Vous êtes sur la page 1sur 8

Journal of Power Sources 359 (2017) 277e284

Contents lists available at ScienceDirect

Journal of Power Sources


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

Na2MnSiO4 as an attractive high capacity cathode material for


sodium-ion battery
Markas Law, Vishwanathan Ramar, Palani Balaya*
Department of Mechanical Engineering, National University of Singapore, 117575, Singapore

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Na2MnSiO4 is prepared via a modi-


fied two-step route.
 It delivered the highest discharge
capacity for a polyanion-based
compound.
 Na2MnSiO4 registered impressive so-
dium storage performance of
210 mAh g1 at 0.1 C.
 Electrolyte additive VC forms a meta-
stable passivation film on electrode
surface.
 The amount of VC added greatly in-
fluences electrochemical perfor-
mance of Na2MnSiO4.

a r t i c l e i n f o a b s t r a c t

Article history: Here we report a polyanion-based cathode material for sodium-ion batteries, Na2MnSiO4, registering
Received 25 March 2017 impressive sodium storage performances with discharge capacity of 210 mAh g1 at an average voltage of
Received in revised form 3 V at 0.1 C, along with excellent long-term cycling stability (500 cycles at 1 C). Insertion/extraction of
12 May 2017
~1.5 mol of sodium ion per formula unit of the silicate-based compound is reported and the utilisation of
Accepted 19 May 2017
Available online 29 May 2017
Mn2þ ! Mn4þ redox couple is also demonstrated by ex-situ XPS. Besides, this study involves a systematic
investigation of influence of the electrolyte additive (with different content) on the sodium storage
performance of Na2MnSiO4. The electrolyte additive forms an optimum protective passivation film on the
Keywords:
Sodium manganese silicate
electrode surface, successfully reducing manganese dissolution.
Sodium-ion battery © 2017 Elsevier B.V. All rights reserved.
Cathode material
Energy storage
Polyanion compound

1. Introduction has soared to unprecedented heights. Large-scale energy storage


systems utilised in the micro-grids and the hybrid vehicles require
Rechargeable lithium battery technologies have been the fore- huge amounts of lithium [1], which resulted in the anxiety con-
runner in stationary and mobile energy storage, and have now cerning the shortage of lithium reserves. This issue has become the
evolved to empower green electric vehicles. Ever since its com- driving force to recognise sodium battery as a promising replace-
mercialisation in 1990, demand for the lithium battery technology ment because of the abundance and wide distribution of sodium
resources for power-grid applications where space/volume is not a
big concern [2,3]. However, energy density of this technology is
* Corresponding author. limited because of its low capacity and slow kinetics, owing to the
E-mail address: mpepb@nus.edu.sg (P. Balaya).

http://dx.doi.org/10.1016/j.jpowsour.2017.05.069
0378-7753/© 2017 Elsevier B.V. All rights reserved.
278 M. Law et al. / Journal of Power Sources 359 (2017) 277e284

relatively heavier and larger ionic radius of the sodium ion [4]. Due 2. Experimental section
to this fact, compounds which have lower molecular weight or
those which could exchange more than 1 mol of sodium per for- 2.1. Material synthesis
mula unit are crucial to achieve high energy density. Besides, ma-
terials with an open framework are essential for facile sodium ion Nanostructured pure Na2MnSiO4 is synthesised via a two-step
intercalation reaction. method. All precursors are mixed in stoichiometric ratio unless
In view of the above, significant interests have been devoted to stated otherwise. In glass bottle A, tetraethyl orthosilicate (TEOS)
sodium-based oxide compounds (NaxMO2, where M ¼ transition (C8H20O4Si; Alfa Aesar; purity 98%) is added to a well-mixed solvent
metal) because of their relatively lower molecular weight, as well as of ethanol (Merck Millipore) and Milli-Q water (2:1 vol ratio), and
open structure. Driven by high discharge capacity and high oper- stirred at room temperature for 1 h. In a separate glass bottle B,
ating potential, early contributions have been made mostly by 1 mmol of citric acid (C6H8O7; Alfa Aesar; 99þ%) is added to ethanol
Delmas et al. in the beginning of 1980s [5e7]. Very recently how- and Milli-Q water mixture (2:3 vol ratio) and stirred until citric acid
ever, Liu et al. [8] reported a new O3-type Na0.78Li0.18Ni0.25Mn0.58O2 is fully dissolved. The addition of citric acid provides an acidic
which is obtained by electrochemical Na-Li ionic exchange process medium for phase formation [22]. This is then followed by addition
of Li1.167Ni0.25Mn0.583O2. This oxide material shows the highest of manganese (II) acetate tetrahydrate (Mn(CH3COO)2$4H2O; Strem
discharge capacity of 240 mAh g1 in the voltage range 1.5e4.5 V. Chemicals; 99þ%) and further stirred for 1 h. These solutions in
Prior to that, Yabuuchi et al. [9] demonstrated the highest obtained bottle A and bottle B are then combined and stirred further at room
reversible capacity of 220 mAh g1 with their new P2-type Na2/ temperature for 24 h to ensure homogeneous mixture of the
3Mg0.28Mn0.72O2 cycled over a voltage window 1.5e4.4 V. Another chemicals.
compound is the P2-type Na2/3Fe1/2Mn1/2O2 proposed by Yabuuchi The ethanol-water solvent is evaporated using a rotary evapo-
et al. [10], which delivers 190 mAh g1 of reversible capacity in the rator (Heidolph Hei-VAP Precision ML/G3). The obtained white
voltage range 1.5e4.2 V, utilising electrochemically active Fe3þ/ precipitate is ground to powder and calcined in a tubular furnace
Fe4þ redox couple. Yet another O3-type NaNi0.5Mn0.5O2 is reported (Carbolite Limited, UK) at 700  C for 6 h under argon atmosphere.
in 2012 by Komaba et al. [11] demonstrating high discharge ca- The resultant black powder is mixed with sodium carbonate (20%
pacity of 185 mAh g1 in the voltage range 2.2e4.5 V. The revers- excess) (Na2CO3; Alfa Aesar; 98%) using a high-energy ball mill,
ibility is further improved by adding fluoroethylene carbonate into HEBM (Fritsch Planetary Micro Mill PULVERISETTE 7 premium line)
electrolyte solution. More recently, the same group [12] discovered at 500 rpm for 30 min. A 20 ml zirconium oxide grinding bowl with
a new lithium-excess P2-type Na5/6Li1/4Mn3/4O2 which delivers stainless steel casing and five 10 mm zirconium oxide balls are used
high reversible capacity of 190 mAh g1 when cycled at a voltage for the HEBM process. This is followed by addition of 1 mmol of
window 1.5e4.4 V. However, it is a known fact that oxide-based citric acid as a carbonising agent and ball milled at 500 rpm for
compounds suffer numerous safety issues, resulting from the 10 min. The sample is calcined in the tubular furnace at 750  C for
weak M-O bond which makes them unstable at too high potentials 8 h under argon. The obtained Na2MnSiO4 sample is further ball
and requires special attention using battery management system to milled at 500 rpm for 4 h with Super P carbon black (Alfa Aesar,
assure safely [1,13]. >99%) in the weight ratio 4:1 [17,24]. This sample is finally annealed
In the case of polyanion, only a handful of compounds exceeded in the tubular furnace at 750  C for 6 h under argon to release any
reversible storage capacity of 110 mAh g1 because of its relatively accumulated lattice strain during the course of ball milling [25].
higher molecular weight and sluggish kinetics [14e16], and plau-
sible failure to utilise more than 1 mol of Naþ per formula unit
[17e19]. Nonetheless, Park et al. [20] revealed a fluorophosphate- 2.2. Material characterisation
based Na1.5VPO4.8F0.7 cathode capable of delivering a discharge
capacity of 134 mAh g1 in the voltage window 2.0e4.5 V, involving Powder X-ray diffraction (PXRD) is carried out with a D8
1.2 electron exchange per formula unit (V3.8þ ! V5þ). Saravanan ADVANCE ECO X-ray powder diffractometer (Bruker, Germany)
et al. [21] prepared carbon coated Na3V2(PO4)3 cathode material over 2q angle of 10 e100 using Cu-Ka radiation operated at 40 kV
which was able to provide 116 mAh g1 discharge capacity over the and 25 mA. Rietveld refinement is performed using TOPAS version
voltage window 2.3e3.9 V involving 1.88 mol of Naþ. Although 4.2 software on the acquired high resolution PXRD data. Physical
Li2MnSiO4 compounds are well-explored for lithium-ion battery, its characteristics of synthesised sample are examined using a field-
sodium analogue has seen very little development. In 2011, Duncan emission scanning electron microscope (FESEM, JEOL JSM-7000F;
et al. [22] successfully synthesised Pn-Na2MnSiO4 by a sol-gel Japan) operated at 15 kV and 20 mA. Energy-dispersive X-ray
method. Pn-Li2MnSiO4 was obtained via ion exchange of the syn- spectroscopy (EDS) and elemental mapping are performed to
thesised Pn-Na2MnSiO4 and electrochemical performance was obtain elemental composition of the samples. Prior to FESEM and
investigated in a lithium-ion cell. However, no electrochemical EDS, the sample is sputtered with a thin platinum layer using auto
study was performed to investigate sodium storage. Chen et al. [23] fine coater (JFC-1600; JEOL, Japan) to increase surface conductivity.
synthesised Na2MnSiO4 by the sol-gel method and achieved Transmission electron microscopy (TEM, JEM-2010F; JEOL, Japan)
reversible capacity of 125 mAh g1 at 0.1 C in ionic liquid electro- operated at 20 kV is used to observe thickness of the carbon coating
lyte. Only one electron transfer was possible in this work, and around the Na2MnSiO4 particles. For the TEM analysis, the sample is
115 mAh g1 was obtained at moderate 1 C rate. dispersed in ethanol by sonication and a drop of this suspension is
To address various limitations of such previous reports, we loaded on the Cu grid and dried. Thermogravimetric analysis (TGA)
present here a polyanion-based Na2MnSiO4 framework as a po- is carried out using Discovery TGA Analyser (TA Instruments, USA)
tential cathode material with excellent storage capability involving to estimate the total carbon content in the sample. For the TGA
manganese redox couple successfully delivering ~1.5 electrons per study, 12e15 mg of sample is heated at 5  C min1 in air from room
formula unit. This compound registered the highest-ever discharge temperature to 800  C. Composition of the elements present in the
capacity so far for a polyanion-based cathode material. sample as well as Mn dissolution analysis are assessed using Op-
tima 5300 DV (PerkinElmer, USA) inductively coupled plasma
M. Law et al. / Journal of Power Sources 359 (2017) 277e284 279

optical emission spectrometry (ICP-OES) system. For elemental narrow 2q range is illustrated in Fig. S1, confirming that there are no
analysis, the sample is digested with HNO3/HCl and top up to 10 mL unidentified weak peaks in the pattern, as reported by Duncan et al.
with deionised water so as to form a precipitate. Prior to Mn [22] The Rietveld refinement with good reliability factor values
dissolution measurements, the electrodes are aged for 30 days in indicates that the synthesised Na2MnSiO4 crystallises in the
the coin cell (cells are assembled with and without 5 vol% VC) and monoclinic phase with Pn space group. The obtained lattice pa-
the electrodes are carefully retrieved by dismantling the coin cells rameters [a ¼ 7.01944 (45) Å, b ¼ 5.61377 (35) Å, c ¼ 5.33185 (33) Å,
in the glove box. The retrieved electrodes are immersed in the b ¼ 89.8085 (17) ] and atomic positions (Table S1) show close
respective electrolytes at room temperature and the electrolyte agreement to the values obtained by GGA ab initio DFT calculations
samples are taken for examination over a time interval of 5-days for (ICDD Card No. 00-055-0638) that are used as the initial structural
a total of 30 days. The Brunauer-Emmett-Teller (BET) specific sur- data for the Rietveld refinement [22]. In addition, the degree of
face area is determined using a Nova 2200e surface area analyser crystallinity calculated from Rietveld refinement is 97.8%, suggest-
(Quantachrome, USA). Sample degassing is made at 110  C for 16 h ing that the synthesised sample has high crystallinity.
under vacuum prior to the experiment. The BET surface area is The crystal structure (viewed along b-axis) generated from the
determined from nitrogen adsorption isotherm at P/P0 of 0.05e0.3. refinement is displayed in the inset of Fig. 1a (for crystal structure
The total pore volume is evaluated from adsorption region at P/P0 of viewed along a- and c-axes, refer Fig. S2). The FESEM image in
~0.99. X-ray photoelectron spectroscopy (XPS) analysis is carried Fig. 1b of the Na2MnSiO4 reveals small particles with a pseudo-
out with an AXIS Ultra DLD (Kratos Analytical Ltd, UK) at 15 kV and spherical morphology and their size falls in the nano-range
5 mA utilising a mono Al-Ka radiation. During XPS measurement, 60e90 nm. The total carbon content determined by TGA is 24 wt
working pressure of 1  109 Torr and base pressure 1  109 Torr % (Fig. S3a). The BET analysis reveals a specific surface area of
are applied. XPSPEAK software version 4.1 is used to fit XPS spectra 71.1 m2 g1 for the Na2MnSiO4 sample featuring a well-defined
with a Shirley-type background. The accuracy of binding energies Type-IV isotherm, which can be classified as mesoporous mate-
(BE) is within ±0.1 eV. rials (Fig. S5). The molar ratio (normalised to the manganese con-
tent) for the synthesised sample has a stoichiometric ratio of 2:1:1
2.3. Electrochemical characterisation for Na:Mn:Si, corresponding to the molecular formula of Na2Mn-
SiO4 (Table S2). From the TEM image in Fig. S6, it is clear that a
Electrochemical measurement on the reported system Na2Mn- uniform layer of amorphous carbon coating with thickness around
SiO4 is carried out using 2016-type coin cells. Na2MnSiO4 electrode 2e3 nm is present on the Na2MnSiO4 nanoparticles. This layer of
is prepared by mixing the as-prepared sample and polyvinylidene surface carbon coating is essential to ensure good electronic con-
fluoride (PVdF; Kynar 2801) binder in the weight ratio 9:1 with N- ductivity for efficient electron charge transfer between the parti-
methyl-2-pyrrolidone (NMP; Merck) as the solvent. No extra ad- cles. In addition, diffused rings in the SAED pattern confirm the
ditive carbon is used to the prepare slurry. The viscous slurry is amorphous nature of the carbon coating (inset to Fig. S6). Lattice
coated on a 0.015 mm-thick etched aluminium current collector fringes are clearly visible from the TEM image, confirming the high
(Shenzhen Vanlead, China) using a doctor blade and dried at 110  C degree of crystallinity obtained from Rietveld refinement of PXRD
for 6 h in a vacuum oven (Memmert, Germany). The dried electrode data.
is then cold-pressed using a roller-pressing machine (Hohsen
Corporation, Japan) before being cut into circular discs. The elec- 3.2. Sodium storage performance
trodes are dried in a vacuum antechamber at 110  C for 6 h before
the cell assembly. The geometrical area of the studied electrode is The charge-discharge profiles obtained at a current rate of 0.1 C
2.01 cm2 for the 2016-type cell and 0.95 cm2 for the Swagelok-type for the Na2MnSiO4 half-cell (using NaPF6 electrolyte in EC:PC vol-
cell. In both the cells, the electrode has a typical active material ume ratio 1:1, with 5 vol% vinylene carbonate (VC) additive) with
loading of 1.4e1.6 mg cm2. The coin cells are assembled in an the constant current-constant voltage (CCCV) mode of charging at
argon-filled glove box (MBraun, Germany) with metallic sodium as 4.3 V is presented in Fig. 2a. The 5 vol% VC used here is an optimised
the counter electrode and glass microfiber filters (Whatman; Grade content upon a systematic variation, which will be discussed later.
GF/C) as the separator. In a three-electrode setup (MTI Corporation, It can be seen that the initial sodium extraction follows a sloppy
USA), metallic sodium acts as both counter and reference elec- profile ascending to the upper cut-off voltage of 4.3 V at which a
trodes, with Na2MnSiO4 as working electrode. The electrolyte used constant voltage mode is applied for 2 h. It should be noted that the
is 1 M NaPF6 in ethylene carbonate (EC; Sigma-Aldrich)-propylene operated voltage during constant voltage mode is well within the
carbonate (PC; Sigma-Aldrich) (1:1 vol ratio) with varying amount stability window of the electrolyte [26], which means that the extra
of vinylene carbonate (VC; Sigma-Aldrich) as additive. Sodium capacity obtained is purely from the active material and is not
storage performance is analysed using a battery testing system (BT- related to any electrolyte degradations. The discharge process be-
2000 Arbin Instruments, USA) operated in galvanostatic mode be- gins at 4.3 V and decreases steeply until a shallow plateau at ~3.4 V
tween 2.0 and 4.3 V at room temperature. Constant voltage (CV) before descending to the lower cut-off voltage of 2.0 V, giving rise
mode is initiated at the end of the charge cycle at 4.3 V for 2 h. to discharge capacity of 155 mAh g1. Surprisingly, as cycling pro-
Electrochemical impedance spectroscopy (EIS) is performed over gresses, a charge plateau begins to appear at ~3.9 V resulting in an
the frequency range 10 mHze1 MHz on a three-electrode cell using increase in charge capacity. As a result, discharge capacity increases
a VMP3 multi-channel potentiostat (BioLogic Science Instruments, to 210 mAh g1 in the 10th cycle corresponding to ~1.5 mol of Naþ
France) equipped with an analysis software EC-Lab version 10.19. storage per formula unit of Na2MnSiO4 (1 mol contributes to
discharge capacity of 139 mAh g1). In order to confirm that the
3. Results and discussion change in the shape of the voltage profiles as seen in Fig. 2a is not
due to any structural changes, ex-situ XRD measurements are per-
3.1. Structural and morphological analysis formed at the end of different discharge cycles.
Fig. 3 demonstrates ex-situ PXRD patterns of Na2MnSiO4 elec-
Fig. 1a presents PXRD pattern of the as-synthesised Na2MnSiO4, trode at the end of selected discharge cycles. All the major peaks of
which clearly shows single-phase formation of Na2MnSiO4 without Na2MnSiO4 are seen, suggesting that the sample retains its struc-
any observable impurity phases. A zoomed-in PXRD pattern over ture during cycling. This confirms the fact that the change in the
280 M. Law et al. / Journal of Power Sources 359 (2017) 277e284

Fig. 1. (a) PXRD pattern, with inset showing refined crystal structure viewed along b-axis, Naþ yellow sphere, Mn2þ magenta sphere, Si4þ green sphere, O2 red sphere; and (b)
FESEM image of the synthesised Na2MnSiO4. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 2. Charge-discharge profiles cycled at 0.1 C using NaPF6 electrolyte in EC:PC v/v 1:1 (a) with 5 vol% VC and (b) without 5 vol% VC.

cycles. The obtained discharge capacity in this control experiment


is similar to a recent report on Na2MnSiO4 using ionic liquid elec-
trolyte [23]. Thus, inclusion of 5 vol% VC plays a vital role in sta-
bilising the electrode-electrolyte interface during cycling and the
effect of varying the VC content on the electrochemical perfor-
mance will be discussed in the following section.
Interestingly, high and stable discharge capacity is obtained for
the Na2MnSiO4 sample with 5 vol% VC electrolyte additive at
various current rates, as shown in Fig. 4a. Even at 5 C rate, a
discharge capacity close to 100 mAh g1 could still be achieved,
which promises this material to be a high power density cathode.
Notably, the electrode is able to regain 100% of its initial capacity
when cycling is reverted to 0.1 C after its performance at 5 C. Be-
sides the good rate performance, this material exhibits 89.5% ca-
pacity retention even after 500 cycles at 1 C (note that the capacity
retention is calculated from the 12th cycle onwards) as shown in
Fig. 4b. A similar increasing trend is observed in the initial
discharge capacity at 1 C, which will be further elaborated in the
following section. More importantly, a high average coulombic ef-
ficiency of 99.5% is achieved throughout the cycling. The impressive
rate performance and durable cyclability of Na2MnSiO4 implies
Fig. 3. Ex-situ PXRD of the Na2MnSiO4 electrode recorded at the end of different high structural integrity and robustness of the host framework,
discharge cycles.
which is manifested by the ex-situ PXRD shown in Fig. 3.

voltage profile is not due to any structural changes. It is also worth


3.3. The effect of electrolyte additive
noting that Na2MnSiO4 does not result in amorphisation during
cycling, unlike its lithium analogue, Li2MnSiO4 [27e29].
EIS is employed to understand the influence of electrolyte ad-
As a control experiment, a Na2MnSiO4 half-cell is fabricated
ditive on the sodium storage performance and formation mecha-
using NaPF6 electrolyte in EC:PC (volume ratio 1:1) without 5 vol%
nism of the solid-electrolyte interface. Impedance spectra is
VC, and its charge-discharge profile at 0.1 C is illustrated in Fig. 2b. It
recorded on Na2MnSiO4 working electrode (WE) with different VC
is clearly seen that this control experiment does not exhibit the
content (namely 0, 3, 5, 7 and 10 vol% VC) at the end of different
similar trend as described earlier with 5 vol% VC even after 20
discharge cycles. The measurements are performed using a three-
M. Law et al. / Journal of Power Sources 359 (2017) 277e284 281

Fig. 4. (a) Rate performance and (b) long-term cycling performance at 1 C of the Na2MnSiO4 sample with 5 vol% VC.

Fig. 5. Nyquist plots at end of different discharge cycles of Na2MnSiO4 WE with various amount of VC content; (a) 0 vol%, (b) 3 vol%, (c) 5 vol%, (d) 7 vol% and (e) 10 vol% VC.

electrode setup where only the impedance response of WE is the high- and medium-frequency regions. The high-frequency
measured. All measurements are performed at a fully discharged semicircle is attributed to resistance of the passivation film, while
state, when the open-circuit potential of the WE vs. reference the semicircle at the medium-frequency region refers to charge
electrode (RE) reaches equilibrium. transfer resistances across the surface film and electrode interfaces
The Nyquist plots of the Na2MnSiO4 sample with various VC for electrons and Naþ [30,31]. The size or diameter of the depressed
content at the end of different discharge cycles are shown in Fig. 5. semicircle gives the magnitude of resistance of each process. As it is
The Nyquist plots constitute mainly two depressed semicircles at more interesting to analyse how the diameter of the semicircle
282 M. Law et al. / Journal of Power Sources 359 (2017) 277e284

evolves during cycling with different amount of VC, the fitting to


the EIS Nyquist plots is purposefully avoided to prevent any
complicated analysis, which is also suggested by Burns et al. and
Aurbach et al. [32,33].
When no VC is used in the electrolyte, the size of the semicircle
remains unchanged at different charge-discharge cycles, as can be
seen in Fig. 5a. When the amount of VC is increased to 3 and 7 vol%,
a small but steady decrease in the size of the semicircle is observed
(Fig. 5b and d, respectively). Interestingly, when 5 vol% VC is added,
the diameter of the semicircle of the EIS plots decrease significantly
during cycling (Fig. 5c). When 10 vol% VC is added, a detrimental
outcome is noted in the resistances associated with the storage
processes (Fig. 5e), where the size of the depressed semicircle in-
creases drastically after each cycle.
Analogous to the impedance response, it is interesting to note
that increasing additive VC content up to 5 vol% resulted in an
improved discharge capacity at 0.1C (Fig. S7). However, as the VC
concentration is increased above 5 vol%, a decrease in discharge
capacity is noticed. In other words, although inclusion of VC up to a Fig. 6. Amount of manganese dissolution estimated by immersing the Na2MnSiO4
electrodes in respective electrolytes (with 0 vol% vs. 5 vol% VC) at room temperature
certain limit (5 vol%) improves discharge capacity and decreases
for 30 days.
cell impedance as cycling progresses (Fig. 5), excess additive
beyond 5 vol% leads to detrimental effect in cell impedance and
subsequently on sodium storage performance. Thus, the addition of surface explains the variation in cell impedance, as well as the
optimum amount of VC into 1 M NaPF6 in EC:PC electrolyte forms a degree of Mn dissolution and electrochemical performance. When
meta-stable passivation film on Na2MnSiO4 WE, resulting in less than 5 vol% VC is used (Fig. 7a and b), the passivation film
obvious decrease in the cell impedance as a function of cycle life. formed does not sufficiently wrap the entire electrode surface
The influence of VC on the surface film resistance of the WE is leading to high Mn dissolution. Conversely, when the concentration
consistent with the earlier report by Aurbach et al. [33]. Besides, the of VC is increased beyond the optimum amount of 5 vol%, the
influence of electrolyte additives on storage performance is very passivation film becomes much thicker, thus explaining the in-
well documented for lithium-ion battery [34e37]. This observation crease in the passivation film and the associated charge transfer
confirms that VC plays a crucial role in controlling the meta-stable resistances during storage processes with cycle life as evidenced
passivation film on the surface of the working electrode. We believe from the increase in size of semicircles of Nyquist plots of the
that this passivation film dissolves during the course of cycling electrode (see Fig. 5e).
which reflects the decrease in passivation film and charge transfer Please note that presently, the actual reason for the observed
resistances (as demonstrated by a decrease in the diameter of the increase in both the charge and discharge capacities when 5 vol%
semicircle in the EIS plots), and the degree of dissolution depends VC is added is still unclear (as seen in Figs. 2a and 4). However, we
largely on the amount of VC. On the other hand, no such observa- are claiming that the reason leading to this increase in capacity
tion is made from the EIS measurement of the sodium counter might be due to a significantly suppressed manganese dissolution,
electrode (CE) (as seen in Fig. S8) for the 5 vol% VC, which clearly which is caused by an optimised passivation layer formed on the
reveals that the impedance changes arise only from the passivation electrode surface. Nonetheless, more in-depth analysis has to be
film occurring on the surface of the Na2MnSiO4 WE. conducted to fully understand the role of this passivation layer.
Another possible explanation for the observed increase in storage
3.4. Manganese dissolution capacity in the first 10 cycles is that the liquid electrolyte might
penetrate slowly into the bulk of the mesoporous electrode mate-
Transition metal dissolution can be suppressed by creating op- rial investigated here as cycling progresses, thus promoting more
timum protective surface film on the electrode surface using elec- electrochemically active sites with cycling [41,42].
trolyte additives [38,39]. This technique is proven to be one of the
most effective methods to improve the electrode-electrolyte
interface stability [40]. To understand the nature of the surface 3.5. Redox reaction of Mn2þ during cycling
passivation film, Mn dissolution test is carried out on an aged
electrode using ICP-OES analysis. The observed high discharge capacity of over 200 mAh g1 is
Fig. 6 represents the amount of Mn dissolution in the respective mostly associated with the transition of Mn redox species to higher
electrolytes (with 0 vol% and 5 vol% VC) of the Na2MnSiO4 elec- oxidation state (Mn4þ). The reversible compositional changes of
trodes over an interval of 5 days for a total of 30 days. It can be seen Na2MnSiO4 ! Na2xMnSiO4 (0  x  2) is expected during charge-
that the Mn dissolution in the Na2MnSiO4 electrodes are signifi- discharge cycles, where Mn is oxidised from þ2 to þ4 during
cantly suppressed with 5 vol% VC in electrolyte as compared to charge, and reduced from þ4 to þ2 during discharge. In order to
electrolyte without VC. Less Mn dissolution of this electrode in confirm such a redox reaction, ex-situ XPS spectra are recorded on a
electrolyte over time for 5 vol% of VC additive results in good rate fresh, 20, 40, 50, 60, 80 and 100% state-of-charge (SoC) electrodes,
performance (Fig. 4a) and stable long-term cycling (Fig. 4b). Be- and a fully discharged electrode. It should be noted that the fittings
sides, the battery is expected to have greater shelf life. yield a c2 value of less than 0.3, which indicates a good fit. From
Based on the analyses of impedance measurement on electrodes Fig. 8a, Mn in the fresh electrode has binding energy (BE) of
upon cycling and Mn dissolution on aged electrodes, we present 640.2 eV, corresponding to the BE of Mn2þ 2p3/2 [43e47]. There is
schematic drawings of formation of the passivation film as a no Mn3þ or Mn4þ species detected at this stage. At a 20% SoC,
function of VC content as shown in Fig. 7. It is believed that the deconvolution of the spectra results in two peaks located at 640.2
difference in the thickness of passivation film on the electrode and 641.2 eV, which belong to Mn2þ 2p3/2 and Mn3þ 2p3/2,
M. Law et al. / Journal of Power Sources 359 (2017) 277e284 283

Fig. 7. Schematic illustrations of formation of the passivation film with (a) 0 vol%, (b) 3 vol%, (c) 5 vol%, (d) 7 vol% and (e) 10 vol% VC.

Fig. 8. Ex-situ XPS spectra of Mn2þ/Mn3þ/Mn4þ redox couples obtained on electrode at (a) fresh, 20%, 40% and 50% SoC and (b) 60%, 80% and 100% SoC and 100% DoD.
284 M. Law et al. / Journal of Power Sources 359 (2017) 277e284

respectively [43,46,47]. Upon continuation of charge, the Mn3þ (2013) 2067.


[3] B.L. Ellis, L.F. Nazar, Curr. Opin. Solid State Mater. Sci. 16 (2012) 168e177.
peak is seen to increase in intensity at the expense of the Mn2þ
[4] M.D. Slater, D. Kim, E. Lee, C.S. Johnson, Adv. Funct. Mater. 23 (2013) 947e958.
peaks until a 60% SoC (Fig. 8b). The Mn2þ peaks disappear [5] C. Delmas, J.-J. Braconnier, C. Fouassier, P. Hagenmuller, Solid State Ion. 3e4
completely as charge progresses, and an onset of Mn4þ peaks are (1981) 165e169.
detected at 642.2 eV corresponding to Mn4þ 2p3/2 [43,47,48]. The [6] J. Molenda, Claude Delmas, P. Hagenmuller, Solid State Ion. 9e10 (1983)
431e435.
intensity of this peak increases until a 100% SoC, where the co- [7] A. Mendiboure, C. Delmas, P. Hagenmuller, J. Solid State Chem. 57 (1985)
existence of both Mn3þ and Mn4þ peaks are observed. Upon a 323e331.
100% DoD, all Mn3þ and Mn4þ species are fully reduced to Mn2þ. [8] H. Liu, J. Xu, C. Ma, Y.S. Meng, Chem. Commun. 51 (2015) 4693.
[9] N. Yabuuchi, R. Hara, K. Kubota, J. Paulsen, S. Kumakura, S. Komaba, J. Mater.
We recognise the fact that the XPS experiment is surface- Chem. A 2 (2014) 16851.
sensitive and the results represent information only on surface [10] N. Yabuuchi, M. Kajiyama, J. Iwatate, H. Nishikawa, S. Hitomi, R. Okuyama,
processes during battery operation. Careful fitting of the XPS data R. Usui, Y. Yamada, S. Komaba, Nat. Mater. 11 (2012) 512.
[11] S. Komaba, N. Yabuuchi, T. Nakayama, A. Ogata, T. Ishikawa, I. Nakai, Inorg.
reveals that Mn ions on the surface are oxidised from Mn2þ to Mn4þ Chem. 51 (2012) 6211e6220.
at the end of the charge process and reduced to Mn2þ at the end of [12] N. Yabuuchi, R. Hara, M. Kajiyama, K. Kubota, T. Ishigaki, A. Hoshikawa,
the discharge process. As a result, charge-discharge processes S. Komaba, Adv. Energy Mater. 4 (2014) 1301453.
involve multiple moles of Naþ, which is confirmed by utilisation of
[13] D. Kundu, E. Talaie, V. Duffort, L.F. Nazar, Angew. Chem. Int. Ed. 54 (2015),
3431e3344.
~1.5 mol of Naþ during galvanostatic cycling. This explains the [14] J.M. Clark, P. Barpanda, A. Yamada, M.S. Islam, J. Mater. Chem. A 2 (2014)
impressive discharge capacity of 210 mAh g1 achieved for the 11807.
[15] P. Barpanda, G. Liu, C.D. Ling, M. Tamaru, M. Avdeev, S.-C. Chung, Y. Yamada,
Na2MnSiO4 compound in this study.
A. Yamada, Chem. Mater. 25 (2013) 3480.
[16] P. Barpanda, G. Oyama, S.-i. Nishimura, S.-C. Chung, A. Yamada, Nat. Commun.
4. Conclusions 5 (2015) 4358.
[17] M. Law, V. Ramar, P. Balaya, RSC Adv. 5 (2015) 50155.
[18] H. Zou, S. Li, X. Wu, M.J. McDonald, Y. Yang, ECS Electrochem. Lett. 4 (2015)
In summary, a silicate-based, Na2MnSiO4 compound is proposed A53.
as an excellent cathode material for sodium-ion battery registering [19] Y. Zhong, Z. Wu, Y. Tang, W. Xiang, X. Guo, B. Zhong, Mater. Lett. 145 (2015)
impressive discharge capacity of 210 mAh g1, highest amongst 269.
[20] Y.-U. Park, D.-H. Seo, H.-S. Kwon, B. Kim, J. Kim, H. Kim, I. Kim, H.-I. Yoo,
those reported so far for polyanion framework. Rate performance K. Kang, J. Am. Chem. Soc. 135 (2013) 13870.
shows that this material is able to deliver a reversible discharge [21] K. Saravanan, C.W. Mason, A. Rudola, K.H. Wong, P. Balaya, Adv. Energy Mater.
capacity close to 100 mAh g1 at 5 C. In addition, it exhibits 3 (2013) 444.
[22] H. Duncan, A. Kondamreddy, P.H.J. Mercier, Y.L. Page, Y. Abu-Lebdeh,
excellent long-term cycling performance at 1 C, achieving capacity M. Couillard, P.S. Whitfield, I.J. Davidson, Chem. Mater. 23 (2011) 5446e5456.
retention close to 90% after 500 cycles. Electrochemical perfor- [23] C.-Y. Chen, K. Matsumoto, T. Nohira, R. Hagiwara, Electrochem. Commun. 45
mance of this compound is found to be extremely dependent on (2014) 63e66.
[24] V. Ramar, P. Balaya, Phys. Chem. Chem. Phys. 15 (2013) 17240.
additive VC content; the better performance being obtained at 5 vol [25] N. Yabuuchi, M. Sugano, Y. Yamakawa, I. Nakai, K. Sakamoto, H. Muramatsu,
% VC. It is seen that adding an optimum amount of VC in the S. Komaba, J. Mater. Chem. 21 (2011) 10035.
electrolyte forms a meta-stable passivation film on Na2MnSiO4, [26] A. Ponrouch, E. Marchante, M. Courty, J.-M. Tarascon, M.R. Palacin, Energy
Environ. Sci. 5 (2012) 8572e8583.
which decreases the cell impedance as a function of cycle number.
[27] V. Ramar, P. Balaya, J. Power Sources 306 (2016) 552e558.
Through chemical analysis of the electrolyte, it is found that the [28] M. Moriya, M. Miyahara, M. Hokazono, H. Sasaki, A. Nemoto, S. Katayama,
surface film provides a passivation on the electrode surface, Y. Akimoto, S.-i. Hirano, J. Electrochem. Soc. 161 (2014) A97eA101.
resulting in significantly suppressed Mn dissolution over time. Ex- [29] R. Dominko, J. Power Sources 184 (2008) 462e468.
[30] A. Rudola, D. Aurbach, P. Balaya, Electrochem. Commun. 46 (2014) 56e59.
situ XPS analysis reveals that Mn2þ/Mn3þ/Mn4þ is the active redox [31] D. Aurbach, J. Power Sources 89 (2000) 206e218.
species during charge and discharge cycles, involving utilisation of [32] J.C. Burns, R. Petibon, K.J. Nelson, N.N. Sinha, A. Kassam, B.M. Way, J.R. Dahn,
~1.5 mol of Naþ per formula unit of Na2MnSiO4 cathode material. J. Electrochem. Soc. 160 (2013) A1668eA1674.
[33] D. Aurbach, K. Gamolsky, B. Markovsky, Y. Gofer, M. Schmidt, U. Heider,
The superior electrochemical performance obtained in this study is Electrochimica Acta 47 (2002) 1423e1439.
due to the interdependent effect of both the structural stability of [34] J. Li, W. Yao, Y.S. Meng, Y. Yang, J. Phys. Chem. C 112 (2008) 12550e12556.
the Na2MnSiO4 electrode, as well as the optimised electrolyte and [35] K.S. Kang, S. Choi, J.-H. Song, S.-G. Woo, Y.N. Jo, J. Choi, T. Yim, J.-S. Yu, Y.-
J. Kim, J. Power Sources 253 (2014) 48e54.
additive mixture. [36] E.-G. Shim, T.-H. Nam, J.-G. Kim, H.-S. Kim, S.-I. Moon, J. Power Sources 172
(2007) 901e907.
Acknowledgements [37] T.-H. Nam, E.-G. Shim, J.-G. Kim, H.-S. Kim, S.-I. Moon, J. Electrochem. Soc. 154
(2007) A957eA963.
[38] C.M. Julien, A. Mauger, K. Zaghib, H. Groult, Inorganics 2 (2014) 132e154.
Markas thanks NUS, Singapore for research scholarship. PB [39] J. Wang, Q. Zhang, X. Li, Z. Wang, K. Zhang, H. Guo, G. Yan, B. Huang, Z. He,
thanks NUS for funding sodium battery research (WBS No. R-261- Electrochem. Commun. 36 (2013) 6e9.
[40] K. Xu, Chem. Rev. 104 (2004) 4303e4418.
510-001-646). The authors thank MoE, Singapore (WBS: R-265-
[41] R. Dominko, M. Bele, M. Gaberscek, M. Remskar, D. Hanzel, J.M. Goupil,
000-510-112) and EMA, Singapore (WBS No. R-265-000-568-279) S. Pejovnik, J. Jamnik, J. Power Sources 153 (2006) 274e280.
for funding support. [42] C. Sun, S. Rajasekhara, J.B. Goodenough, F. Zhou, J. Am. Chem. Soc. 133 (2011)
2132e2135.
[43] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson,
Appendix A. Supplementary data R.S.C. Smart, Appl. Surf. Sci. 257 (2011) 2717e2730.
[44] D. Choi, J. Xiao, Y.J. Choi, J.S. Hardy, M. Vijayakumar, M.S. Bhuvaneswari, J. Liu,
Supplementary data related to this article can be found at http:// W. Xu, W. Wang, Z. Yang, G.L. Graff, J.-G. Zhang, Energy Environ. Sci. 4 (2011)
4560e4566.
dx.doi.org/10.1016/j.jpowsour.2017.05.069. [45] V. Ramar, P. Balaya, Phys. Chem. Chem. Phys. 15 (2013) 17240e17249.
[46] Y. Liu, Q. Wang, Y. Gao, J. Pan, M. Su, B. Hai, G. Zhu, S. Liu, J. Alloys Compd. 646
References (2015) 112e118.
[47] M. Kuezma, S. Devaraj, P. Balaya, J. Mater. Chem. 22 (2012) 21279.
[48] L. Yunjian, L. Xinhai, G. Huajun, W. Zhixing, H. Qiyang, P. Wenjie, Y. Yong,
[1] S.-W. Kim, D.-H. Seo, X. Ma, G. Ceder, K. Kang, Adv. Energy Mater. 2 (2012)
J. Power Sources 189 (2009) 721e725.
710e721.
[2] S.Y. Hong, Y. Kim, Y. Park, A. Choi, N.-S. Choi, K.T. Lee, Energy Environ. Sci. 6

Vous aimerez peut-être aussi