructure of the ceramic pellets
In this work, the structure and electrical behaviour of was examined through scanning electron microscopy (SEM),
Bi/vacancy substituted NaNbO3 is investigated systematically. using an FEI Inspect F (Hillsboro, OR) equipped with energy-
It is shown that Na0.7Bi0.1NbO3, which possesses a cubic dispersive X-ray spectroscopy (EDS).
perovskite structure with 20% A-site vacancies, shows good For electrical characterisation, ceramic pellets were first cut
stability with high dielectric energy storage density. Relaxor and polished into rectangles of ca. 3 mm  3 mm and thickness
ferroelectric (RFE) behaviour is induced in NaNbO3 through around 0.3 mm. Electrodes were applied as silver paste
compositional control, which is more favourable than AFE type (Gwent Electronic Materials Ltd, C2011004D5, Pontypool, U.K.)
phases for high power energy storage based on the fact that AFE onto the two largest parallel faces and fired at 600 1C for 30 min

4290 | J. Mater. Chem. C, 2021, 9, 4289–4299 This journal is © The Royal Society of Chemistry 2021
View Article Online

Journal of Materials Chemistry C Paper

to obtain smooth electrode surfaces. The temperature


dependent dielectric data were obtained by measuring the
capacitance and loss at different frequencies in the range
100 Hz to 1 MHz, using an LCR meter (Agilent, 4284A, Hyogo,
Japan) in a temperature-controlled furnace. The electric
displacement–electric field (D–E) and current–electric field
(I–E) hysteresis loops were measured using a hysteresis tester
Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

(NPL, Teddington, U.K.) with a triangle waveform electric field


at a frequency of 10 Hz.

Results and discussion


Structural and thermal characterisation
SEM images for Na13xBixNbO3 pellets after polishing and
thermal etching are shown in Fig. 1. The microstructure of all Fig. 2 (a) XRD patterns for the studied compositions in the Na13xBixNbO3
system, with detail shown in (b).
compositions revealed a similar particle size of around 5 to
10 mm with good densification. Relative densities of 95.6%,
96.3%, 96.9%, 94.1% and 93.7% of theoretical density were The XRD patterns for the studied compositions are summarised
achieved for pellets of compositions x = 0.015, 0.05, 0.10, 0.15 in Fig. 2a and confirm high purity with good crystallinity. On
and 0.20, respectively. Element maps (Fig. S1, ESI†) confirm a close inspection (Fig. 2b), significant differences are seen in the
uniform distribution of the cations, while EDS results (Table patterns indicating phase evolution with increasing Bi/vacancy
S1, ESI†) are reasonably close to the expected values, indicating content. The fitted diffraction profiles are given in the ESI† as
that the evaporation of Na and Bi is reduced through covering Fig. S2, with the fits in the range 351 to 451 2y shown in Fig. 3.
the pellets with powder during sintering. At low Bi content (x r 0.05) the data are readily modelled on
the orthorhombic P-phase structure of NaNbO3 in space group
Pbcm.41 In the compositional range 0.10 r x r 0.15 the pattern
changes significantly, corresponding to the appearance of a
tetragonal phase in space group P4/mbm, analogous to the
T2-phase in the undoped system.42 At x = 0.20 the pattern
simplifies, as a cubic structure is adopted in space group Pm3% m,
corresponding to the U-phase in NaNbO3.43 A summary of
crystal and refinement parameters is given in Table S2 (ESI†),
with refined structural parameters and significant bond lengths
and angles given in Tables S3 and S4 (ESI†), respectively.
In order to assess the thermal evolution of the lattice
parameters, the orthorhombic and tetragonal unit cell parameters
were transformed to the pseudo-cubic equivalents. These are
plotted as a function of composition in Fig. 4. The a- and c- axes
show general increasing trends throughout the studied compo-
sition range. As x increases up to x = 0.10, the difference
between the a- and b- cell parameters decreases as the structure
transforms from orthorhombic to tetragonal symmetry.
Similarly, as x approaches 0.20, the c-axis curve approaches
that of the a-axis until at x = 0.20 they are equal as the structure
adopts cubic symmetry. The unit cell volume increases steadily
over the studied composition range, with linear increases from
x = 0.015 to x = 0.10 and from x = 0.10 to x = 0.20. Although it is
tempting to explain this in terms of simple ionic radii
differences, there are little published data on the radius of
Bi3+ in 12 coordinate geometry. A number of authors have cited
a value of 1.32 Å for rBi3+ in 12 coordinate geometry,44,45 which
is considerably smaller than the value of 1.39 Å for rNa+ in the
same coordination, available in Shannon’s tables.46 However,
Fig. 1 SEM images for (a) x = 0.015, (b) x = 0.05, (c) x = 0.10, (d) x = 0.15 this value of rBi3+ is much smaller than the value of 1.45 Å
and (e) x = 0.20 compositions in the Na13xBixNbO3 system. obtained through extrapolation of lower coordination number

This journal is © The Royal Society of Chemistry 2021 J. Mater. Chem. C, 2021, 9, 4289–4299 | 4291
View Article Online

Paper Journal of Materials Chemistry C


Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

Fig. 3 Detail of fitted diffraction profiles for (a) x = 0.015, (b) x = 0.05, (c) x = 0.10, (d) x = 0.15 and (e) x = 0.20 compositions in the Na13xBixNbO3 system.

radii in Shannon’s tables, which would suggest that rBi3+ is observed trend in unit cell volume, one must consider three
significantly larger than rNa+ in 12 coordinate geometry. This possible processes for charge compensation on the introduction
contrasts with the situation in 8 coordinate geometry where of bismuth into the NaNbO3 lattice:
rNa+ is slightly larger than rBi3+ (rNa+ = 1.18 Å, rBi3+ = 1.17 Å). A-site vacancy creation
Perhaps more reliably, extrapolation to x = 0.0 of the average Bi2 O3

A-site cation radius in the cubic perovskite system Bi1x 3Na 0
Na ! BiNa þ 2V Na (4)
CaxFeO3x/2 where x r 0.147 (assuming rO2 = 1.40 Å for
6 coordinate geometry) gives a value of 1.40 Å for rBi3+, i.e. B-site cation reduction
slightly larger than rNa+. Whilst this small difference might  Bi2 O3
 0
Na
Na þ 2NbNb ! BiNa þ 2NbNb (5)
explain the observed trend in unit cell volume, another
explanation may lie in the extent of disorder in this system. O interstitials
Indeed, analysis of the average A–O (A = Na, Bi, Vac) and Nb–O
Bi2 O3
 00
bond lengths from the bond length and angle data presented in Na
Na ! BiNa þ Oi (6)
Table S4 (ESI†) show no general trend in the A–O distances,
while the average Nb–O distance shows a general decreasing Mechanisms (5) and (6) would result in compounds with full
trend (Fig. S3a and b, ESI†). Thus, in order to explain the site occupancy of both the A and B cation sites. However, according

4292 | J. Mater. Chem. C, 2021, 9, 4289–4299 This journal is © The Royal Society of Chemistry 2021
View Article Online

Journal of Materials Chemistry C Paper


Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

Fig. 4 Compositional variation of (a) (pseudo) cubic lattice parameters and (b) unit cell volume in the Na13xBixNbO3 system (estimated standard
deviations are given in Table S2, ESI†).

to the stoichiometry of the starting materials used here, if the


charge balance was achieved by mechanisms (5) or (6) then the
compositions would contain excess Nb2O5 which should be
evident in XRD patterns especially at the higher levels of
substitution. Additionally, mechanism (5) would be expected
to enhance the electronic conductivity due to the additional
electrons on niobium, while mechanism (6) would likely lead to
increased ionic conductivity due to the motion of the inter-
stitial oxygen ions. As shown below, the studied compositions
are dielectric insulators, and therefore the evidence suggests
that mechanism (4) occurs. In other words, the substitution of
Na+ by Bi3+ in NaNbO3 creates A-site vacancies, and the solid
solution formula can be expressed as Na13xBixV2xNbO3. It
should be noted that A-site vacancies can also occur intrinsically
through volatilisation of Bi2O3 and/or Na2O during synthesis and
can result in changes in dielectric behaviour, as seen for example
in (Bi0.5Na0.5)TiO3 based systems.48
The co-existence of multiple species, i.e., Bi3+, Na+ and VNa
on the A-site yields a disordered system with increased Fig. 5 Structure of Na13xBixNbO3 showing (a) P-phase, (b) T2a-phase
configurational entropy. With increasing level of substitution, and (c) U-phase. Unit cells are denoted by dashed lines.
the disorder increases and the symmetry changes from the
ordered orthorhombic phase through the tetragonal phase to
the fully disordered cubic phase. The increasing level of
disorder on the A-site is accommodated by an expansion of Pm3% m (Fig. 5c). These phase transitions are reflected in the
the lattice, as is commonly observed at order - disorder phase variation of the average Nb–O–Nb angle (Fig. S3c, ESI†), which
transitions. Further evidence for this increased disorder is seen shows a general increase with increasing x-value, with clear
in the compositional variation of the isotropic displacement steps observed as the crystal symmetry changes from orthor-
parameter, Uiso, for the A-site cations (Table S3, ESI†), which hombic to tetragonal to cubic.
increases significantly with increasing x-value. Raman spectroscopy was employed to monitor the structural
The structures of the observed room temperature phases are variation as a function of Bi content in Na13xBixNbO3 over the
shown in Fig. 5. The AFE orthorhombic phase in space group wavelength range 100 cm1 to 1000 cm1 (Fig. 6 and Table S5,
Pbcm (Fig. 5a), which contains 4 perovskite layers along the ESI†). In common with other perovskite based systems, the
c-axis in the unit cell, shows ab+a/aba/ab+a octahedral Raman spectra can be divided into two main regions, the
tilting with respect to the ideal perovskite structure.49,50 On external and internal modes, i.e. the vibration of Na+/Bi3+
increasing the Bi/vacancy content, the structure increases in against the NbO6 octahedra which occurs below 100 cm1
symmetry with the adoption of the tetragonal P4/mbm structure and the vibration of the NbO6 octahedra themselves above
(Fig. 5b), accompanied by the octahedral tilting changing to 100 cm1.51 The spectrum of Na0.955Bi0.015NbO3 in Fig. 6a has
a0b0c+. Finally, at a high enough level of substitution, the the same mode distribution as that reported for the room
undistorted perovskite structure is adopted in space group temperature phase of the end member, NaNbO3, indicative of

This journal is © The Royal Society of Chemistry 2021 J. Mater. Chem. C, 2021, 9, 4289–4299 | 4293
View Article Online

Paper Journal of Materials Chemistry C

100–300 cm1 and 500–700 cm1 at x = 0.05 is less resolved


than at x = 0.015. This broadening effect reflects the increasing
disorder in the system with increasing incorporation of Bi3+
cations and vacancies. For the higher symmetry compositions
(0.10 r x r 0.20), two broad bands centred around 250 and
600 cm1 are evident, which can be modelled by 8 intense
modes.60 The observed shift in the stretching modes at around
Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

600 cm1, which move to higher wavenumber with increasing


Bi content, reflects that seen in the parent compound NaNbO3
on heating59 and may be associated with the lattice expansion
caused by the Bi3+/vacancy incorporation on neighbouring
A-sites. The observation of Raman active modes for the
x = 0.20 composition, which exhibits the highly symmetric
cubic structure, is likely due to the presence of local distortions
in this highly disordered structure.61
DSC thermograms for the studied compositions on heating
and cooling from room temperature to 350 1C are shown in
Fig. 7. For the x = 0.015 composition, an endothermic peak is
seen on heating with an onset temperature of 298 1C (Tmax =
310 1C) and the corresponding exotherm on cooling is seen at
276 1C (Tmax = 267 1C), analogous to the first order phase
transition between the AFE P-phase and the AFE R-phase in the
parent compound.26,43 At x = 0.05, this phase transition shifts
to lower temperature (Tonset = 151 1C, Tmax = 164 1C on heating
and Tonset = 81 1C and Tmax = 70 1C on cooling). Since the
x = 0.10 and 0.15 compositions, exhibit a tetragonal structure
(akin to the T2 phase in NaNbO3), the P 2 R reversible phase
transition is not visible in the studied temperature range, but
Fig. 6 Raman spectra for Na13xBixNbO3 compositions at room temperature might be expected to occur at lower temperatures. The DSC
showing (a) spectrum for composition x = 0.015, with vibrational modes
thermogram for the x = 0.20 composition which exhibits a cubic
indicated and (b) fitted spectra.
structure like that of the U-phase in NaNbO3 is featureless as
expected.
the orthorhombic structure they share.52–55 Overall, the spectrum
of Na0.955Bi0.015NbO3 can be classified into the librational modes Dielectric response
(which are the roto-vibrational modes of the NbO6 octahedra
The temperature dependencies of relative dielectric permittivity
from 100 to 160 cm1), the bending motion within the bond
(er) and dielectric loss (tan d) for the studied Na13xBixNbO3
angle (O–Nb–O) from 160 to 460 cm1, and the stretching motion
compositions are shown in Fig. 8. In Fig. 8a, the anomaly in
of Nb–O bonds from 460 to 1000 cm1.51 The vibrations of an
relative permittivity at ca. 284 1C for the x = 0.015 composition
isolated equilateral NbO6 octahedron can be described as A1g(n1) +
Eg(n2) + 2F1u(n3, n4) + F2g(n5) + F2u(n6) modes with n1, n2 and n3
attributed to the stretching modes and n4, n5 and n6 to the
bending modes as shown in Fig. 6a.51,56,57
Fig. 6b shows the full fitted Raman spectra for all studied
compositions. The observed spectra change from x = 0.015 to
0.20, reflecting those seen in the stoichiometric end member
NaNbO3 on heating,58–60 consistent with the XRD data which
indicate that the high temperature phases seen in the parent
compound can be stabilised to room temperature through
compositional control. There is a clear change in the spectra
between x = 0.05 and 0.10 as the system changes from orthor-
hombic to tetragonal symmetry, but a less distinct change on
the tetragonal to cubic transition between x = 0.15 and 0.20.
The higher symmetry of the tetragonal and cubic structures is
reflected in fewer Raman active modes, and a decrease in
spectral complexity. The spectrum of the x = 0.05 composition Fig. 7 DSC thermograms for compositions in the system Na13xBixNbO3
is similar to that at x = 0.015, but the peak splitting from on heating and cooling.

4294 | J. Mater. Chem. C, 2021, 9, 4289–4299 This journal is © The Royal Society of Chemistry 2021
View Article Online

Journal of Materials Chemistry C Paper


Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

Fig. 8 Temperature dependencies of dielectric permittivity and loss for (a) x = 0.015, (b) x = 0.05, (c) x = 0.10, (d) x = 0.15 and (e) x = 0.20 compositions
in the Na13xBixNbO3 system; (f) compositional variation of room temperature dielectric permittivity and loss at 1 kHz in the system Na13xBixNbO3.

is associated with the AFE-P to AFE-R phase transition. As seen NaNbO3, we label it T2a. For the x = 0.20 composition, with the
in the DSC data, this transition shifts to lower temperature in cubic U-phase structure at room temperature, frequency
the x = 0.05 composition (ca. 92 1C). For the x = 0.10 and 0.15 dependence of dielectric permittivity is visible below ca. 25 1C,
compositions, which exhibit a tetragonal structure at room characteristic of RFE behaviour. The room temperature relative
temperature, a broad frequency dependent dispersion is seen permittivity at 1 kHz increases from 285 for x = 0.015 (Fig. 8f) to a
in both the relative permittivity and loss spectra. Relative maximum of 1450 at x = 0.10 and then decreases to 560 for
permittivity decreases and the temperature of maximum x = 0.20. The maximum in relative permittivity at x = 0.10 is
permittivity, Tm, increases with increasing frequency. These attributable to the large concentration of PNRs at room
characteristics are typical of relaxor ferroelectrics (RFEs) and temperature for this composition.
have been explained in terms of the existence of polar nanoregions Fig. 9 shows the D–E and I–E loops at room temperature for
(PNRs) within a non-polar matrix.31,62 The average structure for the the studied compositions. The x = 0.015 composition (Fig. 9a)
tetragonal phase in the present case is indistinguishable from that exhibits similar ferroelectric D–E loops to pure NaNbO3, with a
in pure NaNbO3, with both structures showing P4/mbm symmetry. saturation polarisation of 0.31 C m2 and a remnant polarisation
However, while PNRs are undetectable by conventional XRD, they of 0.29 C m2. In the corresponding I–E loop (Fig. 9b) at lower
are very sensitive to dielectric tests63,64 and to distinguish the RFE electric field, no current peaks are observed (Fig. 9b inset).
tetragonal phase identified here from the PE T2 phase in pure However, at fields above 8 kV mm1, four current peaks are

This journal is © The Royal Society of Chemistry 2021 J. Mater. Chem. C, 2021, 9, 4289–4299 | 4295
View Article Online

Paper Journal of Materials Chemistry C


Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

Fig. 9 D–E and I–E loops at different electric field amplitudes for (a and b) x = 0.015, (c and d) x = 0.05, (e and f) x = 0.10, (g and h) x = 0.15 and (i and j)
x = 0.20 compositions in the Na13xBixNbO3 system.

seen, two peaks in the 1st quadrant and two in the 3rd quadrant On increasing the Bi/vacancy content further to x = 0.10
in Fig. 9b. Unlike ferroelectric materials which show two current (Fig. 9e and f), where the X-ray data confirm a tetragonal phase
peaks in their I–E loops (one in the 1st quadrant and the other in (T2a), samples exhibit RFE behaviour with Ds = 0.23 C m2 and
the 3rd quadrant), associated with domain switching, the four Dr = 0.015 C m2. In the I–E loop, when the applied electric field
current peaks here represent irreversible field induced phase is below 6 kV mm1, the electric field is too weak to give rise to
transitions. On increasing field, EF at around 5.4 kV mm1 is due a response of the polar regions. Therefore, the low electric field
to the AFE (anti-ferroelectric) to FE (ferroelectric) phase transition, D–E loop (Fig. 9e inset) exhibits near-linear dielectric
while EB at around 3.3 kV mm1 is due to the FE to AFE phase behaviour, with the corresponding low field I–E loops
transition under reverse field.65 The irreversible AFE behaviour of rectangular in shape. On increasing the electric field further,
this sample with micron-sized grains is consistent with that in the shape of the I–E loop gradually changes to give a single peak
pure NaNbO3 of similar grain size (NaNbO3 samples with nano- with a maximum current near 0 kV mm1, characteristic of RFE
metre sized grains exhibit the polar FE phase Pmc21).19 Since there behaviour and indicating the response of PNRs under strong
is little difference between Dr and Ds, the recoverable energy electric field. The low Dr is attributable to the extent of disorder
density is very low. For the x = 0.05 composition (Fig. 9c and d), on the A-site at this composition, which leads to local PNRs
irreversible AFE behaviour is maintained, but the D–E loop is rather than the ordered anti-parallel dipoles of an AFE phase.
more tilted than at x = 0.015 with a larger difference between Ds A Wrec value of around 1.4 J cm3 with a high efficiency, Z, of
and Dr (0.29 C m2 and 0.15 C m2, respectively). The decreased 76.5% is obtained at an electric field of 15 kV mm1 for this
value of Dr allows for an increased Wrec for this composition, composition. This is just below the breakdown field strength of
although still relatively low. The I–E loops show a decrease in EB 16 kV mm1. RFE behaviour is maintained at x = 0.15, but the
(2.4 kV mm1), but an increase in EF (the EF peak is only I–E loops are somewhat tilted and the D–E loop is narrower and
partially visible due to breakdown at high fields). Thus, while straighter than at x = 0.10, with a much-reduced saturation
the AFE to FE transition can be induced at an applied electric field polarisation of 0.10 C m2.
of 9 kV mm1 for x = 0.015, the FE phase is less stable, requiring a At x = 0.20, the I–E loops show that the maximum current
much larger applied electric field for x = 0.05 (above 12 kV mm1), occurs at the maximum applied field and the loops are significantly
indicating that increasing Bi/vacancy content leads to decreasing tilted, indicative of a high leakage current.66 The D–E loops
stability of the FE phase. show hysteresis, which is associated with loss produced by the

4296 | J. Mater. Chem. C, 2021, 9, 4289–4299 This journal is © The Royal Society of Chemistry 2021
View Article Online

Journal of Materials Chemistry C Paper


Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

Fig. 10 (a) Compositional variation of energy storage density and efficiency (Z) in Na13xBixNbO3 measured at room temperature with 10 Hz A.C. and (b)
comparison of recoverable energy density in Na0.7Bi0.1O3 with those in other lead-free ceramics systems: BaTiO3 (BT), (Bi0.5Na0.5)TiO3 (BNT),
(Ba0.6Sr0.4)TiO3, (BST) BiFeO3 (BFO) and AgNbO3 (AN).67–74

heating effect of the leakage current. Bearing in mind the applied field is 15 kV mm1, which is notable among
permittivity and loss data (Fig. 8e), which show frequency current lead-free ceramic systems at similar applied fields
dependent peaks below 25 1C, and the observed cubic structure (Fig. 10b).
of this composition at room temperature, the I–E and D–E Fig. 11 is a schematic illustrating the local electrical
behaviour suggest that the composition may be described as a structure of the different phases of Na13xBixNbO3, with and
lossy paraelectric. without applied electric field. For low x-value compositions, the
Fig. 10a summarises the compositional variation of energy AFE phase is dominant (Fig. 11a) in the absence of an applied
storage density and efficiency in the studied compositions. field and is accompanied by the presence of the FE phase
Among these compositions, the energy storage efficiency is separated by areas of incommensurate structure as seen in the
increased from a low level (less than 25%) for the irreversible TEM images of pure NaNbO3.75 On application of an applied
AFE phase (x r 0.05) to a high level (above 75%) for the field, a phase transition from the AFE phase to the FE phase
RFE phase (0.1 r x r 0.15). The x = 0.10 composition exhibits occurs (Fig. 11b), leaving only the polar FE phase domains.
the highest energy storage density of 1.4 J cm3 when the At intermediate compositions where RFE behaviour occurs,

Fig. 11 Schematic illustrating the dielectric behaviour in (a and b) compositions x r 0.05 (blue, yellow and green colours represent the FE Q-phase,
incommensurate phase and the AFE P phase, respectively), (c and d) 0.10 r x r 0.15 and (e and f) x = 0.20 in the Na13xBixNbO3 system, (a, c and e) prior
to application of an applied electric field and (b, d and f) in the presence of an applied electric field.

This journal is © The Royal Society of Chemistry 2021 J. Mater. Chem. C, 2021, 9, 4289–4299 | 4297
View Article Online

Paper Journal of Materials Chemistry C

small PNRs are orientated randomly within grains in the References


absence of an applied field, giving a net polarisation of
zero (Fig. 11c). In the presence of an applied field, the PNRs 1 Z. Yan, D. Zhang, X. Zhou, H. Qi, H. Luo, K. Zhou,
grow in number and size, exhibiting the same polarisation I. Abrahams and H. Yan, J. Mater. Chem. A, 2019, 7,
direction as the electric field (Fig. 11d). At high x-values, the 10702–10711.
lossy paraelectric phase shows randomly oriented dipoles, 2 X. Qiao, F. Zhang, D. Wu, B. Chen, X. Zhao, Z. Peng, X. Ren,
illustrated as spring dipoles in Fig. 11e. On application of an P. Liang, X. Chao and Z. Yang, Chem. Eng. J., 2020,
Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

applied field, these dipoles orientate in the field direction 388, 124158.
(Fig. 11f). 3 L. Zhang, Y. Pu, M. Chen, T. Wei and X. Peng, Chem. Eng. J.,
2020, 383, 123154.
4 G. A. Samara, Phys. Rev. B: Solid State, 1970, 1, 3777–3786.
Conclusions 5 X. Hao, J. Zhai, L. B. Kong and Z. Xu, Prog. Mater. Sci., 2014,
63, 1–57.
A-site substitution of sodium in NaNbO3 by Bi and 6 P. Niemiec, R. Skulski, D. Bochenek and P. Wawrzała, Phase
vacancies leads to the room temperature stabilisation of phases Transitions, 2013, 86, 267–274.
analogous to the high temperature phases of NaNbO3, 7 Z.-G. Ye, Ferroelectrics, 1996, 184, 193–208.
including tetragonal and cubic phases with high purity and 8 A. K. Yadav, H. Fan, B. Yan, C. Wang, M. Zhang, J. Ma,
good crystallinity. In contrast to pure NaNbO3, the tetragonal W. Wang, W. Dong and S. Wang, Ceram. Int., 2020, 46,
phase exhibits RFE behaviour and to distinguish it from 5681–5688.
the PE T2 phase in NaNbO3, it has been labelled here as T2a. 9 Q. Li, W. Zhang, C. Wang, L. Ning, C. Wang, Y. Wen, B. Hu
The stabilisation of these phases may be attributed to the and H. Fan, J. Alloys Compd., 2019, 775, 116–123.
increased disorder caused by the enhanced configurational 10 H. Qi and R. Zuo, J. Mater. Chem. A, 2019, 7, 3971–3978.
entropy due to the mixture of three species (Na+, Bi3+, 11 D. Yang, J. Gao, L. Shu, Y.-X. Liu, J. Yu, Y. Zhang, X. Wang,
and VNa) on the A-site. The structural evolution is accompanied B.-P. Zhang and J.-F. Li, J. Mater. Chem. A, 2020, 8,
by changes in the electrical behaviour from AFE to RFE to 23724–23737.
lossy PE. 12 H. F. Shi, G. Q. Chen, C. L. Zhang and Z. G. Zou, ACS Catal.,
The highest energy storage density of 1.4 J cm3 at 15 kV mm1 2014, 4, 3637–3643.
was obtained for the T2a phase, Na0.7Bi0.1NbO3, which shows 13 P. Li, S. Ouyang, G. Xi, T. Kako and J. Ye, J. Phys. Chem. C,
RFE behaviour with an efficiency of 76.5%. The composition 2012, 116, 7621–7628.
shows the highest room temperature permittivity and exhibits a 14 J. Lv, T. Kako, Z. Li, Z. Zou and J. Ye, J. Phys. Chem. C, 2010,
tetragonal structure in space group P4/mbm. While the values of 114, 6157–6162.
energy storage density are not the highest, significantly higher 15 P. Li, S. Ouyang, Y. Zhang, T. Kako and J. Ye, J. Mater. Chem.
values are expected in this system if prepared as thin films and A, 2013, 1, 1185–1191.
in the case of bulk ceramics on further optimisation of densi- 16 C. Yan, L. Nikolova, A. Dadvand, C. Harnagea, A. Sarkissian,
fication processes. D. F. Perepichka, D. Xue and F. Rosei, Adv. Mater., 2010, 22,
1741–1745.
17 J. H. Jung, M. Lee, J. I. Hong, Y. Ding, C. Y. Chen, L. J. Chou
Author contributions and Z. L. Wang, ACS Nano, 2011, 5, 10041–10046.
18 L. A. Reznitchenko, A. V. Turik, E. M. Kuznetsova and
L. Zhang: investigation, writing – original draft; Z. Yan:
V. P. Sakhnenko, J. Phys.: Condens. Matter, 2001, 13,
investigation; T. Chen: investigation; H. Luo, investigation,
3875–3881.
supervision; H. Zhang, formal analysis, validation, T. Khanom,
19 H. Zhang, B. Yang, H. Yan and I. Abrahams, Acta Mater.,
investigation; D. Zhang, supervision, I. Abrahams: supervision,
2019, 179, 255–261.
formal analysis, writing – review & editing and H. Yan:
20 M. Zhou, R. Liang, Z. Zhou and X. Dong, J. Mater. Chem. A,
conceptualisation, supervision, writing – review & editing.
2018, 6, 17896–17904.
21 M. Zhou, R. Liang, Z. Zhou, S. Yan and X. Dong, ACS
Sustainable Chem. Eng., 2018, 6, 12755–12765.
Conflicts of interest 22 A. M. Glazer and H. D. Megaw, Acta Crystallogr., Sect. A: Cryst.
There are no conflicts of interest to declare. Phys., Diffr., Theor. Gen. Crystallogr., 1973, 29, 489–495.
23 H. D. Megaw, Ferroelectrics, 1974, 7, 87–89.
24 S. K. Mishra, N. Choudhury, S. L. Chaplot, P. S. R. Krishna
Acknowledgements and R. Mittal, Phys. Rev. B: Condens. Matter Mater. Phys.,
2007, 76, 024110.
Ludan Zhang thanks Queen Mary University of London and the 25 K. E. Johnston, C. C. Tang, J. E. Parker, K. S. Knight,
China Scholarship Council (CSC) for Financial support through P. Lightfoot and S. E. Ashbrook, J. Am. Chem. Soc., 2010,
a joint scholarship (201606100039). 132, 8732–8746.

4298 | J. Mater. Chem. C, 2021, 9, 4289–4299 This journal is © The Royal Society of Chemistry 2021
View Article Online

Journal of Materials Chemistry C Paper

26 H. Guo, H. Shimizu and C. A. Randall, Appl. Phys. Lett., 2015, 52 K. K. Mishra, V. Sivasubramanian and A. K. Arora, J. Raman
107, 112904. Spectrosc., 2011, 42, 517–521.
27 S. Wu, W. Li, M. Lin, Q. Burlingame, Q. Chen, A. Payzant, 53 S. J. Lin, D. P. Chiang, Y. F. Chen, C. H. Peng, H. T. Liu,
K. Xiao and Q. M. Zhang, Adv. Mater., 2013, 25, 1734–1738. J. K. Mei, W. S. Tse, T. R. Tsai and H. P. Chiang, J. Raman
28 Z. M. Dang, J. K. Yuan, S. H. Yao and R. J. Liao, Adv. Mater., Spectrosc., 2006, 37, 1442–1446.
2013, 25, 6334–6365. 54 E. Bouziane, M. D. Fontana and M. Ayadi, J. Phys.: Condens.
29 L. Gao, H. Guo, S. Zhang and C. A. Randall, Appl. Phys. Lett., Matter, 2003, 15, 1387–1395.
Published on 15 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:17:05 AM.

2018, 112, 092905. 55 Z. X. Shen, X. B. Wang, M. H. Kuok and S. H. Tang, J. Raman


30 Z. Y. Liu, J. S. Lu, Y. Q. Mao, P. R. Ren and H. Q. Fan, J. Eur. Spectrosc., 1998, 29, 379–384.
Ceram. Soc., 2018, 38, 4939–4945. 56 Y. Shiratori, A. Magrez, J. Dornseiffer, F. H. Haegel, C. Pithan
31 L. E. Cross, Ferroelectrics, 1987, 76, 241–267. and R. Waser, J. Phys. Chem. B, 2005, 109, 20122–20130.
32 Z. Yang, H. Du, L. Jin, Q. Hu, S. Qu, Z. Yang, Y. Yu, X. Wei 57 H. U. Khan, K. Alam, M. Mateenullah, T. Blaschke and
and Z. Xu, J. Eur. Ceram. Soc., 2019, 39, 2899–2907. B. S. Haq, J. Eur. Ceram. Soc., 2015, 35, 2775–2789.
33 Y. Fan, Z. Zhou, R. Liang and X. Dong, J. Eur. Ceram. Soc., 58 X. B. Wang, Z. X. Shen, Z. P. Hu, L. Qin, S. H. Tang and
2019, 39, 4770–4777. M. H. Kuok, J. Mol. Struct., 1996, 385, 1–6.
34 H. Qi, R. Zuo, A. Xie, A. Tian, J. Fu, Y. Zhang and S. Zhang, 59 I. Y. Yu, P. Simon, E. Gagarina, L. Hennet, D. Thiaudière,
Adv. Funct. Mater., 2019, 0, 1903877. V. I. Torgashev, S. I. Raevskaya, I. P. Raevskii,
35 Z. B. Pan, B. H. Liu, J. W. Zhai, L. M. Yao, K. Yang and L. A. Reznitchenko and J. L. Sauvajol, J. Phys.: Condens.
B. Shen, Nano Energy, 2017, 40, 587–595. Matter, 2005, 17, 4977–4990.
36 R. E. Cohen, Nature, 1992, 358, 136–138. 60 P. S. R. Krishna, S. K. Mishra, A. B. Shinde, S. Kesari and
37 H. Pan, F. Li, Y. Liu, Q. Zhang, M. Wang, S. Lan, Y. Zheng, R. Rao, Ferroelectrics, 2017, 510, 34–42.
J. Ma, L. Gu, Y. Shen, P. Yu, S. Zhang, L. Chen, Y. Lin and 61 O. Svitelskiy, J. Toulouse, G. Yong and Z. G. Ye, Phys. Rev. B:
C. Nan, Science, 2019, 365, 578. Condens. Matter Mater. Phys., 2003, 68, 104107.
38 H. Qi, R. Zuo, A. Xie, A. Tian, J. Fu, Y. Zhang and S. Zhang, 62 V. V. Shvartsman, D. C. Lupascu and D. J. Green, J. Am.
Adv. Funct. Mater., 2019, 29, 1903877. Ceram. Soc., 2012, 95, 1–26.
39 L. Jin, F. Li and S. Zhang, J. Am. Ceram. Soc., 2014, 97, 1–27. 63 H. Yu and Z.-G. Ye, J. Appl. Phys., 2008, 103, 034114.
40 A. C. Larson and R. B. Von Dreele, Los Alamos National 64 L. Bian, X. Qi, K. Li, Y. Yu, L. Liu, Y. Chang, W. Cao and
Laboratory Report, No. LAUR-86–748, 1987. S. Dong, Adv. Funct. Mater., 2020, 30, 2001846.
41 K. E. Johnston, C. C. Tang, J. E. Parker, K. S. Knight, 65 G. Viola, R. McKinnon, V. Koval, A. Adomkevicius, S. Dunn
P. Lightfoot and S. E. Ashbrook, J. Am. Chem. Soc., 2010, and H. X. Yan, J. Phys. Chem. C, 2014, 118, 8564–8570.
132, 8732–8746. 66 H. Yan, F. Inam, G. Viola, H. Ning, H. Zhang, Q. Jiang,
42 R. H. Mitchell, B. J. Kennedy and K. S. Knight, Phys. Chem. T. A. O. Zeng, Z. Gao and M. J. Reece, J. Adv. Dielectr., 2012,
Miner., 2018, 45, 77–83. 01, 107–118.
43 S. K. Mishra, R. Mittal, V. Y. Pomjakushin and S. L. Chaplot, 67 T. Wang, L. Jin, Y. Tian, L. Shu, Q. Hu and X. Wei, Mater.
Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 134105. Lett., 2014, 137, 79–81.
44 Y. Yuan, C. J. Zhao, X. H. Zhou, B. Tang and S. R. Zhang, 68 T. Wang, L. Jin, C. Li, Q. Hu and X. Wei, J. Am. Ceram. Soc.,
J. Electroceram., 2010, 25, 212–217. 2015, 98, 559–566.
45 C. M. Lau, X. W. Xu and K. W. Kwok, Appl. Surf. Sci., 2015, 69 W. Cao, W. Li, Y. Feng, T. Bai, Y. Qiao, Y. Hou, T. Zhang,
336, 314–320. Y. Yu and W. Fei, Appl. Phys. Lett., 2016, 108, 202902.
46 R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., 70 Z. C. Liu, P. R. Ren, C. B. Long, X. Wang, Y. H. Wan and
Theor. Gen. Crystallogr., 1976, 32, 751–767. G. Y. Zhao, J. Alloys Compd., 2017, 721, 538–544.
47 K. Sardar, J. Hong, G. Catalan, P. K. Biswas, M. R. Lees, 71 Y. Yao, Y. Li, N. Sun, J. Du, X. Li, L. Zhang, Q. Zhang and
R. I. Walton, J. F. Scott and S. A. T. Redfern, J. Phys.: X. Hao, Ceram. Int., 2018, 44, 5961–5966.
Condens. Matter, 2012, 24, 045905. 72 Z. Song, S. Zhang, H. Liu, H. Hao, M. Cao, Q. Li, Q. Wang,
48 J. Shi, H. Fan, X. Liu, Y. Ma and Q. Li, J. Alloys Compd., 2015, Z. Yao, Z. Wang and M. T. Lanagan, J. Am. Ceram. Soc., 2015,
627, 463–467. 98, 3212–3222.
49 A. M. Glazer, Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. 73 D. Zheng, R. Zuo, D. Zhang and Y. Li, J. Am. Ceram. Soc.,
Mater., 1972, 28, 3384–3392. 2015, 98, 2692–2695.
50 A. Glazer, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. 74 J. Ma, S. Yan, C. Xu, G. Cheng, C. Mao, J. Bian and G. Wang,
Gen. Crystallogr., 1975, 31, 756–762. Mater. Lett., 2019, 247, 40–43.
51 M. Jauhari, S. K. Mishra, R. Mittal and S. L. Chaplot, 75 H. Guo, H. Shimizu, Y. Mizuno and C. A. Randall, J. Appl.
J. Raman Spectrosc., 2019, 0, 1–9. Phys., 2015, 118, 054102.

This journal is © The Royal Society of Chemistry 2021 J. Mater. Chem. C, 2021, 9, 4289–4299 | 4299

You might also like