You are on page 1of 8

Journal of Power Sources 452 (2020) 227832

Contents lists available at ScienceDirect

Journal of Power Sources


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

Chemically sodiated ammonium vanadium oxide as a new generation


high-performance cathode
Ananta Sarkar, Sagar Mitra *
Electrochemical Energy Laboratory, Department of Energy Science and Engineering, Indian Institute of Technology Bombay, Powai, Mumbai, 400076, Maharashtra, India

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

� Chemically sodiated ammonium vana­


date cathode (SNVO) for SIBs.
� Enhancement of thermal and electro­
chemical stability.
� Nominal potential enhancement from
2.1 to 2.8 V.
� Doped-SNVO cathode provides an
extremely high energy density of 770
Wh kg-1.
� High energy density (220 Wh kg-1) full-
cell prototype (Zr-SHNVO-3) // H-
Na2Ti3O7.

A R T I C L E I N F O A B S T R A C T

Keywords: Sodium-ion battery technologies must have low cost, safe, and high energy density cathode with prolonging cycle
Sodiated ammonium vanadium oxide life propositions for any commercial utilizations. Herein, we present for the first time an extremely high energy
Cathode density (770 W h kg 1) and excellent electrochemical performance of a sodiated and doped ammonium vana­
High energy density
dium oxide (Zr-SNVO) cathode. The cathode delivers a high discharge capacity of 255 mAh g 1 with 95% ca­
Full-cell
Sodium-ion battery
pacity retention after 100 cycles at a current rate of 200 mA g 1. Furthermore, the present report demonstrates
the full-cell performance with our cathode against a hydrogenated doped sodium titanium oxide (Zr-HNTO)
anode, capable of delivering a high energy density of 220 W h kg 1 (both active mass of cathode and anode are
considered). The present study shows a way to produce a high-performance rechargeable sodium-ion battery full-
cell using alternative cathode and anode with a comparable energy density of current lithium-ion technology.
The present study gives us a direction to make such cells for future renewable energy storage systems.

1. Introduction competitive cycling performance) compare to any other transition metal


oxides [6–10]. As vanadium oxides have a large interlayer spacing,
In recent times, transition metal-based layered oxides became a state which accumulates more number of sodium-ions that facilitate the
of arts for higher energy density and safe cathode materials for sodium- diffusion process of larger size sodium-ion into/from the host matrix
ion battery [1–5]. Among all, vanadium-based oxide materials have [11,12]. Besides, the electronic behavior of most of the vanadium ox­
been attracted attention due to its superior redox behavior (wide range ides, changes from an insulator to semiconductor or metallic character,
of oxidation state, higher specific capacity, better rate capability, and during the intercalation process at normal temperature (273–333 K)

* Corresponding author.
E-mail address: sagar.mitra@iitb.ac.in (S. Mitra).

https://doi.org/10.1016/j.jpowsour.2020.227832
Received 12 September 2019; Received in revised form 13 December 2019; Accepted 31 January 2020
Available online 2 March 2020
0378-7753/© 2020 Elsevier B.V. All rights reserved.
A. Sarkar and S. Mitra Journal of Power Sources 452 (2020) 227832

may help to enhance the rate capability of the electrode [13,14]. the highest specific capacity (375 mAh g 1 at 100 mA g 1 current
Many vanadium-based materials like V2O5, NH4V4O10, Na3V2(PO4)3, density) among all other reported cathode materials existing in
Na3V2(PO4)3Fx, VOPO4, etc. have been extensively studied in recent sodium-ion battery technology with an efficient average Coulombic ef­
years as battery cathode materials [9,15–18]. Among all, Na3V2(PO4)3, ficiency of 99.5%. We improved the energy density of about 1.5 times of
Na3V2(PO4)3F, Na2V6O16.nH2O, and VOPO4 cathodes showed a high the cathode material by metal doping. However, these cathode materials
rate and stable cycling performances against sodium metal, however, is having inherent lower potential operation characteristics leads to low
limited by lower specific capacity (nearly 120 mAh g 1) [16,19–21]. energy density and more importantly it is a sodium-free cathode causes
Whereas, the other vanadium cathode materials like V2O5, Na2VTi problem in battery manufacturing process.
(PO4)3 and NaxV2O5⋅nH2O have a higher specific capacity but an issue of Therefore, it is necessary to enhance our cathode potential and need
either lower nominal potential or/and poor cycling stability [22–24]. to incorporate sodium-ion into the cathode material during the chemical
Recently, we investigated ammonium vanadium oxide (NH4V4O10) synthesis process. Here, we introduce a chemically sodiated NH4V4O10
based cathode material for a sodium-ion battery. It provided a discharge (SNVO) cathode materials for SIB first time in literature. The same hy­
capacity of 225 mAh g 1 at a higher current density of 200 mA g 1 [25]. drothermal procedure was followed to prepare SNVO materials; only
We further enhanced the specific capacity and rate capability of the sodium-containing salt was added to this process. By this process, we are
electrode by zirconium metal doping to it [26]. This cathode provides not only incorporate the sodium source into the cathode materials and

Fig. 1. Structural and morphological characterizations of as prepared sodiated ammonium vanadium oxide (SNVO-n); (a) X-ray diffraction (XRD) pattern of different
SNVO-n cathode materials with corresponding JCPDS card no. 035–0075 showed beyond 4 mol sodiation, the monoclinic structure completely converted to different
phase, (b) characteristic XRD pattern of SNVO-3 analyzed by commercial FullProf Suite software, (c) Crystal structure of chemically pre-sodiated NVO (SNVO-3),
yellow Na, red O, blue NHþ 4 , brown V, (d) FEG-SEM image of SNVO-3, (e) FEG-TEM image of SNVO-3 materials showed nanobelts morphology with an average length
of 2–3 μm and width 300–600 nm, (f) HR-TEM image of SNVO-3 indicating the interlayer d spacing is 0.355, 0.203 nm corresponds (110) and ( 205) plane
respectively and inset image shows SEAD pattern revealed single-crystalline pattern of SNVO-3 cathode material, (g,h) In-situ high-temperature XRD of SNVO
electrode material at 5 � C min 1 heating rate under Ar atmosphere; (g) NH4V4O10 (NVO) and (h) SNVO material. (For interpretation of the references to colour in this
figure legend, the reader is referred to the Web version of this article.)

2
A. Sarkar and S. Mitra Journal of Power Sources 452 (2020) 227832

also improved the overall potential and exhibited improved electro­ and SNVO-6, these 2D particles were becoming the 3D materials. The
chemical performances. average dimension of SNVO-5 material was observed as length 2.5–6
μm, width 500 nm and height ca. 200 nm (Figs. S1i and j). However,
2. Results and discussion SNVO-6 shape is converted to irregular shapes with 1–1.3 μm average
diameter (Fig. S1k and l). Fig. 1f and Figs. S2a–d showed the HR images
To construct a sodium-ion full-cell battery, either cathode or anode of SNVO-3 at different magnifications. Figs. S2b and c further indicates
should have a sodium source in its matrix. The electrochemical pre- that the preferred growth direction of the SNVO-3 particles were in
sodiation process is not an obvious choice to make it commercially (001) plane. Here, the inter planer d spacing was estimated as ~0.353
viable. The energy density is considered one of the significant charac­ nm corresponds to (110) plane, whereas 0.203 nm inter planer d spacing
teristic parameters for the practical application, and it should be as high represents the ( 205) plane (Fig. S2d). Fig. 1f (inset) showed a single
as possible. The overall energy density of a battery can be improved by crystal selected area electron diffraction (SAED) pattern of SNVO-3
(1) enhancing the specific capacity of the active materials, and (2) sample which also matched with the corresponding XRD data. To
increasing the cell voltage because most of the consumer electronics know the composition of as-prepared SNVO-3 materials, SEM-EDS
items and ICS are powered above 3 V. In this manuscript; we attempt to pattern was shown in Fig. S2e. It shows the presence of Na, V, N, O,
enhance the overall energy density of the battery by improving the and very minute amounts of F in the materials. ICP analysis was done to
nominal potential of the cell by sodiated our cathode (NH4V4O10) ma­ find out the final product for each level sodiation process as shown in
terial by a simple chemical process. Table S1. The final formula of SNVO-1, SNVO-2, SNVO-3, SNVO-4, and
Herein, we adopted a single-step hydrothermal process to make SNVO-5 materials are Na0.28NH4V4O10, Na0.40NH4V4O10,
sodiated NH4V4O10 cathode material. The different concentration of Na0.74NH4V4O10, Na0.96NH4V4O10, Na1.12NH4V4O10 respectively.
sodiated NH4V4O10 (SNVO-n) samples were synthesized by adding To know the thermal stability of different ammonium vanadium
different mole concentration of sodium salt in the precursor during the oxides, in-situ high-temperature XRD patterns, and thermogravimetric
hydrothermal process at 190 � C for 6 h (the synthesis procedure was analysis (TGA) data were analyzed very carefully (Fig. 1g,h and Fig. S3).
discussed in details in the experimental section in Supporting Informa­ Sodiated NH4V4O10 is thermally stable under the nitrogen atmosphere
tion). SNVO-n symbol means n-mole sodium-containing salt was added till 700 � C where pure NH4V4O10 (NVO) and zirconium-doped
to 1 mol of V2O5 precursor by an in-situ process (where n ¼ 1, 2, 3, 4, 5, NH4V4O10 (Zr-NVO) materials are stable up to 350 � C in the same
6). X-ray diffraction pattern (XRD) of SNVO-n at different mole con­ environment. Initial decomposition of the materials is attributed to the
centration with corresponding JCPDS file no 031–0075 was shown in decomposition of absorbed solvent molecules like water, sodium hy­
Fig. 1a. Fig. 1a showed that with chemical sodiation, the major peaks at droxide, and ammonium hydroxide for SNVO-3 (325 � C) and water and
around 9.3, 49.8, and 64.6� of two-theta value were shifted a little to the ammonium hydroxide for Zr-NVO and NVO (bellow 100 � C) shown in
higher two theta value, indicating the shrinkage of the interlayer Fig. S3 [29–31]. After 350 � C, NVO and Zr-NVO cathode materials start
d spacing (Fig. 1a). When sodium-ion is intercalated during the chemical to decompose and form V2O5, NH3, and water (12.5% weight loss due to
pre-sodiation process, strong interaction builds between the sodium-ion evaporation of water and ammonia) [29,31]. However, SNVO-3 mate­
and vanadium layers, and that reduces the interlayer spacing. As Fig. 1a rials are stable up to 650 � C (only 5.7% weight degradation) (Fig. S3).
also indicated that with increasing the sodiation concentration above 4 Similar phenomena also observed in (in-situ high-temperature) XRD
mol, new peaks were appeared at ca. 18 and 21� of two theta scale and at experiment of different NVO materials (Fig. 1g and h). After 350 � C,
6-mol sodiation concentration (SNVO-6), monoclinic NH4V4O10 (NVO) (001) main characteristic peak of NVO was disappeared (as the material
phase was converted to a different phase (mixed phase of monoclinic decomposed and ammonia is coming out at this temperature), means
NH4V4O10 and NaxV2O5) [27]. Fig. 1a also described that (111) to (400) monoclinic NVO material was converted to a different phase. The
plane intensity ratio was increased with sodiation (preferable sodium possible thermal decomposition reaction is 2NH4V4O10 → 2NH3 þ
diffusion). The lattice parameters of as prepared SNVO-3 sample were 4V2O5 þ H2. However, this monoclinic phase of SNVO-3 sample was
evaluated by FullProf Suite software shown in Fig. 1b [28]. The evalu­ sustained up to 600 � C (after 650 � C V2O5 itself start to decompose).
ated lattice parameters of SNVO-3 are a ¼ 11.698 Å, b ¼ 3.690 Å, c ¼ During the chemical sodiation process, Naþ intercalated into the NVO
9.735 Å, α ¼ 90.0, β ¼ 101.121 and γ ¼ 90.0 with space group C12/m1 layers as well as in the gallery space whereas some of the NHþ 4 (ions) are
significantly changed from the standard JCPDS data i.e. a ¼ 11.71 Å, b ¼ attached with the oxygen by weak van der Waals attraction (Fig. 1c) and
3.76 Å, c ¼ 9.85 Å, α ¼ 90.0, β ¼ 101.0 and γ ¼ 90.0. Here, the lattice as a result, the thermal stability of the SNVO-3 materials was improved.
parameters a, b, c values are reduced which signify the contraction of the However, in NVO, NHþ 4 (ions) are occupying the gallery space which was
lattice due to the increase of attraction forces between the atoms. This easily decomposed at 350 � C (Fig. 1g and h).
indicates that the individual Naþ intercalation into the matrix. Here, the The X-ray photo absorption spectroscopy (XPS) was shown in Fig. 2
growth of the materials preferred in z-axis and sodium-ion are interacted to analyze the effect of the sodiation process on the individual compo­
in b-direction as well as in the gallery space. Some of the NHþ 4 ions are nent’s oxidation state of various SNVO-n. Fig. 2a–d showed the vana­
connected with O atoms by weak Van Der Wall bond, and others occupy dium oxidation state of various SNVO-n, which revealed that vanadium
the gallery space (Fig. 1c) [12,26]. was in the mixed-phase of V4þ of V5þ. The vanadium oxidation state of
The morphology of as-prepared materials of different mole concen­ pristine (NVO) and zirconium-doped NH4V4O10 (ZrNVO) showed the
tration SNVO-n was analyzed by the field emission gun scanning elec­ mixed phase of V5þ and V4þ in 3:1 ratio (shown in our previous studies)
tron microscopy (FEG-SEM) and field emission gun transmission [25,26]. Fig. 2a showed the vanadium oxidation state of the SNVO-1
electron microscopy (FEG-TEM). Fig. S1 indicates nanorod types sample, and the V5þ and V4þ 2p3/2 peaks are observed at 516.94 eV
morphology for all sodiated NH4V4O10 cases, except SNVO-6. Fig. S1 and 515.50 eV, and the ratio of the peak is ~2.60:1 (estimated by the
clearly shows that with increasing the sodiation concentration, the area under the curve method). Whereas for SNVO-2 sample, the peaks
particle size of the nanorods were increased. The average length of the were noticed at 516.96, and 515.52 eV corresponds to V5þ and V4þ
SNVO-1 particles (Figs. S1a and b) was observed around 500 nm to 1.2 oxidation state respectively, and the ratio was estimated ~1.66 : 1
μm and width was ~100 nm, whereas SNVO-2 particles length (Figs. S1c (Fig. 2b). However, for SNVO-3, SNVO-4 electrodes, the 2p3/2 peaks of
and d) was ~ 2–3 μm and width was about 150–200 nm. With a further V5þ and V4þ were noticed at 516.97 and 515.55 eV; 516.94 and 515.64
increase in the sodiation concentration to 3 mol (SNVO-3), the particle- eV respectively, and the peak ratio of V5þ to V4þ was further decreased
sized was enlarged to 500 nm in width and 2.5–3.5 μm in length, shown to 1.51:1 and 1.32:1 respectively (Fig. 2c and d). It was also noticed that
in Fig. 1d,e. Whereas, the particles size of SNVO-4 further increased to with increasing sodiation concentration in the composition, V5þ to V4þ
above 3–4 μm, observed in Figs. S1g and h. But in the case of SNVO-5 ratio was decreased, and both V5þ and V4þ peaks position shifted to the

3
A. Sarkar and S. Mitra Journal of Power Sources 452 (2020) 227832

Fig. 2. X-ray photo absorption spectroscopy (XPS) analysis of as prepared different SNVO-n electrode; (a) vanadium V2p spectrum of SNVO-1, (b) vanadium
spectrum of SNVO-2, (c) vanadium spectrum of SNVO-3, (d) vanadium spectrum of SNVO-4; XPS spectrum of vanadium of all different SNVO-n suggested the
coexistence of V4þ and V5þ phase in the materials and ratio of V5þ to V4þ was decreased, (e–h) O1s spectrum of different SNVO-n; (e) SNVO-1, (f) SNVO-2, (g) SNVO-
3, (h) SNVO-4; O1s spectra also suggested that V5þ to V4þ state was decreased with sodiation concentration increased, (i) Na1s spectrums of different SNVO-n, (j)
nitrogen absorption peaks of different SNVO-n, (k) fluorine spectrums of different SNVO-n suggested the absence of F1s peak in SNVO-1, SNVO-2 and SNVO-3 but
intense spectrum was observed in SNVO-4 and (l) survey scan of SNVO-3 confirmed the presence of individual elements vanadium, sodium, oxygen, and nitrogen.

higher binding energy state. As more and more sodium was inserted into limit. Further, sodium (Na1s) peak position at 1070.92 eV was shifted
NH4V4O10 matrix, more no of V5þ was converted to V4þ state, and as a towards lower binding energy and became wider with more sodiation
result, V5þ to V4þ oxidation state ratio was decreasing from 3 to 1.32 (Fig. 2i). Fig. 2j showed N1s oxidation peak was a slightly moved to­
(NVO to SNVO-4) which again confirmed the sodium-ion insertion into wards the lower binding energy from its original position (at 400.80 eV)
the matrix by the chemical process. A small V5þ satellite peak is also after more sodium insertion into NVO. Generally, when sodium is
noticed in V2p spectrum for all SNVO-n materials. For further confir­ inserted into a layered oxide, both N1s and Na1s peak intensity increases
mation, the high-resolution XPS of SNVO-n electrodes of oxygen (O1s) and it becomes broader, and the peak position shifted towards the lower
spectra were recorded as shown in Fig. 2e–h. It was noticed that O1s binding energy with more number of sodium-ion insertion [34]. A sharp
peaks divided into two prominent peaks after the sodiation (only single fluorine peak at 684.72 eV binding energy was observed for the SNVO-4
O1s peak observed for NH4V4O10 with doped or without doped; Fig. S4) sample. However, no significant F1s peak was observed for SNVO-1,
process. The O1s peak at lower binding energy ~529.50 eV is due to the SNVO-2, and SNVO-3 samples (Fig. 2k), and that indicates that there
bonding of oxygen with higher vanadium oxidation (V5þ) state, and O1s is no presence of fluorine in SNVO-1, SNVO-2, and SNVO-3 electrode
peak at higher binding energy ~ 532.0 eV represents the oxygen and reconfirms again the chemical sodiation process. The survey scan of
bonding to the vanadium having lower oxidation state (V4þ) of SNVO-n SNVO-3 presented in Fig. 2l and the scan confirmed the all elements like
(Fig. 2e–h) [32,33]. N, V, O, and Na are present in SNVO-n matrix but no fluorine. The
The O1s peak at lower binding energy was noticed at 529.49 eV, detailed XPS fitting parameters of all SNVO-n samples were evaluated to
529.44 eV, and 529.52 eV, whereas the higher energy O1s peaks know the phase of the materials accurately and tabulated in Table S2.
observed at 531.96 eV, 532.02 eV and 531.80 eV for SNVO-2, SNVO-3 In previous reports, the NH4V4O10 nanobelts as a high capacity and
and SNVO-4 electrode respectively (Fig. 2f–h). Another small peak was useful cathode material have been documented for both sodium and
also seen at around 530.6 eV is due to the V–O–V bond formation lithium-ion battery [12,26]. But this cathode material provides a low
(Fig. 2e–h). It was also interesting to notice that with increasing the nominal potential of 1.95 V, which corresponds to 1.68 V full-cell
chemical sodiation process, the higher binding energy (oxygen bind voltage. Not only that, this cathode material is a sodium-free elec­
with Vþ4) O1s peak intensity (area under the curve) is rising signifi­ trode. These two combined effects hinder this particular cathode from
cantly compared with lower binding energy O1s peak (oxygen bind with the commercial applications, although it has a very high energy density
higher state Vþ5) [32,33]. Therefore, XPS analysis of both oxygen and ca. 655 W h kg 1 [26]. Herein, we addressed these issues and conse­
vanadium oxidation states again confirmed the sodium insertion by the quently enhanced the nominal potential as well as incorporated
chemical process into ammonium vanadium oxide matrix, and the so­ sodium-ion into this cathode materials by a simple combined chemical
dium intercalation is also increased with sodiation concentration up to a sodiation process.

4
A. Sarkar and S. Mitra Journal of Power Sources 452 (2020) 227832

Among all chemically sodiated ammonium vanadium oxide, SNVO-3 more favorable at higher reduction potential, and deintercalation pro­
showed the best electrochemical performances (details electrochemical cess fascinated at lower oxidation potential which helps to enhance the
performances of all SNVO-n were shown in supporting information). overall nominal potential of the cathode material.
Here, we have selected SNVO-3 composition as cathode materials for The charge-discharge curve of different SNVO cathode materials at
further experiments. The cyclic voltammetry (CV) of the SNVO-3 cath­ the current rate of 100 mA g 1 over 1.4 V–4.2 V potential window was
ode materials against Naþ/Na was performed in a half-cell configuration displayed in Fig. S6. From this Fig. S6a, it was noticed that only SNVO-1
at a scanning rate of 0.1 mV s 1 over the potential window 1.4 V–4.3 V delivered poor cycling stability. Fig. S6 showed an initial discharge ca­
vs. Na/Naþ (Fig. 3a). The CV curve exhibited multiple oxidation- pacity of ca. 190 mAh g 1, 225 mAh g 1, 200 mAh g 1, 125 mAh g 1, 95
reduction peaks of vanadium as usual. The oxidation and reduction mAh g 1 and 70 mAh g 1 for SNVO-1, SNVO-2, SNVO-3, SNVO-4,
peaks at 3.65 V and 3.52 V corresponding to V5þ/V4þ redox couple, SNVO-5, and SNVO-6 respectively. It was also noticed that after sodia­
whereas at the lower potential oxidation and reduction peaks at 2.04 V tion, the nominal voltage of the cathode is increased to 2.43 V from 2.1
and 1.78 V represent to V4þ/V3þ redox couple. Some minor reduction V. Among all SNVO-3 exhibited best electrochemical performances,
peaks were also noticed at 2.3 V, 2.5 V, 2.75 V, and 3.25 V whereas, the showing an initial discharge capacity of 200 mAh g 1 with stable cycling
corresponding oxidation peaks appeared at the potential of 2.55 V, 2.7 performance at 100 mA g 1 current density. So, we choose SNVO-3 as
V, 2.9 V, and 3.4 V, respectively. These minor peaks appeared due to the cathode materials for further study.
step by step phase transformation process during sodium insertion/ Fig. 3b and c shows the cycling performance of SNVO-3 cathode
deinsertion of/from the host materials, as shown in our previous report against Naþ/Na at 100 mA g 1 current density over 1.4 V–4.0 V po­
[25,26]. The active elements have mixed phase of V5þ/V4þ (also tential window. SNVO-3 sample exhibited an initial discharge capacity
confirmed by XPS), and in the first cycle, during the reduction process, of 170 mAh g 1 with 93% capacity retention after 100 cycles with a high
the only V4þ phase was transfer to V3þ phase by inserting 3–4 Naþ average Coulombic efficiency of 99.8%. It showed a high rate electro­
stepwise at 2.04 V. During the oxidation process (V4þ to V5þ), phase chemical performances and displayed an initial discharge capacity of
formation was happened stepwise due to 0.5, 1, 3 and 4 Naþ insertions ~250 mAh g 1, 200 mAh g 1, 170 mAh g 1, 133 mAh g 1 and 112 mAh
in the electrode. At 3.53 V all V5þ phases were converted to V4þ phase by g 1 at the current rate of 50 mA g 1, 100 mA g 1, 200 mA g 1, 500 mA
reduction, and the process continued. Here, no irreversible peak at 3.91 g 1 and 1000 mA g 1 current rate with ~84%, 96%, 98%, 96% and
V was observed for SNVO-3 electrode which means these sodiated ma­ 100% capacity retention after 50 cycles respectively (Fig. S7). It was
terials become more stable at higher potential (for better comparison CV interesting to notice that the nominal potential of the cell was enhanced
of NH4V4O10 was shown in Fig. S5). As we inserted more and more to 2.43 V. We further increased the electrochemical performances by
sodium-ion (up to 5 sodium-ion insertions) into NVO matrix, it becomes zirconium doping on vanadium side of SNVO-3
more stable, and ammonium-ion did not come out from the structure (as (Zr0.03Na0.78NH4V3.88O9.88) samples without incorporating any extra
it is bounded with O atoms by week van der Waals attraction) at higher process. Details characterization of zirconium-doped sodiated ammo­
potential (3.91 V) during the charging process. It was also noticed that nium vanadium oxide (Zr-SNVO-3) was shown in the Supporting In­
the intensity of reduction peak at 3.52 V (V5þ → V4þ) was increased formation, Figs. S8 and 9). Fig. 3d reveled the charge-discharge cycling
whereas peak intensity at 1.78 V (V5þ → V4þ) was reduced. However, for performance of Zr-SNVO-3 at 200 mA g 1 current density over 1.5–3.9 V
oxidation case, exact opposite phenomena were noticed; lower oxida­ potential window. This electrode was capable of delivering a high
tion peak (V3þ → V4þ) intensity increases and while higher oxidation discharge capacity of 275 mAh g 1 in its initial cycle and stable to 255
peak (V4þ → V5þ) intensity reduces. So, the intercalation process was mAh g 1 after 2nd cycle. The electrode sustained its 95% initial capacity

Fig. 3. Electrochemical performances of SNVO-3 and Zr-doped SNVO-3 electrode at 20 � 2 � C; (a) cyclic voltammetry (CV) of SNVO-3 at a scan rate of 0.1 mV s 1
over 1.4–4.3 V potential window, (b) cycling performance of SNVO-3 at 200 mA g 1 current density over 1.4–4.0 V potential window up to100 cycles, (c) charge-
discharge profile of SNVO-3, (d) cycling performance of Zr-SNVO-3 at 200 mA g 1 current rate with Coulombic efficiency over 1.5–3.9 V potential window, (e) 2nd
cycle charge-discharge profile of Zr-SNVO-3 at different current rate and (f) electrochemical performance comparison of different ammonium vanadium oxides; NVO,
Zr-NVO, SNVO-3 and Zr-SNVO-3 at 100 mA g 1 current density.

5
A. Sarkar and S. Mitra Journal of Power Sources 452 (2020) 227832

after 100 cycles (Fig. 3d), and 88% capacity after 200 cycles (Fig. S10a) in Table S4 and Table S5).
with an average 97.98% Coulombic efficiency. The 2nd cycle charge- Encouraged by the above outstanding electrochemical performance
discharge profile at the different current rates was shown in Fig. 3e of zirconium-doped sodiated NH4V4O10 cathode, we constructed a
for all the samples as a comparison. The Zr-doped SNVO-3 electrode sodium-ion full-cell based on this cathode material and zirconium-doped
exhibited a high discharge capacity of 275 mAh g 1, 255 mAh g 1, 182 hydrogenated sodium titanium oxide (Zr–HNa2Ti3O7) anode material
mAh g 1, and 145 mAh g 1 at a current density of 100 mA g 1, 200 mA (half-cell electrochemical performance of anode as shown in Fig. S11)
g 1, 500 mA g 1, and 1000 mA g 1 respectively, (Fig. 3e and Fig. S10b). [35]. In the full-cell, loading of an active cathode to anode mass ratio
The nominal potential of the cathode was further increased to 2.80 V was about 1:1. Fig. 4 shows the electrochemical performances of the
from 2.43 V by the chemical sodiation and doping processes. It was sodium-ion full-cell constructed with the above-mentioned cathode and
interesting to notice that a prolong discharge plateau at 3.52 V was anode materials without further modification. The cyclic voltammetry
observed in Zr-SNVO-3 electrode (similar behavior also found in the CV, (CV) of the full-cell was carried out at a scan rate of 0.05 mV s 1 over
Fig. 3a), which help to enhance the nominal potential of the cathode 1.4–4.0 V to identify the intercalation-deintercalation potential range
materials. However, other electrode did not show the prominent (Fig. 4a).
discharge plateau at 3.52 V. The chemical sodiation process opens some The CV exhibited one prominent reduction peak ca. 2.80 V and three
active sites (like (010), (110) and (111) planes become more active in oxidation peaks at 3.10, 3.23, and 3.38 V. Some minor reduction (at 1.78
XRD pattern, Fig. S8) for sodium-ion insertion at higher discharge state and 2.05 V) and oxidation peaks (at 1.86 and 2.13 V) were also
(reduction of V5þ to V4þ) and with zirconium doping the sodium observed. Fig. 4b displayed the charge-discharge profile of the full-cell
insertion channels becomes wider, helping to enhance the nominal po­ at a current density of 200 mA g 1 (cathode base) over 1.2–3.7 V po­
tential. The comparison of the electrochemical performance of tential window. At this current rate, the full-cell provided a specific
NH4V4O10 (NVO), Zr– NH4V4O10 (Zr-NVO), SNVO-3 and Zr-SNVO-3 discharge capacity of 176 mAh g 1 (cathode base) with a high nominal
(Zr0.03Na0.78NH4V3.88O9.88) were shown in Fig. 3f and Table S3 at voltage of 2.5 V. This full-cell is capable of delivering a high energy
100 mA g 1 current density vs. Naþ/Na. The NVO electrode delivered a density of 440 W h kg 1 (cathode base). However, if we combine both
specific discharge capacity of 210 mAh g 1 with the nominal potential of active mass of anode and cathode materials, it still provided a high
2.1 V, corresponds to a high energy density of 441 W h kg 1. However, energy density of 220 W h kg 1, which is comparable with the existing
with zirconium doping, the specific capacity was enhanced to 367 mAh commercial lithium-ion battery technology. The cycling performance of
g 1, and the overall energy density increased to 702 W h kg 1 (nominal the full-cell was shown in Fig. 4c. The full-cell showed initial discharge
potential 1.91 V). The discharge capacity is decreased to 190 from 205 capacity ca. 176 mAh g 1, and after 90 cycles and sustained its 83%
mAh g 1 after chemical sodiation of NVO materials, but the nominal initial capacity with an average Coulombic efficiency of 99.58% with
voltage is increased to 2.37 V from 2.1 V. Therefore, we obtained almost excellent rate capability (Fig. 4d). The comparison study of the essential
similar energy density (450 W h kg 1) from the current studied material. electrochemical parameters such as specific capacity, voltage, energy
However, after zirconium doping on SNVO-3, the specific capacity of the density and power density of recent sodium-ion full-cell prototype was
materials further improved to 275 mAh g 1 with an average nominal exhibited in Fig. 4e and f, and Table S6. Fig. 4e and f attributed that our
potential of 2.80 V. Therefore, we achieved an overall energy density of full-cell prototype delivered higher energy density and superior power
770 W h kg 1 from the current cell in coin cell configuration. This is one density compare with any other sodium-ion full-cell exit in the litera­
of the highest energy density obtained in any kind of cathode materials ture. We also demonstrate a practical application of this sodium-ion full-
used in lithium or sodium-ion battery (details comparison study shown cell in coin cell configuration (Fig. 4h). A single coin-cell constructed

Fig. 4. Electrochemical performance of the Zr-doped sodiated NH4V4O10 (Zr–SHNVO-3)//H–Na2Ti3O7 full-cell prototype with an active mass loading of cathode to
anode is 1:1 at 20 � 2 � C, (a) cyclic voltammetry at scan rate 0.05 mV s 1 over potential window 1.4–4.0 V, (b) charge-discharge at 200 mA g 1 current rate (with
respect to cathode) (c) cycling performance over 100 cycles at 200 mA g 1 current density, (d) Power cycling performance up to 200 cycles, (e–f) comparison of all
recent work on sodium-ion full-cell with our Zr-SNVO-3//H–Na2Ti3O7 full-cell, (e) Specific capacity vs. full-cell voltage profile of different of sodium-ion full-cell
developed in recent time (f) Ragone plot of different sodium-ion full-cell published in recent years with our prototype, and (g) a demonstration of prototype using a
light-emitting diode (LED) study lamp powered by a single coin-cell using our prototype full-cell.

6
A. Sarkar and S. Mitra Journal of Power Sources 452 (2020) 227832

with this cathode and anode material, lighten a study table lamp made References
with 36 LED bulbs in series.
[1] B.L. Ellis, W.R.M. Makahnouk, Y. Makimura, K. Toghill, L.F. Nazar,
A multifunctional 3.5 V iron-based phosphate cathode for rechargeable batteries,
3. Conclusions Nat. Mater. 6 (2007) 749–753, https://doi.org/10.1038/nmat2007.
[2] M.H. Han, E. Gonzalo, G. Singha, T. Rojo, A comprehensive review of sodium
In brief, ammonium vanadium oxide was chemically sodiated by a layered oxides: powerful cathodes for Na-ion batteries, Energy Environ. Sci. 8
(2015) 81–102, https://doi.org/10.1039/c4ee03192j.
single step hydrothermal technique without adding any extra process. In [3] C. Fang, Y. Huang, W. Zhang, J. Han, Z. Deng, Y. Cao, H. Yang, Routes to high
SNVO materials, the ammonium-ions are bounded with oxygen by weak energy cathodes of sodium-ion batteries, Adv. Energy Mater. (2016) 61501727,
van der Waals’s forces, and as a result, it showed better thermal and https://doi.org/10.1002/aenm.201501727.
[4] P.F. Wang, Y.J. Guo, H. Duan, T.T. Zuo, E. Hu, K. Attenkofer, H. Li, X.S. Zhao, Y.
electrochemical stability than NVO materials. This 1-D growth SNVO X. Yin, X. Yu, Y.G. Guo, Honeycomb-ordered Na3Ni1.5M0.5BiO6 (M ¼ Ni, Cu, Mg,
nanobelts cathode materials exhibited better electrochemical perfor­ Zn) as high-voltage layered cathodes for sodium-ion batteries, ACS Energy Lett. 2
mance for SIB with the highest discharge capacity of ~200 mAh g 1 (at (2017) 2715–2722, https://doi.org/10.1021/acsenergylett.7b00930.
[5] A. Ponrouch, E. Marchante, M. Courty, J. Tarascon, M.R. Palac, In search of an
the current rate of 100 mA g 1) and high stability with excellent rate
optimized electrolyte for Na-ion batteries, Energy Environ. Sci. 5 (2012)
capability. Further, the specific capacity was enhanced by zirconium 8572–8583, https://doi.org/10.1039/c2ee22258b.
doping into SNVO. Here, we achieved a high discharge capacity of ~225 [6] J. Liu, K. Tang, K. Song, P.A. van Aken, Y. Yu, J. Maier, Electrospun Na3V2(PO4)3/C
mAh g 1 at 200 mA g 1 current density and the nominal potential nanofibers as stable cathode materials for sodium-ion batteries, Nanoscale 6 (2014)
5081, https://doi.org/10.1039/c3nr05329f.
further improved to 2.80 V from 2.1 V. Hence, the overall energy density [7] Z. Jian, L. Zhao, H. Pan, Y. Hu, H. Li, W. Chen, L. Chen, Carbon coated Na3V2(PO4)3
of the cathode material was achieved to 770 W h kg 1 which is one the as novel electrode material for sodium ion batteries, Electrochem. Commun. 14
best among all other cathode materials used in SIB or LIB system. Here, a (2012) 86–89, https://doi.org/10.1016/j.elecom.2011.11.009.
[8] D. Su, G. Wang, Single-crystalline bilayered V2O5 nanobelts for high-capacity
new class of sodium-ion full-cell was developed by constructing a cell sodium-ion batteries, ACS Nano 7 (2013) 11218–11226, https://doi.org/10.1021/
consists of Zr-SNVO-3//Zr–HNa2Ti3O7, which showed an excellent nn405014d.
reversible capacity of 176 mAh g 1 at 200 mA g 1 current rate with [9] Y. Fang, L. Xiao, X. Ai, Y. Cao, H. Yang, Hierarchical carbon framework wrapped
Na3V2(PO4)3 as a superior high-rate and extended lifespan cathode for sodium-ion
stable cyclic performance up to 200 cycles with an average 99.58% batteries, Adv. Mater. 27 (2015) 5895–5900, https://doi.org/10.1002/
Coulombic efficiency. This high-performance sodium-ion full-cell adma.201502018.
exhibited an excellent and high energy density of 220 W h Kg 1 (w.r.t [10] Y. Huang, Y. Zheng, X. Li, F. Adams, W. Luo, Y. Huang, L. Hu, Electrode materials
of sodium-ion batteries toward practical application, ACS Energy Lett. 3 (2018)
cathode and anode mass) with a high nominal potential of 2.5 V even at 1604–1612, https://doi.org/10.1021/acsenergylett.8b00609.
a high rate of 200 mA g 1. The sodium-ion full-cell prototype showed [11] S. Sarkar, A. Bhowmik, D. Bharadwaj, S. Mitra, Phase transition , electrochemistry ,
one of the best performing and promising performance in the sodium- and structural studies of high rate LixV3O8 cathode with nanoplate morphology,
J. Electrochem. Soc. 161 (2014) A14–A22, https://doi.org/10.1149/2.006401jes.
ion battery literature. This technology could be one of the best alter­
[12] S. Sarkar, P.S. Veluri, S. Mitra, Morphology controlled synthesis of layered
nates to the existing lithium-ion battery technology for renewable en­ NH4V4O10 and the impact of binder on stable high rate electrochemical
ergy storage applications shortly. performance, Electrochim. Acta 132 (2014) 448–456, https://doi.org/10.1016/j.
electacta.2014.03.144.
[13] Y. Sun, B. Qu, S. Jiang, C. Wu, B. Pan, Y. Xie, Highly depressed temperature-
Author contributions induced metal-insulator transition in synthetic monodisperse 10-nm V2O3
pseudocubes enclosed by {012} facets, Nano Today 3 (2011) 2609–2614, https://
The manuscript was written through the contributions of all authors. doi.org/10.1039/c1nr10179j.
[14] A. Sarkar, A. Kumar, S. Mitra, Nanostructured vanadium tri-oxides , as a long life
All authors have approved the final version of the manuscript. and high performance anode for sodium-ion battery, Electrochim. Acta 299 (2019)
914–925, https://doi.org/10.1016/j.electacta.2019.01.076.
Funding sources [15] H. Fei, X. Liu, Y. Lin, M. Wei, Facile synthesis of ammonium vanadium oxide
nanorods for Na-ion battery cathodes, J. Colloid Interface Sci. 428 (2014) 73–77,
https://doi.org/10.1016/j.jcis.2014.04.029.
The authors are thankful for the financial support provided by MHRD [16] G. He, W.H. Kan, A. Manthiram, A 3.4 V layered VOPO4 cathode for Na-ion
(Ministry of Human Resource Development) and NCPRE (Grant No. 31/ batteries, Chem. Mater. (2016) 682–688, https://doi.org/10.1021/acs.
chemmater.5b04605.
09/2015-16/PVSE-R&D) funded by the Ministry of New Renewable [17] V. Raju, J. Rains, C. Gates, W. Luo, X. Wang, W.F. Stickle, G.D. Stucky, X. Ji,
Energy, Govt. of India. Superior cathode of sodium-ion batteries: orthorhombic V2O5 nanoparticles
generated in nanoporous carbon by ambient hydrolysis deposition, Nano Lett. 14
(2014) 4119–4124, https://doi.org/10.1021/nl501692p.
Decalartion of competing interest
[18] H. Yi, L. Lin, M. Ling, Z. Lv, R. Li, Q. Fu, H. Zhang, Q. Zheng, X. Li, Scalable and
economic synthesis of high-performance Na3V2(PO4)2F3 by a solvothermal–ball-
The authors declare that the research was conducted in the absence milling method, ACS Energy Lett. 2 (2019) 1565–1571, https://doi.org/10.1021/
of any commercial or financial relationships that could be construed as a acsenergylett.9b00748.
[19] C.V. Manohar, A.R. K, M. Kar, M. Forsyth, D.R. MacFarlane, S. Mitra, Stability
potential conflict of interest. enhancing ionic liquid hybrid electrolyte for NVP@C cathode based sodium
batteries, Sustain. Energy Fuels 2 (2018) 566–576, https://doi.org/10.1039/
Acknowledgments C7SE00537G.
[20] D. Chao, C.M. Lai, P. Liang, Q. Wei, Y. Wang, C.R. Zhu, G. Deng, V.V.T. Doan-
nguyen, J. Lin, L. Mai, H.J. Fan, B. Dunn, Z.X. Shen, Sodium vanadium
The authors are indebted to SAIF, IIT-B, for their assistance in FEG- fluorophosphates (NVOPF) array cathode designed for high-rate full sodium ion
TEM, NCPRE for FEG-SEM facility, and DESE for XRD analysis and storage device, Adv. Energy Mater. 8 (2018) 1800058, https://doi.org/10.1002/
aenm.201800058.
IRCC, IITB for their assistance in XPS. We are very thankful to Arnab [21] C. Deng, S. Zhang, Z. Dong, Y. Shang, 1D nanostructured sodium vanadium oxide
Ghosh for helping us to propose the reaction mechanism. We are also as a novel anode material for aqueous sodium ion batteries, Nano Energy 4 (2014)
grateful to Dr. Sudeep Sarkar and all the ECEL members for their 49–55, https://doi.org/10.1016/j.nanoen.2013.12.014.
[22] S. Batteries, S. Tepavcevic, H. Xiong, V.R. Stamenkovic, X. Zuo,
continuous support and encouragement. M. Balasubramanian, Nanostructured bilayered vanadium oxide electrodes for
rechargeable, ACS Nano 6 (2012) 530–538, https://doi.org/10.1021/nn203869a.
Appendix A. Supplementary data [23] D. Wang, X. Bie, Q. Fu, D. Dixon, N. Bramnik, Y. Hu, F. Fauth, Y. Wei,
H. Ehrenberg, G. Chen, F. Du, Sodium vanadium titanium phosphate electrode for
symmetric sodium-ion batteries with high power and long lifespan, Nat. Commun.
Supplementary data to this article can be found online at https://doi. 8 (2017) 15888, https://doi.org/10.1038/ncomms15888.
org/10.1016/j.jpowsour.2020.227832. [24] C. Lee, A.C. Marschilok, A. Subramanian, K.J. Takeuchi, E.S. Takeuchi, Synthesis
and characterization of sodium vanadium oxide gels : the effects of water (n) and
sodium (x) content on the electrochemistry of NaxV2O5.nH2O, Phys. Chem. Chem.
Phys. 13 (2011) 18047–18054, https://doi.org/10.1039/c1cp21658a.

7
A. Sarkar and S. Mitra Journal of Power Sources 452 (2020) 227832

[25] A. Sarkar, S. Sarkar, T. Sarkar, P. Kumar, M.D. Bharadwaj, S. Mitra, Rechargeable [31] K. Range, R. Zintl, A.M. Heyns, The thermal decomposition of ammonium
sodium-ion battery: high-capacity ammonium vanadate cathode with enhanced metavanadate (V) in open and closed systems, Verlag Z. Naturforsch. B Chem. Sci.
stability at high rate, ACS Appl. Mater. Interfaces 7 (2015) 17044–17053, https:// 43b (1988) 309–317. Coop. with Max Planck Soc.
doi.org/10.1021/acsami.5b03210. [32] S. Kundu, B. Satpati, M. Mukherjee, T. Kar, S.K. Pradhan, Hydrothermal synthesis
[26] A. Sarkar, S. Sarkar, S. Mitra, Exceptionally high sodium-ion battery cathode of polyaniline intercalated vanadium oxide xerogel hybrid nanocomposites:
capacity based on doped ammonium vanadium oxide and a full cell SIB prototype effective control of morphology and structural characterization, New J. Chem. 41
study, J. Mater. Chem. A. 5 (2017) 24929–24941, https://doi.org/10.1039/ (2017) 3634–3645, https://doi.org/10.1039/c7nj00372b.
c7ta08104a. [33] Q.H. Lu, R. Huang, L.S. Wang, Z.G. Wu, C. Li, Q. Luo, S.Y. Zuo, J. Li, D.L. Peng, G.
[27] H. Si, L. Seidl, E. Miao, L. Chu, S. Martens, J. Ma, X. Qiu, U. Stimming, L. Han, P.X. Yan, Thermal annealing and magnetic anisotropy of NiFe thin films on
O. Schneider, Impact of the morphology of V2O5 electrodes on the electrochemical n þ -Si for spintronic device applications, J. Magn. Magn Mater. 394 (2015)
Na þ -ion intercalation, J. Electrochem. Soc. 165 (2018) 2709–2717, https://doi. 253–259, https://doi.org/10.1016/j.jmmm.2015.06.066.
org/10.1149/2.0621811jes. [34] R. Precht, S. Stolz, E. Mankel, T. Mayer, W. Jaegermann, R. Hausbrand,
[28] J. Rodríguez-Carvajal, Recent advances in magnetic structure determination by Investigation of sodium insertion into tetracyanoquinodimethane (TCNQ): results
neutron powder diffraction, Phys. B Condens. Matter 192 (1993) 55–69, https:// for a TCNQ thin film obtained by a surface science approach, Phys. Chem. Chem.
doi.org/10.1016/0921-4526(93)90108-I. Phys. 18 (2016) 3056–3064, https://doi.org/10.1039/c5cp06659j.
[29] M. Taniguchi, T.R. Ingraham, Mechanism of thermal decomposition of ammonium [35] A. Sarkar, C.V. Manohar, S. Mitra, A simple approach to minimize the first cycle
metavanadate, Can. J. Chem. 42 (1964) 2467–2473. irreversible loss of sodium titanate anode towards the development of sodium-ion
[30] A.K. Sarkar, S. Saha, S. Ganguly, D. Banerjee, K. Kargupta, Hydrogen storage on battery, Nano Energy 70 (2020) 104520, https://doi.org/10.1016/j.
graphene using Benkeser reaction, Int. J. Energy Res. 38 (2014) 1889–1895, nanoen.2020.104520.
https://doi.org/10.1002/er.

You might also like