Sunteți pe pagina 1din 9

Article

pubs.acs.org/cm

Interplay between Electrochemistry and Phase Evolution of the P2type Nax(Fe1/2Mn1/2)O2 Cathode for Use in Sodium-Ion Batteries
Wei Kong Pang,,, Sujith Kalluri,, Vanessa K. Peterson,*, Neeraj Sharma, Justin Kimpton,
Bernt Johannessen, Hua Kun Liu, Shi Xue Dou, and Zaiping Guo*,

School of Mechanical, Materials, and Mechatronic Engineering, Institute for Superconducting & Electronic Materials, Faculty of
Engineering, University of Wollongong, Wollongong, NSW 2522, Australia

Australian Nuclear Science and Technology Organisation, Locked Bag 2001, Kirrawee DC, NSW 2232, Australia

School of Chemistry, University of New South Wales, Sydney, NSW 2052, Australia

Australian Synchrotron, Clayton, VIC 3168, Australia


S Supporting Information
*

ABSTRACT: Sodium-ion batteries are the next-generation in


battery technology; however, their commercial development is
hampered by electrode performance. The P2-type
Na2/3(Fe1/2Mn1/2)O2 with a hexagonal structure and P63/mmc
space group is considered a candidate sodium-ion battery
cathode material due to its high capacity (190 mAhg1) and
energy density (520 mWhg1), which are comparable to
those of the commercial LiFePO4 and LiMn2O4 lithium-ion
battery cathodes, with previously unexplained poor cycling
performance being the major barrier to its commercial
application. We use operando synchrotron X-ray powder
diraction to understand the origins of the capacity fade of
the Na2/3(Fe1/2Mn1/2)O2 material during cycling over the
relatively wide 1.54.2 V (vs Na) window. We found a complex phase-evolution, involving transitions from P63/mmc (P2-type at
the open-circuit voltage) to P63 (OP4-type when fully charged) to P63/mmc (P2-type at 3.42.0 V) to Cmcm (P2-type at 2.0
1.5 V) symmetry structures during the desodiation and sodiation of the Na2/3(Fe1/2Mn1/2)O2 cathode. The associated large cellvolume changes with the multiple two-phase reactions are likely to be responsible for the poor cycling performance, clearly
suggesting a 2.04.0 V window of operation as a strategy to improve cycling performance. We demonstrated here that the P2type Na2/3(Fe1/2Mn1/2)O2 cathode is able to deliver 25% better cycling performance with the strategic operation window. This
signicant improvement in cycling performance implies that by characterizing the phase evolution and reaction mechanisms
during battery function we are able to propose these modications to the conditions of battery use that improve performance,
highlighting the importance of the interplay between structure and electrochemistry.

INTRODUCTION
Since the commercialization of the rst lithium-ion battery
(LIB), featuring a LiCoO2 cathode and graphite anode, by Sony
Corporation in 1991, LIBs are now extensively used for energy
storage due to their relatively high energy-density, long cyclelife, and low cost.1 However, the search is underway for a
replacement technology, mainly because of the limited
abundance of lithium. Sodium is the sixth-highest abundant
element in the Earths crust, making it cheaper to produce, and
sodium-ion batteries (SIBs) are now attracting attention as
promising alternatives for large-scale energy storage applications. Sodium is also the next lightest and smallest alkali-metal
after lithium, therefore providing useful gravimetric and
volumetric energy-storage density.25 Beyond these advantages
that sodium has to oer battery technology, sodium also
exhibits a redox potential well suited to battery applications (0.3
V vs Li) with electrochemical function similar to that of lithium
2015 American Chemical Society

in LIBs. Taken together, these attributes make sodium a serious


contender for future alternative battery chemistries.25
Recently, alternative cathode materials for SIBs based on
suldes, uorides, phosphates, sulfates, and oxides (including
layered transition-metal oxides) have been investigated.3,4,6 Ellis
et al. indicated that the layered sodiummetal oxides, including
O3-type and P2-types, using Delmass notation,7 exhibit
promising electrode properties.4 The O3-type layered oxides
NaMO2 (including M = Co, Cr, V, Ni0.5Mn0.5, Fe0.5Mn0.5, and
Ni1/3Co1/3Mn1/3)814 are suitable cathode materials for SIBs
but suer from poor cycle and rate performance, with an initial
discharge capacity limited to 160 mAhg1 or lower.
Monoclinic NaMnO2 was also found to have an initial
Received: March 12, 2015
Revised: April 9, 2015
Published: April 10, 2015
3150

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials
discharge capacity of 160 mAhg1, where the two electrochemical reactions involved lead to structural instability and
poor cycling performance.15 While P2-type layered oxides such
as Na0.6MnO216 and Na0.7CoO217 have a better initial discharge
capacity than the O3-type layered oxides, again, poor cycling
performance remains a major issue. Environmental concerns
are also driving research aimed at replacing toxic and costly
metals such as Cr and Co with safe and abundant metals such
as Mn, Fe, and Ni, possibly with these intermixed.1820 Safer
and aordable layered sodium metal-oxides containing mixedvalence Fe and Mn, such as the P2-type Nax(FeyMn1y)O2
material (x = 2/3; y = 1/2), have been obtained by solid-state,
coprecipitation, autocombustion, and solgel processes.2024
Importantly, these environmentally friendly materials deliver an
excellent initial discharge capacity of 190 mAhg1 at low
current density in the 1.54.3 V range with an energy density
of 520 mWhg11. These electrodes are therefore comparable in performance with the LiFePO4 (530 mWhg11)
and LiMn2O4 (460 mWhg11) cathodes used commercially
in LIBs and open up a new avenue for the development of
future cathode materials. Unfortunately, the poor cycling
performance of the Nax(FeyMn1y)O2 (x = 2/3; y = 1/2)
materials presents a major hurdle to large-scale commercial
application. The electrochemical behavior of materials is
governed by their structural and chemical evolution. For
example, the delithiation of the P2LixVO2 (x = 0.80) cathode
proceeds through multiple two-phase and solidsolution
reactions to form P2LixVO2 (x = 0.50), leading to structural
instability and poor cycling performance. 1 0 P2
Na2/3(Fe1/2Mn1/2)O2 crystallizes in the hexagonal structure
with a P63/mmc space group and is considered a candidate
cathode material for SIBs, with the sodium insertion and
extraction mechanisms studied using ex situ synchrotron X-ray
powder diraction (SXRPD) and X-ray absorption spectroscopy (XAS).20 It is reported that the Nax(Fe1/2Mn1/2)O2 cathode
rst undergoes a solidsolution reaction up to 3.8 V (from x =
0.66 to 0.4) before a two-phase reaction between P63/mmc
(P2) and P6m2 (OP4) is observed, alongside Fe3+/Fe4+ and
Mn3+/Mn4+ transitions.20 While the established mechanism is
seemingly well-accepted, the ex situ study involved the removal
and postprocessing of composite electrodes from a coin-cell
and therefore did not directly reveal the entire mechanism of
cathode function. Recent studies of the Nax(Fe1/2Mn1/2)O2
cathode using operando SXRPD reported that Nax(Fe1/2Mn1/2)O2 (x 1) crystallized in the Cmcm space group (also P2-type)
and rst transformed to the P63/mmc symmetry (P2) through a
two-phase reaction upon desodiation. Further desodiation
caused the P2 structure to undergo both solidsolution (0.35
< x < 0.82) and two-phase reactions, transforming into a Z
phase with a high degree of disorder. However, neither report
has explained the structurefunction relationship that underpins the electrochemical properties of this promising SIB
cathode.
Since the energy density of a battery mainly depends
predominately on the cathode material, the cathode is the
component of focus in the consideration of new SIB systems.
The properties and features of the Nax(Fe1/2Mn1/2)O2 cathode
are suciently promising for SIB applications to warrant an indepth understanding of this material both structurally and
electrochemically in the pursuit of overcoming its poor cycling
performance and thus is the subject of the present study. In
particular, an understanding of the reaction pathways that
control performance will enable the rational improvement of

electrode materials. In the past, operando SXRPD using


customized coin-cells has successfully examined the structural
evolution of electrode materials.2528 In this study, we prepare
the nanosized P2-type Nax(Fe1/2Mn1/2)O2 (x = 2/3) material
via a facile single-step solgel method. The nanoparticulate
cathode delivers excellent discharge capacity and energy density
in a coin-cell but relatively poor cycling performance and
energy eciency. We study the structurefunction relationship
of the Nax(Fe1/2Mn1/2)O2 cathode using operando SXRPD
during sodium insertion and deinsertion and suggest a rational
improvement in the use of the P2-type cathode. Our study
reveals that the P2 hexagonal phase undergoes a solidsolution
reaction during the Mn3+/Mn4+ transition and a solidsolutionlike two-phase reaction at 4.1 V and above to form an OP4
phase with a P63 space group during the Fe3+/Fe4+ transition.
The OP4 phase returns to the P2-type hexagonal phase upon
discharge, and then transforms to a P2-type orthorhombic
phase with a Cmcm space group upon discharge below 2.0 V.
Importantly, we identify the signicant volumetric and
structural changes associated with the orderdisorder (P2OP4) and hexagonalorthorhombic two-phase transitions, and
suggest an alternative operating scheme for the battery that
avoids this, enhancing cycle life.

EXPERIMENTAL SECTION

The precursor solution was prepared by dissolving stoichiometric


amounts of sodium acetate, iron nitrate, and manganese acetate (all
from Sigma-Aldrich) in ethanol and N,N-dimethylformamide (DMF).
After stirring for an hour, 10 wt % of polyvinylpyrrolidone (PVP,
molecular weight 1,300,000 gmol1), which acts as a gelling agent,
was added to the resultant solution. The precursor solution was ovendried at 100 C overnight before calcination at 900 C for 2 h in air.
The obtained P2-type Na2/3(Fe1/2Mn1/2)O2 powder was quenched to
room temperature and stored in an Ar-lled glovebox.
The as-prepared P2-type Na2/3(Fe1/2Mn1/2)O2 (NFMO) was
characterized using X-ray powder diraction (XRPD) (GBC, MMA)
equipped with CuK radiation and high-resolution neutron powder
diraction (NPD) using ECHIDNA29 at the Australian Nuclear
Science and Technology Organisation. The wavelength of the neutron
beam was 1.62380(3) , determined using the La11B6 NIST standard
reference material (SRM) 660b. The NPD data were obtained in the
2 angular range 6.5 to 165.2 with a step size of 0.125 over 6 h.
Fullprof/WinPlotR30 were employed to perform Rietveld analysis of
the high-resolution NPD and XRPD data. The parameters including
the background coecients, zero-shift, peak shape parameters, lattice
parameters, oxygen positional parameter, sodium occupancy, and
isotropic atomic displacement parameters (B) were rened. For the
NPD measurement, powders were packed into an airtight 9 mm
vanadium can. The morphology and particle size of the as-prepared
NFMO powder were examined using scanning electron microscopy
(SEM, JEOL JSM-7500, Japan).
A customized CR2032 coin-cell for use in synchrotron X-ray
powder diraction (SXRPD) experiments was designed and
made.2527 Electrodes were prepared by mixing the as-prepared
NFMO powder with carbon black (Super P, TIMCAL) and
polyvinylidene uoride (in a 80:10:10 weight ratio) in anhydrous Nmethyl-2-pyrrolidinone (Sigma-Aldrich, 99.5%) to form a homogeneous slurry. The slurry was uniformly pasted onto aluminum foil
before being dried in a vacuum oven at 100 C for 24 h and pressed
prior to the assembly of the coin-cell. The coin-cell was assembled
using sodium disks (Sigma-Aldrich) as the counter electrode, porous
glass ber (Millipore) as a separator, and 1 M NaClO4 in propylene
carbonate as an electrolyte with 2 wt % uorinated ethyl carbonate
electrolyte additive. Holes for permitting synchrotron X-ray beam
transmission were punched in the top and bottom casing and then
sealed with polyimide lm (Kapton, DuPont) and wax. The cell was
galvanostatically charged and discharged over 1.54.2 V vs Na at a
3151

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials
constant current of 0.07 mA (equivalent to 0.05 C) during data
collection. SXRPD experiments were conducted on the powder
diraction beamline at the Australian Synchrotron where data were
collected every 3.40 min at 0.68922(1) (determined using the NIST
SRM 660b) during the battery cycle using a MYTHEN microstrip
detector. The lattice response of P2Na2/3(Fe1/2Mn1/2)O2 during the
cycle was extracted and examined using a single peak-tting analysis
with the large-array manipulation program (LAMP)31 to track the
changes in peak position, intensity, and width of the 002 and 100
cathode reections during charging and discharging.
Fe and Mn K-edge X-ray absorption near-edge structure (XANES)
spectra were collected using the same cells as per the SXRPD
experiment on the X-ray absorption beamline at the Australian
Synchrotron. The energy step size through the edge was 0.3 eV. The
exposure time per step was xed to 1 s. For the Fe K-edge, FeO and
Fe2O3 powders were used as Fe2+ and Fe3+ references, respectively. Fe
K-edge spectra of [Fe(O)(N4Py)](ClO4)2 were sourced from Rohde
et al.32 For the Mn K-edge, LiMnO2 and MnO2 were used as Mn3+ and
Mn4+ references, respectively.
The electrochemical properties of the P2-type Na2/3(Fe1/2Mn1/2)O2
cathode were evaluated in CR2032 coin-cells with a half-cell
conguration as assembled in an Ar-lled glovebox (MBraun,
Germany). Galvanostatic chargedischarge behavior was assessed
using an automatic battery analyzer (Land, China).

Table 1. Crystallographic Details of the As-prepared NFMO


Cathode Obtained from Rietveld Analysis Using NPD Dataa
Na0.70(2)(Fe1/2Mn1/2)O2
space group = P63/mmc
a = b = 2.91883(1) and c = 11.2830(2)
atom

site

B (2)

occupancy

Na
Na
Mn
Fe
O

2b
2d
2a
2a
4f

0
1/3
0
0
1/3

0
2/3
0
0
2/3

1/4
3/4
0
0
0.0903(2)

1.5(3)
4.9(4)
0.8(1)b
0.8(1)b
0.54(2)

0.24(1)
0.46(2)
1/2
1/2
1

Rwp = 18.6, 2 = 2.45, and Bragg-R factor = 4.18. bThe B values of Fe


and Mn are constrained to be the same.
a

Table 2. Crystallographic Details of NFMO Obtained from


Rietveld Analysis Using XRPD Dataa
Na0.70(2)(Fe1/2Mn1/2)O2
space group = P63/mmc
a = b = 2.9163(2) and c = 11.295(1)

RESULTS AND DISCUSSION


To understand the structurefunction relationship of the
nanoparticulate P2Nax(Fe1/2Mn1/2)O2 (x = 2/3) (NFMO)
cathode, XRPD, high-resolution NPD, and XANES were
employed to study the crystallographic and electronic structure,
and its evolution. The XRPD and NPD patterns are shown in
Figure 1a and b, respectively. The rened structures obtained
using NPD and XRPD data are summarized in Tables 1 and 2,
respectively. Both XRPD and NPD data show that the NFMO

atom

site

B (2)

occupancy

Na
Na
Mn
Fe
O

2b
2d
2a
2a
4f

0
1/3
0
0
1/3

0
2/3
0
0
2/3

1/4
3/4
0
0
0.06378(8)

1.0(5)
1.0(5)
1.0(5)b
1.0(5)b
0.1(2)

0.26(1)
0.30(1)
1/2
1/2
1

Rwp = 36.4, 2= 6.73, and Braggg-R factor = 17.5. bThe B values of Fe


and Mn are constrained to be the same.
a

is hexagonal and adopts the P63/mmc space group. Interestingly, the two give dierent lattice parameters (a = 0.0025(2)
and c = 0.012(1) , with NPD yielding a larger a and
smaller c parameter) and dissimilar sodium occupancies (x =
0.14(2)), with the larger Na content determined from the NPD
data approaching the expected value. We note that the
elemental contrast in the NPD data is more advantageous for
understanding the structure than the XRPD data, most notably
the contrast available between the site-sharing Fe and Mn.
Perhaps more importantly, the powders were isolated from air
during the NPD experiment but not the XRPD measurement,
where moisture is known to sometimes cause the decomposition of NFMO and the partial removal of Na. The larger c
and smaller a lattice parameters are consistent with a lower
sodium content. For these reasons, the NPD result was
considered more reliable, and the rened NMFO structure is
illustrated in Figure 2.
In the P63/mmc hexagonal P2-type structure, Na ions occupy
two prismatic sites (2b and 2d, with 24(1) and 46(2)%
occupancies, respectively) that are sandwiched between layers
of hybrid FeO6 and MnO6. An SEM image (Figure 3) shows
the particle morphology and size, and reveals that the material
is composed of nanoparticles without a well-dened morphology that aggregate with particle size 200700 nm.
The electrochemical properties of the NFMO cathode were
measured in coin-cells within the 1.54.2 V range to avoid
unnecessary reactions associated with the degradation of the
carbonate-based electrolytes at high potential, as suggested in
Yabuuchi et al.33 Figure 4a shows the chargedischarge
behavior of a typical NMFO-containing coin-cell for the initial,
second, and 80th cycles at 0.1 C. The rst charge exhibits a at
plateau at around 4.07 V, whereas on the rst discharge two

Figure 1. Rietveld renement prole for the as-prepared NFMO


powder using (a) XRPD and (b) NPD data. Vertical bars are reection
markers.
3152

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials

Figure 2. Rened crystal structure of the as-prepared NFMO powder


with the unit cell shown in gray.

Figure 4. (a) Chargedischarge curves of a typical NFMO-containing


coin-cell at the initial and 80th cycle at 0.1 C and (b) the
corresponding cycle stability. The potential range was 1.54.2 V.

cathode only maintains 55% of the initial capacity at the 80th


cycle when cycled at 0.1 C. A similar decay phenomenon was
previously observed.20
The self-aggregation of nanoparticles is commonly associated
with reduced contact with the electrolyte. During the solgel
synthesis, nanoparticles without a well-dened morphology
have a tendency to self-aggregate. Further, application of a
conductive coating is expected to improve the electrochemical
performance of the NFMO cathode.34,35 Nevertheless, the
origin of the large capacity decay (83.8 mAhg1 in our work)
and relatively low energy eciency (78%) of the NFMO
cathode remain poorly understood.
XANES (see Figure 5) estimates the oxidation states of Fe
and Mn in the uncycled cell to be 3.4+ and 4.0+, respectively,
using the analysis method suggested in Berry et al.36 (Figure S2,
Supporting Information). Given that Mn4+ is electrochemically
inert and the possibility of Fe3+ oxidation in SIBs,20 the XANES
results explain why the rst charge curve of the NMFO cathode
does not exhibit the 2.2 V plateau and delivers a lower chargecapacity since only the Fe3+/Fe4+ is active and part of the Fe is
4+. Further, these results conrm that the 4.1/3.4 V plateaus
correspond to the Fe4+/Fe3+ redox couple and that it is
therefore the Mn4+/Mn3+ redox centers responsible for the 2.2/
2.0 V plateaus, in agreement with Park et al.24 We note that the
oxidation states determined using XANES are for the material
in the coin-cell before cycling where the material may have
undergone desodiation, in contrast to the Rietveld renementderived results which are for the as-prepared powder.
We investigated the structurefunction relationship that
underpins the electrochemical properties of the cathode using
operando SXRPD. Figure 6 shows selected regions of operando

Figure 3. SEM micrograph of the as-prepared NFMO powder


showing the self-aggregation of nanoparticles.

plateaus are observed at 3.40 and 2.06 V (shown in the


incremental capacity plot in Figure S1, Supporting Information).
In the second cycle, the charge and discharge curves contain
two plateau-like features at 2.15 and 4.05 V, and 3.43 and
2.07 V, respectively. The plateaus at 2.15/2.07 V correspond
to the Mn3+/Mn4+ redox couple and those at 4.05/3.43 V to
the Fe3+/Fe4+ couple, as reported in Yabucchi et al.20 It is
notable that the 2.2 V plateau presented in the second cycle is
absent in the rst charge curve. The signicant potential
dierences, especially at the 4.05/3.43 V plateaus, indicate poor
energy and cycling performance. The NMFO cathode has the
capacity of 121.3 and 184.4 mAhg1 during the rst charge and
discharge, respectively. This equates to 0.467 and 0.709 Na per
formula unit being inserted into and extracted from the
structure, forming Na0.203(Fe1/2Mn1/2)O2 and
Na0.912(Fe1/2Mn1/2)O2 at the charged and discharged states,
respectively. It is this dierence that is responsible for the
abnormal energy eciency (>100%) of the cathode. The initial
discharge capacity is comparable to that previously reported,20
and the energy density of 495 mWhg11 closely approaches
that of the commercial LiFePO4 (530 mWhg11) cathode
and exceeds that of the LiMn2O4 (450 mWhg11) cathode.
In the second cycle, the cathode delivers the reasonably high
discharge capacity of 178.3 mAhg1, although this occurs
alongside Coulombic and energy eciencies of only 89% and
78%, respectively. Moreover, as shown in Figure 4b, the
3153

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials

representative of the phase evolution during charge (desodiation) and discharge (sodiation) for that part of the battery.
The capacity of the custom coin-cell during the operando
SXRPD experiment was lower than anticipated (Figure S4,
Supporting Information), indicating that within this sampled
volume there are parts of the cathode that are inactive. The
synchrotron beam passes through a hole in the outer battery
casing, and it is well-established that this can result in poor
contact between the electrode layers and cause areas of
inactivity.37 Despite the lower amount of extracted and inserted
Na calculated from the SXRPD data, the data show denitively
the mechanism of transformations for the NFMO cathode
during the chargedischarge process.
Figures 7, 8, and 9 show the results from the single peaktting routines applied to the 002 and 100 NFMO reections.

Figure 5. Normalized XANES (a) Fe and (b) Mn K-edge spectrum of


the P2-Nax(Fe1/2Mn1/2)O2 cathode within an uncycled coin-cell. Note
that the Fe4+ data are taken from Rohde et al.32

Figure 7. Results of peak tting of the NFMO (a) 002 and (b) 100
reections illustrating changes in the 2-values (peak shifts). The
voltage prole is also presented, and the shaded regions represent twophase transitions (purple and orange), the OP4 phase (green) and the
P2-orthorhombic phase (blue).

The evolution of the 002 and 100 reections indicate changes


to the c and a (= b) lattice parameters of the P2-hexagonal
structure. The 002 reection (Figures 6a and 7a) indicates an
increase of the lattice parameter c during initial desodiation up
to 4.06 V, while the 100 reection (Figures 6b and 7b)
indicates a decrease of the a lattice parameter, with the latter
likely a results of shrinkage of the MO6 (M = Fe, Mn)
octahedra as a result of the increase in the average oxidationstate of the redox centers. The continuous c-axis expansion and
a-axis contraction during desodiation up to 4.06 V (the Fe3+/
Fe4+ redox plateau) suggests that the Mn3+/Mn4+ redox couple
proceeds through a solidsolution reaction. On further
desodiation above 4.06 V, the intensity of both the 002 and
100 reections (Figures 6 and 8) changes signicantly
alongside their shift in position. Similar changes are also

Figure 6. Contour plots of operando SXRPD data for NFMO within a


coin cell in the (a) 6.58.5 and (b) 13.516.8 2 regions.

SXRPD data during charge and discharge within 1.5 and 4.2 V
vs Na, respectively, at 0.07 mA (equivalent to 0.05 C). The
operando SXRPD data for the other regions are shown in Figure
S3 (Supporting Information). We note that the synchrotron
beam interacts with the sample in a region 2 1 mm in
transmission geometry and that the operando data are therefore
3154

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials

Figure 8. Results from peak ts of the NFMO (a) 002 and (b) 100
reections illustrating changes in the reection intensity. The voltage
prole is also presented, and the shaded regions represent two-phase
transitions (purple and orange), the OP4 phase (green) and the P2orthorhombic phase (blue).

Figure 9. Results of peak ts of the NFMO (a) 002 and (b) 100
reections illustrating changes in the peak width. The voltage prole is
also presented, and the shaded regions represent two-phase transitions
(purple and orange), the OP4 phase (green) and the P2-orthorhombic
phase (blue).

observed for other reections (Figure S3, Supporting


Information). Some reections become nearly invisible before
reappearing during the cycle. The continuous peak shift
alongside changing intensity signies concurrent solidsolution
reaction and atomic rearrangement. We consider and briey
summarize previous work investigating the NFMO cathode
phase evolution at high charge (low-sodium-content). Lu and
Dahn studied a P2Nax(Ni1/3Mn2/3)O2 cathode and suggested
that an O2-type phase with stacking faults is formed at x 1/3
during desodiation.38 Yabucchi et al. report the formation of an
OP4-type phase with a P6m2 space group after charging P2Nax(Fe1/2Mn1/2)O2 to 4.2 V as a result of gliding of the MO6
(M = Fe, Mn) slabs.20 Mortemard de Boisse et al. studied the
P2Nax(Fe1/2Mn1/2)O2 cathode with x < 0.35, suggesting a
new phase with hybrid FeO6 and MnO6 octahedra-containing
layers that exhibits a high degree of disorder.39 The order
disorder transition is thought to originate from instability of the
prismatic sodium sites. We note that part of the cathode is
unresponsive, and this inactivity, which is irreversible, arises
from the orderdisorder transition and suggests that the upper
cuto voltage may be too high for the P2Nax(Fe1/2Mn1/2)O2
cathode. Given that the amount of the nonresponsive cathode
increases with cycle, as evidenced by the increasing intensity
with time in Figure 6 (and Figure S5, Supporting Information),
the unresponsive cathode may be responsible for the capacity
decay of the cathode.
Our results reveal a dramatic change in the full-width at halfmaximum (FWHM) of the NFMO 002 and 100 reections
(Figure 9) during charging at 4.1 V or above, suggesting
increased structural disorder. Despite the disorder, we were able
to index the reections at the high-charge state (i.e., 4.2 V and

lowest sodium content) to a hexagonal OP4 phase with a P63


space group, which is disagrees with the model suggested by
Yabucchi et al. that was informed by ex situ SXRPD and XAS
data.20 We nd that the hybrid MO6 octahedra-containing
layers are maintained, which is in agreement with Mortemard
de Boisse et al.,39 during the P2-OP4 transition. Taken
together, these results suggest that the P2-OP4 transition
proceeds through a solidsolution-like two-phase reaction. The
OP4 phase exhibits a distribution of Na ions at the 6c site,
giving rise to disorder that is reduced during further
desodiation. The OP4 phase is formed via gliding of the
MO6 (M = Fe, Mn) slabs, enabled through a reduction in the
repulsion between them as prismatic (6c) sodium sites are
depopulated during the Fe3+/Fe4+ transition. This formation
mechanism is similar to that found for the P2Na
0.7CoO2O3-LiCoO2 system, where the OP4 (LixNay)CoO2 crystallizes in P63/mmc,40 and where our OP4 phase has
both sodium and vacancies () at the 6c sites and can be
considered a mixed P2NaxMO2O3-MO2 system.
The 004 and 100 (of OP4) reections shift dramatically to a
higher angle, indicating a large lattice shrinkage, with further
sodium removal. This c-axis contraction likely arises due to the
increasing average charge of the O-ions at the high-charge state,
meaning the layers are less negatively charged and that the
repulsion between them is reduced, along with the interlayer
distance.4143 Concurrently, the increasing oxidation state of
the redox centers enhances the M-O bonding and causes
further contraction of the MO6 octahedra, and consequently,
the a-axis. Nevertheless, the time evolution of the 100 reection
position is nearly linear and its rate of change during the Mn3+/
Mn4+ transition (within the P2-type structure) is smaller than
3155

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials
that during the Fe3+/Fe4+ transition (within the OP4-type
structure), further evidencing the dierent redox-active centers
at the two voltage ranges. As one can expect from the Shannon
radii of Mn (rMn3+ = 0.58 and rMn4+ = 0.53 ) and Fe (rFe3+ =
0.645 and rFe4+ = 0.565 ), the 2.2/2.0 and 4.1/3.4 V plateaus
likely arise from the Mn3+/Mn4+ and Fe3+/Fe4+ couples,
respectively.44 Upon discharge (sodiation), the reverse
reactions (OP4-P2) are observed. When the coin-cell is
discharged below 2.0 V, the P2-hexagonal phase undergoes a
two-phase reaction, transforming into a P2-type orthorhombic
phase with a Cmcm space group, the same structure reported by
Mortemard de Boisse et al.39 On the basis of the operando
SXRPD results, the phase evolution of the P2
Nax(Fe1/2Mn1/2)O2 cathode during charge and discharge
obtained from operando SXRPD is shown in Scheme 1.

(t32ge1g )) destabilize the cathode at various states-of-charge,


with the P63/mmc to Cmcm transition associated with
considerable structural anisotropy. Notably, it is the noncooperative JahnTeller distortion of Mn3+ during deep
discharge that initiates this transition.
It is likely that the poor cycling performance of the cathode
originates from this complex phase/structure evolution. These
results suggest that by narrowing the operating voltage of the
battery from 1.54.2 to 2.04.0 V the cycle stability of the
cathode may be signicantly improved, and this is corroborated
by reports of signicantly improved cycle stability in the 1.5
4.0 V21 and 2.04.2 V windows. Xu et al. hypothesized that the
better cycling performance obtained in the 1.54.0 V window
may avoid the structural distortion present in the 1.54.3 V
window, and Park et al. considered that stacking-fault eects
could be reduced by cycling within the 2.04.2 V window. We
measured the cycling performance of the P2Nax(Fe1/2Mn1/2)O2 cathode in the 1.54.0, 2.04.0, and 2.04.2 V windows at
0.1 C, with the results shown in Figure 10. Unquestionably, the

Scheme 1. Phase Evolution of the NFMO Cathode during


Desodiation and Sodiation

The phase evolution of the cathode is accompanied by


volumetric changes. Given the three dierent structures at 4.2,
2.7, and 1.5 V, we compare the volume per cathode formula
unit of the material in Table 3, noting the 12% volume change

Figure 10. (a) Discharge curves of typical NFMO-containing coincells at the initial and 80th cycle at 0.1 C when cycled in the 1.54.0,
2.04.0, and 2.04.2 V voltage windows and (b) the corresponding
cycle stability.

Table 3. Comparison of the Volume per Formula for the


NFMO Cathode at Various States
phase

P63

P63/mmc

Cmcm

structure type
volume per formula (3)

hexagonal
36.7

hexagonal
41.6

orthorhombic
42.9

P2Nax(Fe1/2Mn1/2)O2 cathode cycling performance was best


within the 2.04.0 V window, although the initial discharge
capacity was only 90 mAhg1. Within the 1.54.0 and 2.0
4.2 V windows, large capacity decays, i.e., 100.6 and 110.4 mAh
g1, are observed after 80 cycles, and these are attributed to the
formation of the Cmcm and disordered P63 phases, respectively.
Our results show that the cycling performance of the cathode
can be enhanced by controlling the phase evolution, as
informed by our SXRPD study. It is notable that, with 2.0
4.0 V window, the capacity retention rate is 69%, showing
25% better cycling performance when compared with the
1.54.0 V window. While there is a signicant reduction in
capacity within this new voltage range (37%), this loss is oset

when the P2Nax(Fe1/2Mn1/2)O2 cathode (P63/mmc) is


charged (desodiated) to form the P63 structure. The relatively
large volume change between the OP4/P2 structures is a likely
contributor to structural instability of the cathode.
Despite the only 3% volume dierence between the P63/
mmc and Cmcm phases, the transition between these two is
structurally signicant, with the symmetry changing from
hexagonal to orthorhombic during desodiation. Notably,
JahnTeller distortion of the Mn3+ and Fe4+ (both 3d4
3156

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials
by a relatively greater gain in the cycle retention (59%). We
note that this material is considered for use in not yet
commercially available sodium-ion batteries, an emerging
technology, where materials development is in its infancy.
Our results demonstrate that electrochemical properties of
electrode materials can be optimized by a detailed knowledge of
the phase and structural evolution gained using operando
structural methods such as SXRPD.

Absorption Spectroscopy beamline at the Australian Synchrotron, Victoria, Australia.

CONCLUSIONS
The P2Nax(Ni1/2Mn1/2)O2 nanoparticulate cathode has been
successfully prepared via a solgel method, and its electrochemical properties are investigated. Although the cathode
delivers a high initial capacity of 180 mAhg1 and excellent
energy-density of 495 mWhg11, the energy eciency and
cycle stability are reasonably poor within the 1.54.2 V vs Na
operating range. To understand the origin of the capacity fade
during cycling, operando synchrotron X-ray powder diraction
was used to determine the phase evolution of the cathode
during charge and discharge within this voltage window. The
work showed that the P2 hexagonal phase undergoes a solid
solution reaction during the Mn3+/Mn4+ redox transition and a
solidsolution-like two-phase reaction at 4.1 V and above,
accompanying desodiation during the Fe3+/Fe4+ transition.
Importantly, by avoiding the large voltage dierence between
the anodic and cathodic plateaus, we show that with the
alternative operating window, i.e., 2.04.0 V (vs Na), as
informed by our comprehensive understanding of the phase
evolution, the cycling performance can be improved by 25%.
This work demonstrates the importance of understanding the
structurefunction relationship of electrode materials and the
feasibility of operando synchrotron X-ray powder diraction for
studying this.

ABBREVIATIONS
LIB, lithium-ion battery; SIB, sodium-ion battery; XRPD, X-ray
powder diraction; SXRPD, synchrotron X-ray powder
diraction; NPD, neutron powder diraction; XAS, X-ray
absorption spectroscopy; SEM, scanning electron microscopy;
XANES, X-ray absorption near-edge structure; standard
reference material, SRM; NFMO, Nax(Ni1/2Mn1/2)O2

ASSOCIATED CONTENT

S Supporting Information
*

Incremental capacity plot; estimation of Fe and Mn oxidation


states; contour plot of operando synchrotron diraction data;
chargedischarge curves of during the operando SXRPD
experiment, including the capacity calculation; and the results
of peak tting of the nonresponsive NFMO 002 illustrating
changes in the intensity with time. This material is available free
of charge via the Internet at http://pubs.acs.org.

REFERENCES

(1) Nishi, Y. J. Power Sources 2001, 100, 101106.


(2) Kim, S.-W.; Seo, D.-H.; Ma, X.; Ceder, G.; Kang, K. Adv. Energy
Mater. 2012, 2, 710721.
(3) Palomares, V.; Serras, P.; Villaluenga, I.; Hueso, K. B.; CarreteroGonzalez, J.; Rojo, T. Energy Environ. Sci. 2012, 5, 58845901.
(4) Ellis, B. L.; Nazar, L. F. Curr. Opin. Solid State Mater. Sci. 2012,
16, 168177.
(5) Slater, M. D.; Kim, D.; Lee, E.; Johnson, C. S. Adv. Funct. Mater.
2013, 23, 947958.
(6) Serras, P.; Palomares, V.; Alonso, J.; Sharma, N.; Lopez del Amo,
J. M.; Kubiak, P.; Fdez-Gubieda, M. L.; Rojo, T. Chem. Mater. 2013,
25, 49174925.
(7) Delmas, C.; Fouassier, C.; Hagenmuller, P. Physica B+C 1980, 99,
8185.
(8) Kikkawa, S.; Miyazaki, S.; Koizumi, M. J. Solid State Chem. 1986,
62, 3539.
(9) Didier, C.; Guignard, M.; Denage, C.; Szajwaj, O.; Ito, S.;
Saadoune, I.; Darriet, J.; Delmas, C. Electrochem. Solid-State Lett. 2011,
14, A75A78.
(10) Guignard, M.; Didier, C.; Darriet, J.; Bordet, P.; Elkam, E.;
Delmas, C. Nat. Mater. 2013, 12, 7480.
(11) Komaba, S.; Takei, C.; Nakayama, T.; Ogata, A.; Yabuuchi, N.
Electrochem. Commun. 2010, 12, 355358.
(12) Komaba, S.; Murata, W.; Ishikawa, T.; Yabuuchi, N.; Ozeki, T.;
Nakayama, T.; Ogata, A.; Gotoh, K.; Fujiwara, K. Adv. Funct. Mater.
2011, 21, 38593867.
(13) Yoshida, H.; Yabuuchi, N.; Komaba, S. Electrochem. Commun.
2013, 34, 6063.
(14) Vassilaras, P.; Toumar, A. J.; Ceder, G. Electrochem. Commun.
2014, 38, 7981.
(15) Ma, X.; Chen, H.; Ceder, G. J. Electrochem. Soc. 2011, 158,
A1307A1312.
(16) Caballero, A.; Hernan, L.; Morales, J.; Sanchez, L.; Santos Pena,
J.; Aranda, M. A. G. J. Mater. Chem. 2002, 12, 11421147.
(17) Berthelot, R.; Carlier, D.; Delmas, C. Nat. Mater. 2011, 10, 74
80.
(18) Kim, D.; Lee, E.; Slater, M.; Lu, W.; Rood, S.; Johnson, C. S.
Electrochem. Commun. 2012, 18, 6669.
(19) Yuan, D.; Hu, X.; Qian, J.; Pei, F.; Wu, F.; Mao, R.; Ai, X.; Yang,
H.; Cao, Y. Electrochim. Acta 2014, 116, 300305.
(20) Yabuuchi, N.; Kajiyama, M.; Iwatate, J.; Nishikawa, H.; Hitomi,
S.; Okuyama, R.; Usui, R.; Yamada, Y.; Komaba, S. Nat. Mater. 2012,
11, 512517.
(21) Xu, J.; Chou, S.-L.; Wang, J.-L.; Liu, H.-K.; Dou, S.-X.
ChemElectroChem 2014, 1, 371374.
(22) Mortemard de Boisse, B.; Carlier, D.; Guignard, M.; Delmas, C.
J. Electrochem. Soc. 2013, 160, A569A574.
(23) Zhu, H.; Lee, K. T.; Hitz, G. T.; Han, X.; Li, Y.; Wan, J.; Lacey,
S.; Cresce, A. v. W.; Xu, K.; Wachsman, E.; Hu, L. ACS Appl. Mater.
Interfaces 2014, 6, 42424247.
(24) Park, K.; Han, D.; Kim, H.; Chang, W.-s.; Choi, B.; Anass, B.;
Lee, S. RSC Adv. 2014, 4, 2279822802.
(25) Pang, W. K.; Peterson, V. K.; Sharma, N.; Zhang, C.; Guo, Z. J.
Phys. Chem. C 2014, 118, 39763983.
(26) Serras, P.; Palomares, V.; Rojo, T.; Brand, H. E. A.; Sharma, N. J.
Mater. Chem. A 2014, 2, 77667779.

AUTHOR INFORMATION

Corresponding Authors

*(V.K.P.) Tel: +61 9717 9401. E-mail: vanessa.peterson@


ansto.gov.au.
*(Z.P.G.) Tel: +61 4221 5225. E-mail: zguo@uow.edu.au.
Author Contributions

W.K.P. and S.K. contributed equally to this work.

Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
We are grateful to the nancial support provided by the
Commonwealth of Australia and Automotive CRC 2020. N.S.
would like to acknowledge AINSE for support through the
Research Fellowship Scheme. The use of infrastructure and
facilities at ISEM and the UOW Electron Microscopy Centre is
gratefully acknowledged. Part of this research was undertaken
on the Powder Diraction beamline and on the X-ray
3157

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

Article

Chemistry of Materials
(27) Kimpton, J.; Gu, Q. Synchrotron Radiat. News 2014, 27, 1820.
(28) Pramudita, J. C.; Schmid, S.; Godfrey, T.; Whittle, T.; Alam, M.;
Hanley, T.; Brand, H. E. A.; Sharma, N. Phys. Chem. Chem. Phys. 2014,
16, 2417824187.
(29) Liss, K.-D.; Hunter, B.; Hagen, M.; Noakes, T.; Kennedy, S.
Physica B 2006, 385386 (Part 2), 10101012.
(30) Roisnel, T.; Rodriguez-Carvajal, J. In Materials Science Forum,
Proceedings of the Seventh European Powder Diraction Conference
(EPDIC 7); Delhez, R., Mittenmeijer, E. J., Eds.; Scitec Publications,
Ltd.,: Switzerland, 2000; pp 118123.
(31) Richard, D.; Ferrand, M.; Kearley, G. J. J. Neutron Res. 1996, 4,
3339.
(32) Rohde, J.-U.; Torelli, S.; Shan, X.; Lim, M. H.; Klinker, E. J.;
Kaizer, J.; Chen, K.; Nam, W.; Que, L. J. Am. Chem. Soc. 2004, 126,
1675016761.
(33) Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Chem. Rev.
2014, 114, 1163611682.
(34) Gu, Y.; Chen, D.; Jiao, X. J. Phys. Chem. B 2005, 109, 17901
17906.
(35) Kalluri, S.; Seng, K.; Pang, W.; Guo, Z.; Chen, Z.; Liu, H.-K.;
Dou, S. X. ACS Appl. Mater. Interfaces 2014, 6, 89538958.
(36) Berry, A. J.; ONeill, H. S.; Jayasuriya, K. D.; Campbell, S. J.;
Foran, G. J. Am. Mineral. 2003, 88, 967977.
(37) Borkiewicz, O. J.; Shyam, B.; Wiaderek, K. M.; Kurtz, C.;
Chupas, P. J.; Chapman, K. W. J. Appl. Crystallogr. 2012, 45, 1261
1269.
(38) Lu, Z.; Dahn, J. R. J. Electrochem. Soc. 2001, 148, A1225A1229.
(39) Mortemard de Boisse, B.; Carlier, D.; Guignard, M.; Bourgeois,
L.; Delmas, C. Inorg. Chem. 2014, 53, 1119711205.
(40) Berthelot, R.; Pollet, M.; Carlier, D.; Delmas, C. Inorg. Chem.
2011, 50, 24202430.
(41) Amatucci, G. G.; Tarascon, J. M.; Klein, L. C. J. Electrochem. Soc.
1996, 143, 11141123.
(42) Van der Ven, A.; Aydinol, M. K.; Ceder, G.; Kresse, G.; Hafner,
J. Phys. Rev. B 1998, 58, 29752987.
(43) Laubach, S.; Laubach, S.; Schmidt, P. C.; Ensling, D.; Schmid,
S.; Jaegermann, W.; Thissen, A.; Nikolowski, K.; Ehrenberg, H. Phys.
Chem. Chem. Phys. 2009, 11, 32783289.
(44) Shannon, R. Acta Crystallogr., Sect. A 1976, 32, 751767.

3158

DOI: 10.1021/acs.chemmater.5b00943
Chem. Mater. 2015, 27, 31503158

S-ar putea să vă placă și