Subbukalai
Vijayakumar
,
Seong-Hun
Lee
and
Kwang-Sun
Ryu
*
Department of Chemistry and Energy Harvest Storage Research Center (EHSRC), University of Ulsan, Muger-dong, Nam-gu, Ulsan 680-749, Republic of Korea. E-mail: ryuks@ulsan.ac.kr; Fax: +82-52-259-2348; Tel: +82-52-259-2763
First published on 30th September 2015
Zn3V2O8 nanoplatelets were successfully synthesized using a hydrothermal method. The formation of the Zn3V2O8 nanoplatelets was explained via splitting, exfoliation and self-aggregation mechanisms. FESEM revealed the nanoplatelet morphology with a thickness of 27.9 nm. HRTEM imaging confirmed the crystalline nature of the Zn3V2O8 nanoplatelets, and the SAED pattern clearly indicated that the prepared sample was Zn3V2O8. The prepared Zn3V2O8 nanoplatelets were further studied for their potential application in Li-ion batteries and supercapacitors. The discharge capacity in the second cycle was 558 mA h g−1 at 100 mA g−1. The Zn3V2O8 nanoplatelets exhibited a maximum specific capacitance of 302 F g−1 at a scan rate of 5 mV s−1. Furthermore, a Zn3V2O8 electrode retained about 98% of its initial specific capacitance after 2000 cycles. The described Zn3V2O8 nanoplatelets were found to be a highly suitable electrode material for energy storage applications.
Transition metal oxides are an important class of functional materials, which have been widely used as electrode materials within energy storage devices, especially in supercapacitors and lithium-ion batteries.4 Recently, it has been shown that a number of mixed transition metal oxides are promising candidates for energy storage devices such as lithium-ion batteries and electrochemical supercapacitors. Mixed transition metal oxides can deliver diverse redox reactions in electrochemical reactions due to the coexistence of two different metal species within a single crystal structure.5 Recently, metal vanadates have been applied as electrode materials within supercapacitors and lithium-ion batteries.6–12 Nickel vanadate on a nickel foam was studied as an electrode material for supercapacitors by Zhang et al.6 They reported the retention of 66.7% capacitance after 5000 cycles. Zhang et al.9 reported the first and second discharge capacities of 875.8 and 472.4 mA h g−1, respectively, at a current density of 100 mA g−1 for MnV2O6 nanobelt anodes. Similarly, Xiao et al.10 reported an initial reversible capacity of 548 mA h g−1 for ZnV2O4 hollow microspheres.
Zinc vanadates (Zn3V2O8 nanomaterials) have been extensively studied for their energy storage and photocatalytic applications. However, there has been far less interest in their optical, photocatalytic and energy storage applications. Very few attempts have been made to investigate the photonic,13 photocatalytic14 or lithium-ion battery applications of Zn3V2O8 nanomaterials.15,16 To the best of our knowledge, there have not been any previously published studies of the supercapacitor performance of Zn3V2O8 nanomaterials. Therefore, we made an attempt to synthesize Zn3V2O8 nanomaterials and study their electrode performance to determine their suitability for applications within supercapacitors.
In this study, we report the synthesis of Zn3V2O8 nanoplatelets by a hydrothermal method using sodium dodecyl sulfate as a surfactant. The structural and morphological features of the prepared Zn3V2O8 nanoplatelets were characterized using FESEM and HRTEM. A possible mechanism of the formation of the platelet structure, as well as the application of Zn3V2O8 nanoplatelets within lithium-ion batteries and supercapacitors, is reported and discussed in detail.
The purity and oxidation state of Zn3V2O8 nanoplatelets were characterized using X-ray photoelectron spectroscopy (XPS). Fig. 3a shows the complete survey spectrum of Zn3V2O8 nanoplatelets. It shows the presence of V 2p, O 1s, C 1s, and Zn 2p states. Fig. 3b shows the Zn 2p spectrum. The two peaks in the Zn 2p spectrum at 1021.58 and 1044.68 eV correspond to Zn 2p3/2 and Zn 2p1/2, respectively.18 The V 2p spectrum is shown in Fig. 3c. The peak positions at 516.78 (V 2p3/2) and 524.38 eV (V 2p1/2) were attributed to the V5+ state.19Fig. 3d shows the XPS spectrum of O 1s. The O 1s spectrum was deconvoluted into three components (O1, O2, and O3), with binding energies of 529.58, 531.0 and 532.4 eV, respectively. The O1 peak at 529.58 eV was attributed to a metal–oxygen bond, i.e. V–O binding in Zn3V2O8.20 The O2 peak at 531.0 eV was attributed to a large number of defect sites with minimum oxygen coordination and small particle size.21 The O3 component at the higher binding energy of 532.4 eV corresponds to chemically and physically bonded water within the surface.21
Fig. 4a–c shows FESEM images of Zn3V2O8 nanomaterials at different magnifications. The FESEM images of the prepared Zn3V2O8 nanomaterials show nanoplatelet morphology. Fig. 4a and b clearly show the even distribution of nanoplatelets with a flake-like arrangement. Fig. 4c depicts the nanoplatelet morphology and dimensions. The platelets are arranged with spacing between neighboring nanoplatelets, which offers effective electron and ion transport and enables the improvement of electrochemical performance in batteries and supercapacitors. The length of the nanoplatelets is about 1 μm and their breadth is between 500 nm and 1 μm.
Fig. 4d shows an AFM image of zinc vanadate nanoplatelets. The thickness of the nanoplatelets measured by AFM is 27.9 nm. The structure of the Zn3V2O8 nanoplatelets was further characterized using HRTEM. Fig. 4e and f shows HRTEM images of Zn3V2O8 nanoplatelets. The TEM images clearly reveal the nanoplatelet arrangement and dimensions. Fig. 4g shows a higher-magnification HRTEM image. The image clearly shows the crystalline structure of the sample. The d-spacing corresponds to the (271) plane of Zn3V2O8. Fig. 4h shows the selected-area electron diffraction (SAED) pattern of Zn3V2O8 nanoplatelets. The SAED pattern can be indexed to Zn3V2O8.
A time-dependent experiment was carried out to explain the formation of nanoplatelets. Fig. 5 shows FESEM images of the precursor sample prepared at different durations of hydrothermal treatment. A mechanism of the formation of the Zn3V2O8 nanoplatelets based on the results of the time-dependent experiment is shown in Fig. 6. The formation of Zn3V2O8 nanoplatelets includes nucleation, growth, splitting and exfoliation processes, as well as self-aggregation. Initially, zinc and vanadium ions reacted to form zinc vanadate nuclei. After nucleation, a growth process started to form a zinc vanadate nanostructure. Herein, the role of SDS was important.
![]() | ||
Fig. 5 FESEM images of Zn3V2O8 precursor sample at different durations of hydrothermal treatment: (a) 6 h, (b) 12 h, (c) 18 h, and (d) 24 h. |
SDS acted as a structure-directing agent that controls the surface free energy.22 Therefore, it promoted the growth process. The structure of the final product was dependent upon the hydrothermal temperature, duration of treatment, concentration of the surfactant and pH of the solution. Initially, thick zinc vanadate platelets were formed. As the hydrothermal reaction continues, the platelets split into thin sheets.19
Fig. 5a shows (after 6 h of hydrothermal reaction) thick nanoplates with sheets. These sheets may have formed due to a splitting process. The splitting process continued for up to 12 hours of hydrothermal treatment (Fig. 5b). The sheets were then exfoliated to form thin platelets (Fig. 5c).19 After exfoliation, the sheets were attached together by self-aggregation, forming zinc vanadate nanoplatelets (Fig. 5d).
The Zn3V2O8 nanoplatelets were subjected to BET measurement to determine the details of specific surface area and pore size distribution. Fig. S2† shows the BET isotherm and corresponding pore size distribution curve of Zn3V2O8 nanoplatelets. The BET surface area of the Zn3V2O8 nanoplatelets was 11.53 m2 g−1. The pore size distribution curve for adsorption by the BJH method clearly indicates that most of the pores were distributed in the region of 2–10 nm. The mean pore diameter of the Zn3V2O8 nanoplatelets was 5.9 nm. The large surface area with a porous nature helps to enhance the diffusion of ions, which improves the electrochemical performance in both Li-ion batteries and supercapacitors.23
The electrochemical performance of Zn3V2O8 nanoplatelets was studied using cyclic voltammetry and charge–discharge analysis and their potential as the active material for lithium-ion batteries was determined. Fig. 7a shows CV curves of Zn3V2O8 nanoplatelets in the potential range of 0.01 to 3.0 V at a scan rate of 0.1 mV s−1. The cathodic scans of the CV curves show reduction peaks at 0.59 V, which are due to the reduction of zinc (Zn2+ to Zn0) and reduction of vanadium (V5+ to V4+ and further to V3+).16,24 The oxidation peaks at 1.23 V are due to the oxidation of vanadium (V3+ to V5+) and zinc (Zn0 to Zn2+). The charge–discharge measurement of Zn3V2O8 nanoplatelets was carried out using a potential range of 0.01 to 3.0 V at a current density of 100 mA g−1. Fig. 7b shows the charge–discharge profile of Zn3V2O8 nanoplatelets at cycles 1, 2, 3, 10, 20, 30 and 40. The discharge curve of the first cycle is different from those of the subsequent cycles. It shows a plateau at about 1.1 to 0.19 V. The capacity of the first-cycle plateau at less than 0.19 V is due to a slow process of lithium storage into orthorhombic-type Zn3V2O8, and the plateau at higher voltage can be attributed to the faradic nature of the material.16
The initial discharge and charge capacities were 752 and 488 mA h g−1, respectively. The irreversible capacity loss in the first cycle was 35.1%. After the first cycle, the electrochemical process was highly reversible. The discharge and charge capacities in the second cycle decreased to 558 and 474 mA h g−1, respectively, and the capacity loss in the second cycle was only 15%. The irreversible capacity loss in the first cycle may be attributed to the loss of lithium due to the formation of a solid electrolyte interphase (SEI) layer.24Fig. 7c and d show the cycling stability curve and coulombic efficiency of Zn3V2O8 nanoplatelets. The specific capacity decreased from the first cycle to the second cycle and further large decreases in specific capacity were observed up to the 10th cycle. Thereafter, the specific capacity decreased linearly. The coulombic efficiency of Zn3V2O8 nanoplatelets increased from the first cycle (64.9%) to the second cycle (85.1%), rising to greater than 95% after the 10th cycle and greater than 99% at the 40th cycle. Zhang et al.15 synthesized Zn3(VO4)2 microspheres as an anode and reported first and second discharge capacities of 1180 and 676 mA h g−1, respectively, at a current density of 20 mA g−1. Although the discharge capacities were high, the current density was very low. Gan et al.16 reported a first discharge capacity of 1522 mA h g−1 and excellent cycling stability for Zn3V2O8 material. However, they used 30% conductive material, which likely resulted in a low specific volumetric capacity.
The electrochemical supercapacitor performance of the Zn3V2O8 nanoplatelets was studied using cyclic voltammetry and galvanostatic charge–discharge studies. The CV study was performed within the potential window of 0–0.5 V at different scan rates. Fig. 8a shows the CV curves of the Zn3V2O8 nanoplatelet electrode at scan rates of 5, 10, 25, and 50 mV s−1. The shapes of the CV curves do not reveal electric double-layer capacitance, which indicates that the capacitance was mainly due to pseudocapacitive behavior.25 The CV curves show a pair of redox peaks at about 0.37 and 0.25 V. These redox peaks may originate from the intercalation and de-intercalation of K+ from the electrolyte into zinc oxide.26 Fig. S3† shows the curve of scan rate versus peak current. It is interesting to note that the variation of peak current with scan rate is linear. This linearity of peak current with scan rate confirms that the electrochemical reaction was a surface redox reaction that led to pseudocapacitive behavior of the electrode.27 The specific capacitance Cc (F g−1) was calculated using the following formula:28
![]() | (1) |
The supercapacitor performance was further evaluated for its specific capacitance and cycling stability using charge–discharge analysis. Fig. 8c shows the discharge curves of the Zn3V2O8 electrode at different specific currents: 2, 4, 6, and 8 A g−1. The discharge profiles reveal nonlinear rather than linear behavior of the electric double-layer capacitance, which suggests that the capacitance was mainly due to pseudocapacitance.30 The specific capacitance Cp (F g−1) was calculated using the following formula:31
![]() | (2) |
Long-term cycling stability is important for practical applications. The long-term cycling stability of the Zn3V2O8 electrode was tested over 2000 continuous charge–discharge cycles at a specific current of 8 A g−1. Fig. 9 shows the cycling stability curve for 2000 cycles. The inset of Fig. 9 shows the charge–discharge curve of the first and last five cycles (cycles 1–5 and 1996–2000). The specific capacitance increased from 100% to 120% after 15 cycles, decreasing slowly thereafter. After 2000 cycles, about 98% of the initial specific capacitance was retained. The Zn3V2O8 nanoplatelet electrode exhibited good specific capacitance with high cycling stability. This appreciable performance is mainly due to the crystal structure and morphology of the Zn3V2O8 nanoplatelets. The crystal structure of Zn3V2O8 is similar to that of Ni3V2O8. Three types of O atom constitute the tetrahedron and octahedron. V is located in the center of the tetrahedron and Zn is located in the center of the octahedron.16 The gap between the octahedron and the tetrahedron, as well as channels of the supercell, allows more effective electron and ion transport, which leads to an improved electrochemical performance.
![]() | ||
Fig. 9 Curve of cycling stability against number of cycles. (Inset shows the initial and final five cycles of charge–discharge.) |
Footnote |
† Electronic supplementary information (ESI) available: XRD pattern of Zn3V2O8 precursor sample, BET isotherm of Zn3V2O8 and scan rate versus peak current curve. See DOI: 10.1039/c5ra13904j |
This journal is © The Royal Society of Chemistry 2015 |