Ediga Umeshbabua,
G. Rajeshkhannaa,
Ponniah Justinb and
G. Ranga Rao*a
aDepartment of Chemistry, Indian Institute of Technology Madras, Chennai – 600036, India. E-mail: grrao@iitm.ac.in; Fax: +91 44 2257 4202; Tel: +91 44 2257 4226
bDepartment of Chemistry, Rajiv Gandhi University of Knowledge Technologies, RK Valley, Kadapa-516330, Andhra Pradesh, India
First published on 29th July 2015
The spinel NiCo2O4 material has received considerable attention as an excellent supercapacitor material. In this study, we report a facile and cost-effective solvothermal method for the synthesis of mesoporous NiCo2O4 anchored on reduced graphene oxide (rGO). The electrochemical activity of the NiCo2O4–rGO and pristine NiCo2O4 materials were evaluated by cyclic voltammetry (CV), chronopotentiometry (CP) and electrochemical impedance spectroscopy (EIS). The NiCo2O4–rGO composite electrode shows a high specific capacitance value of 870 F g−1 at a current density of 2 A g−1 and it retains 600 F g−1 capacitance even at a high current density of 20 A g−1. Pristine NiCo2O4 shows a poor capacitance value of 315 F g−1 at 2 A g−1 and it retains only 191 F g−1 at 10 A g−1. Furthermore, the NiCo2O4–rGO nanocomposite shows an excellent cyclic performance with 90% capacitance retention even after 5000 charge–discharge cycles at a high current density of 4 A g−1, whereas a pristine NiCo2O4 electrode shows only 45% capacitance retention. The high specific capacitance, remarkable rate capability and excellent cycling performance offered by the NiCo2O4–rGO composite is attributed to the high surface area and high conductivity. In addition, rGO is believed to shorten the diffusion, migration paths for electrolyte ions and an easy access for electrolyte ions into redox centers.
Graphene is a 2D atomic sheet of sp2-hybridized carbon atoms arranged in a hexagonal lattice. The planar sp2-hybrid orbitals form σ-bonds in the honeycomb lattice and the remaining 2pz orbitals on carbon atoms are delocalized in π symmetry orientation. The delocalized 2pz electrons are largely responsible for high electronic conductivity of graphene. In addition to conductivity, graphene also has desirable characteristics such as high surface area (theoretical value of 2630 m2 g−1), good thermal/chemical stability and high structural flexibility.29–31 Due to these unique properties, graphene is considered as a good blending material to prepare graphene based metal oxide hybrid nanocomposites for energy storage and conversion applications.30–33 In electrochemical systems, graphene can be effective in improving specific capacitance, rate capability and cyclic permanence of the metal oxide–graphene hybrid nanocomposites by (i) controlling the nanocrystalline morphology of active materials, (ii) providing efficient electronic and ionic transport pathways, and (iii) improving the interfacial contact between active material and conductive graphene.
Recently, different NiCo2O4–graphene hybrid nanocomposites have been designed by different approaches and their performance tested for supercapacitors.26,34–39 For example, Wang et al. synthesised NiCo2O4–rGO composite using self-assembly method by exfoliating Ni–Co hydroxides and assembling with GO followed by heat treatment. These self-assembled 2D nanosheets of NiCo2O4–rGO composite exhibits higher specific capacitance of 835 F g−1 at 1 A g−1.35 Shen et al. fabricated rGO–NiCo2O4 nanocomposite by hydrothermal approach using poly(diallyldimethylammonium chloride) as an additive agent and obtained a specific capacitance of 671 F g−1 at 5 mV s−1.36 In another study, Wei et al. developed 3D mesoporous hybrid NiCo2O4@graphene nanoarchitectures by hydrothermal process using freeze dried graphene oxide and showed a maximum capacitance of 778 F g−1 at 1 A g−1.37 Wang et al. explored NiCo2O4 nanowires-loaded rGO composite synthesized by facile hydrothermal method using hexamethylenetetramine as a structure directing agent and exhibited a higher specific capacitance of 737 F g−1 at 1 A g−1.38 Microwave irradiation followed by mild heating has been adopted by Del Monte et al. to synthesize NiCo2O4–GO nanocomposite. This material contains NiCo2O4 nanoparticles embedded in the GO composite and showed higher specific capacitance value of 925 F g−1 at 1.5 A g−1.39 However, the synthesis methods discussed so far are complex and time-consuming. In addition, there is a possibility of restacking rGO or graphene sheets which can affect the electrochemical performance of the hybrid nanocomposites. There is scope to improve the morphology of NiCo2O4–rGO with uniform particle size and high surface area, while avoiding the aggregation of GO.
Here we report an efficient two step strategy to synthesize mesoporous NiCo2O4–rGO ultrathin nanosheets under solvothermal conditions in the presence of polymeric surfactant, polyvinylpyrrolidone (PVP). NiCo2O4–rGO ultrathin nanosheets are obtained by controlling the nucleation and growth of NiCo–glycolate nanosheets on the surface of rGO followed by thermal treatment. Reduced graphene oxide (rGO) acts as conductive support in this material and improves the electrochemical performance. The role of PVP is to control the size of NiCo2O4 particles and prevent the aggregation of GO.22,40,41 The obtained mesoporous NiCo2O4–rGO ultrathin nanosheets show high specific capacitance, remarkable rate capability and good cycling performance during charge–discharge process.
100 mg of as synthesized GO was dispersed in 100 mL of ethylene glycol (EG). After sonication for 15 min, 50 mg of PVP (weight-averaged molecular weight, Mw = 40000 g mol−1) was added and then stirred for 1 h. 1 g of Ni(CH3COOH)2·4H2O and 2 g of Co(CH3COOH)2·4H2O were added to the above solution and continued stirring for 3 h to get complete homogeneity. The homogeneous mixture was transferred to Teflon lined stainless steel autoclave and kept it in electrical oven at 180 °C for 12 h. The autoclave was then allowed to cool to room temperature, and resulting precipitate was separated by centrifugation, washed with excess ethanol and water, and then dried in oven at 60 °C for 12 h. The dried solid sample was calcined in air at 350 °C for 3 h. Pristine NiCo2O4 was prepared by the same method without adding GO. The NiCo–glycolate–rGO and NiCo–glycolate precursor samples after calcined were denoted as NiCo2O4–rGO and NiCo2O4 respectively.
The powder XRD pattern of uncalcined NiCo–glycolate–rGO precursor is shown in Fig. S1,† a strong diffraction peak at 10.8° is characteristic of metal glycolate (NiCo–glycolate).44–46 The thermal decomposition behavior of GO, NiCo–glycolate–rGO and NiCo–glycolate precursor samples was determined by TGA analysis as shown in Fig. S2† and 1A. For GO, there are three step weight losses (Fig. S2†), the first weight loss observed below 100 °C is due to the elimination of physisorbed water molecules. The second weight loss observed at 218 °C is ascribed to the removal of oxygen functional groups on the surface of GO. The third weight loss observed at 570 °C corresponds to the combustion of carbon skeleton of GO or rGO. For the precursor samples (Fig. 1A), a small weigh loss below 150 °C is attributed to the removal of adsorbed moisture while second weight loss observed between 200 °C to 350 °C corresponds to the decomposition of NiCo–glycolate–rGO/NiCo–glycolate precursors into NiCo2O4–rGO/NiCo2O4. The weight loss observed above 800 °C is ascribed to the complete decomposition of spinel NiCo2O4 structure.15 Further, the content of NiCo2O4 in the NiCo2O4–rGO composite is estimated to be 88.8% and rGO content is 12.2%. The powder XRD patterns of the GO, NiCo2O4–rGO and NiCo2O4 samples are shown in Fig. 1B. For GO, the diffraction peak at 2θ value of 10.5° observed is due to (002) plane of graphitic oxide with interplanar spacing of 0.84 nm. The presence of oxygen containing functional groups on the surface of GO increases the interplanar spacing from 0.34 nm for graphite to 0.84 nm for GO. Further, both NiCo2O4–rGO and NiCo2O4 samples show the reflections (shown in Fig. 1B) corresponding to face centered cubic spinel structure of NiCo2O4 (JCPDS card no. 73-1702) with space group of Fd3m (227). In both samples, no other impurity peaks of the NiCo2O4 were sensed, suggesting a complete conservation to NiCo2O4 without any impurity. In NiCo2O4–rGO composite, the reflection for GO or rGO is not observed, which may be due to insufficient carbon content in the composite.38,47
![]() | ||
Fig. 1 (A) TGA curves of unclacined precursors (Ni–Co–glycolate–rGO and Ni–Co–glycolate); (B) PXRD patterns of GO, NiCo2O4–rGO and NiCo2O4 samples. |
These samples are further analyzed by Raman and FT-IR to confirm the presence of rGO in the composite. Fig. 2A shows Raman spectra of GO, NiCo2O4–rGO and NiCo2O4 samples. Both GO and NiCo2O4–rGO composite samples show two broad peaks at about 1330 cm−1 (D-band) and 1583 cm−1 (G-band). The G-band corresponding to an E2g mode of graphite is associated to the vibration of sp2 bonded carbon atoms in a 2-D hexagonal lattice, whereas D-band is ascribed to the defects and disorders in the hexagonal graphitic carbon layers.29,30,34 In the Raman spectra of NiCo2O4–rGO and pristine NiCo2O4 samples, peaks observed at around 181, 456, 501 and 654 cm−1 could be attributed to F2g, Eg, F2g, and A1g of the phonon modes of NiCo2O4, which are related to the vibrations of Co–O and Ni–O bonds.48 Fig. 2B shows FT-IR spectra of GO, NiCo2O4–rGO and NiCo2O4 samples. The characteristic oxygenated functional groups of GO like CO (∼1728 cm−1) and C–OH (1398 cm−1) carboxylic groups, C–O–C (∼1051 cm−1) epoxy groups, and O–H (∼3400 cm−1) are observed in FT-IR spectra of GO. In addition, an absorption band at ∼1623 cm−1 due to the skeletal C
C vibration of unoxidized graphitic domains is evident. Except the C
C skeletal vibration, all other oxygenated functional groups greatly diminished or extinct in NiCo2O4–rGO composite indicating the reduction of GO into reduced graphene oxide (rGO) by ethylene glycol under solvothermal conditions. In the FT-IR spectra of NiCo2O4–rGO composite and pristine NiCo2O4 samples observed two strong peaks at lower frequencies (553 and 641 cm−1) can be assigned the stretching vibrations of the Ni–O and Co–O bonds in nickel cobalt oxide.38
The NiCo2O4–rGO composite is analyzed by photoemission spectroscopy to identify the chemical species which may contribute to the capacitive performance. Fig. S3† shows the X-ray photoelectron survey spectrum of the composite. It shows the photoemission signatures of Ni, Co, O and C. The deconvoluted spectral regions of C 1s, O 1s, Co 2p and Ni 2p of NiCo2O4–rGO composite are presented in Fig. 3. The C 1s region in Fig. 3A shows peaks at 287.9, 286.0 and 284.5 eV corresponding to oxygenated carbon species (CO), C–OH, and C
C/C–C, respectively.36,50 Accordingly the O 1s region in Fig. 3B shows three components at 529.7, 531.3, and 533 eV which are generally attributed to oxygen bonded to Ni and Co (Ni–O–Co),36 defect sites with minimum oxygen coordination or oxygenated carbon species, and O–H groups on the sample surface, respectively.34,50 The Co 2p spectrum in Fig. 3C consists of spin–orbit doublets and related satellite features characteristic of Co2+ and Co3+ ions.50 The Ni 2p spectrum also shows the spin–orbit doublets due to Ni2+ and Ni3+ ions along with intense shakeup satellites in Fig. 3D.34 Based on the percentage calculations, the sample contains higher amounts of Co3+ and Ni3+ ions than Co2+ and Ni2+ ions. The higher oxidation states are expected to facilitate fast charge transport across the electrode/electrolyte interface. These results show that the surface of NiCo2O4–rGO composite contains Co3+/2+ and Ni3+/2+ couples synergetic to each other and contribute to charge storage processes.34,36,50
![]() | ||
Fig. 3 Core level XPS spectra of (A) C 1s, (B) O 1s, (C) Co 2p and (D) Ni 2p for the NiCo2O4–rGO composite. |
The surface morphologies of the samples are analyzed by FESEM. The images of GO in Fig. S4† show typical homogeneous wrinkled flake-like morphology of very thin GO sheets. The morphology of NiCo2O4–rGO composite, however, resembles porous ultrathin nanosheets (Fig. 4A–C), which resulted from the thermal decomposition of NiCo–glycolate–rGO precursor. Similarly, pristine NiCo2O4 also developed ultrathin porous nanosheets as shown in Fig. 4D–F. The composition of NiCo2O4–rGO composite is further verified by energy X-ray dispersive spectroscopy (EDS) analysis as shown in Fig. S5.† The composite material contains Ni, Co, O, and C with Ni:
Co ratio is virtually 1
:
2. The microstructural properties of NiCo2O4–rGO composite was further analyzed by TEM. Fig. 5A and B shows different magnification TEM images of GO. This clearly shows that GO sheet is very thin, flat and flexible, which conforms that the complete exfoliation of graphite oxide into few layered GO nanosheets. Fig. 5C and D shows the TEM images of NiCo2O4–rGO composite. It clearly indicates that porous ultrathin nanosheets of NiCo2O4 grown on rGO sheets. And also nanosheets are made up by large number of interconnected NiCo2O4 nanoparticles with the uniform size. The interplanar spacing of 0.242 nm, 0.281 nm and 0.468 nm are shown in Fig. 5E, which are all well matched with (311), (220) and (111) planes of spinel NiCo2O4. The polycrystalline nature of NiCo2O4–rGO composite was confirmed by selected-area electron diffraction (SAED) patterns as shown in Fig. 5F. The SAED of NiCo2O4–rGO composite was composed of well-defined rings, the diffraction rings in the SAED patterns can be readily indexed to the planes of (111), (220), (311), (440) and (511) of spinel nickel cobaltite (JCPDS card no. 73-1702).
![]() | ||
Fig. 5 TEM images of GO (A and B) and NiCo2O4–rGO (C to E) at different magnifications. (F) The SAED pattern for the NiCo2O4–rGO composite. |
The pseudocapacitive performance of an electrode material is closely related to the distribution of the pore size and specific surface area of the electroactive material.9,10 The BET specific surface area and porosity of GO, NiCo2O4–rGO and NiCo2O4 samples are determined by N2 adsorption–desorption isotherms at 77 K using N2 as an adsorbent. Fig. 6 shows BET isotherms of NiCo2O4–rGO and NiCo2O4 samples, which characterize “type IV” isotherms with H3 hysteresis loops indicating the mesoporous nature of the respective samples. Inset of Fig. 6 shows the BJH pore size distribution plots for NiCo2O4–rGO and NiCo2O4 samples. The average pore diameters of these samples are found to be in mesoporous region, with tri-modal pore size distribution for NiCo2O4–rGO and bi-modal for NiCo2O4. However, the pore size distribution maxima are centered at 1.7, 4.5 and 22 nm for NiCo2O4–rGO composite, whereas those in pristine NiCo2O4 sample at 1.7 nm and 15 nm. The BET specific surface area for NiCo2O4–rGO composite and pristine NiCo2O4 samples are 84 m2 g−1 and 48 m2 g−1. However, the modified Hummers method route obtained GO nanostructure specific surface area shows 96 m2 g−1 (Fig. S6†), this is relatively higher than the NiCo2O4–rGO composite. The decreased surface area in the NiCo2O4–rGO composite is due to the preferential blocking of the pores by the large size nanocrystalline NiCo2O4. Further, the electrical conductivity of resulting NiCo2O4–rGO and pristine NiCo2O4 materials was measured by four-point probe method and estimated to be 5.23 s m−1 for NiCo2O4–rGO, and 1.21 s m−1 for pristine NiCo2O4. The high surface area and electrical conductivity of NiCo2O4–rGO nanocomposite may facilitate better transport of electrolyte ions through nanochannels and for efficient redox reactions during charge–discharge processes.
![]() | ||
Fig. 6 N2 adsorption–desorption isotherms of NiCo2O4–rGO and NiCo2O4 samples; inset shows corresponding BJH pore size distributions plots of NiCo2O4–rGO and NiCo2O4 samples. |
Trasatti analysis has been employed to evaluate the electrochemical charge storage properties of NiCo2O4–rGO and NiCo2O4 electrodes.51–53 In this analysis, the total voltammetric charge (qt) can be divided into two parts: (i) outer surface charge (qo) or more accessible surface charge, and (ii) inner surface charge (qi) or less accessible surface charge. The total charge exchanged between electrode and electrolyte, including both inner and outer charge storage as: qt = qi + qo. The inner surface charge is a diffusion controlled reaction process, which mainly originates from the less accessible surface such as pores, grain boundaries, crevices and cracks. The outer surface charges, qo storage is assumed to be not dependent on scan rates, it mainly initiated from the direct accessible surface, which corresponds to the region touching the electrolyte solution directly. In this case the reactive species diffusion does not control the reaction. Thus, the relationship of charge stored with scan rate is given as: qt = qo + kν−1/2, where ν is scan rate and k is a constant. The outer surface charge, (qo) can be derived from the extrapolation of voltammetric charge, q as ν = ∞ from the plot of q vs. ν−1/2 (Fig. 7C), and the total charge (qt) can be calculated from an extrapolation of voltammetric charge q at ν = 0 from the plot of q−1/2 vs. ν1/2 (Fig. 7D). Then the inner charge (qi) can be calculated from the subtraction of (qt) and (qo). The estimated outer and inner surface charges of the two samples are displayed in Fig. 8. Further, the ratio of (qo)/qt) for NiCo2O4–rGO composite and pristine NiCo2O4 electrodes are about 0.81 and 0.74, respectively. The higher ratio of (qo)/qt) for NiCo2O4–rGO electrode indicates that a large proportion of charge is located at easily accessible sites which are easily accessible with minimum diffusion limitations. This can impart high-rate capability to NiCo2O4–rGO electrode.
![]() | ||
Fig. 8 Inner and outer surface charge storage profiles of NiCo2O4–rGO and NiCo2O4 electrodes, obtained from cyclic voltammograms. |
Fig. 9A and B show galvanostatic charge–discharge (GCD) curves for NiCo2O4–rGO and NiCo2O4 electrodes at different current densities in the potential range between 0.05 to 0.55 V (vs. Hg/HgO). The nonlinearity in the charge–discharge curves indicates pseudocapacitance behavior owing to faradaic reactions occurring at the electrode/electrolyte interface. This is in contrast with the electric double layer electrode materials like GO which typically shows linear behavior (Fig. S7B†). Under similar current density conditions, NiCo2O4–rGO electrode exhibits longer discharge time than GO and NiCo2O4 electrodes (Fig. S9†). Therefore, NiCo2O4–rGO electrode shows superior charge storage performance and high specific capacitance than pristine NiCo2O4 electrode. The specific capacitance (CS, F g−1) of the resulting electrodes was calculated from the discharge curves at different current densities by using the following equation:35,54
Cycle life of supercapacitor is one of the most important parameter for the practical applications. Fig. 9D shows a cyclic life test for NiCo2O4–rGO and pristine NiCo2O4 electrodes at a current density of 4 A g−1. NiCo2O4–rGO composite electrode shows excellent cyclic stability with 90% capacitance retention over 5000 long term charge–discharge cycles. This is significantly better than the cyclic performance of reported NiCo2O4–rGO composites listed in Table 1. Overall, NiCo2O4–rGO composite electrode exhibits high specific capacitance, rate capability and superior cyclic performance. Such good performance can be attributed to (i) the presence of rGO in the NiCo2O4–rGO hybrid structure enhances the effective specific surface area and shorten the diffusion, migration paths for electrolyte ions (ii) the intimate contact between rGO and NiCo2O4 promoting electrical conductivity, and (iii) the rGO can effectively avoid the agglomeration of NiCo2O4 during charge–discharge process.29,32,33
Synthesis method | Electrode material | CSa (F g−1) | Cyclic performanceb | Ref. |
---|---|---|---|---|
a CS: maximum specific capacitance obtained from CV or CP test in F g−1.b Capacitance retention (%) measured by CP test at a specified current density. | ||||
Solvothermal (PVP surfactant) | NiCo2O4–rGO | 870@2 A g−1 | 90% after 5000 at 4 A g−1 | This work |
Electrodeposition | NiCo2O4/graphene | 1078@1 mA | 60% after 1000 at 3 mA | 24 |
Hydrothermal | rGO–NiCoO2 | 676@5 mV s−1 | No data | 36 |
Hydrothermal | NiCo2O4@RGO | 737@1 A g−1 | 94% after 3000 at 4 A g−1 | 38 |
Microwave | NiCo2O4@GO | 925@1.5 A g−1 | 94% after 700 at 16 A g−1 | 39 |
Sol–gel method | Graphene/NiO | 628@1 A g−1 | 82% after 1000 at 1 A g−1 | 47 |
Hydrothermal | Co3O4/GNS | 157.7@0.1 A g−1 | 70% after 4000 at 0.2 A g−1 | 55 |
Hydrothermal | NRGO–NiCoO2 | 508@0.5 A g−1 | 93% after 2000 at 2 A g−1 | 56 |
Reflux method | NiCo2O4/rGO | 1186@0.5 A g−1 | 97% after 100 at 0.5 A g−1 | 57 |
Electrochemical deposition | NiCo2O4/CNT | 694@1 A g−1 | 91% after 1500 at 4 A g−1 | 58 |
The kinetics and charge transfer mechanism at an electrode/electrolyte interface of both the materials are studied by electrochemical impedance spectroscopy (EIS). Fig. S7C† and 10A shows the Nyquist plots of GO, NiCo2O4–rGO and NiCo2O4 electrodes recorded in the frequency range of 10 mHz to 100 kHz at a bias potential of 0.3 V. The resulting impedance data is fitted by complex nonlinear least squares (CNLS) fitting method to the Randle equivalent circuit as shown in Fig. 10B. In this circuit, different parameters designate different electrochemical process occurring at an electrode/electrolyte interface. For both NiCo2O4–rGO and NiCo2O4 electrodes of the Nyquist plots were characterized by two distinct parts, a semi-circle loop at high-frequency and a linear line at low-frequency regions, suggesting different electrochemical phenomena during the electrochemical process on the electrode surface. At the high frequency region, the intersection with the real part (Z′) represents the combined resistance as a result of ionic resistance of an electrolyte, intrinsic resistance of the substrate and contact resistance at the active material/current collector.15 This value is almost same for both the electrode materials. However, the major difference is seen in the semicircles in which NiCo2O4–rGO electrode shows semicircle of smaller radius compared to pristine NiCo2O4 electrode. This shows that the resistance of NiCo2O4–rGO electrode is considerably less than that of NiCo2O4 electrode. The origin of this electrode resistance is the combination of the double layer capacitance (Cdl) on the oxide surface and ionic charge transfer resistance (Rict) due to faradaic reactions. In the Nyquist plots the inclined portion of the curve at lower frequency is attributed to the Warburg impedance (ZW) which is caused by the diffusion/transport of OH− ions within the pores of NiCo2O4–rGO and NiCo2O4 electrodes during the redox reactions. The near vertical lines along the imaginary axis (Z′′) in the lower frequency region indicate the capacitive behavior of the electrodes.
EIS provides complementary information about the frequency response of NiCo2O4–rGO and pristine NiCo2O4 electrode materials in supercapacitors and allows us to assess the capacitance changes with operating frequency. The complex form of the capacitance is dependent on frequency which can be defined as follows.3
C(ω) = C′(ω) + jC′′(ω) |
Footnote |
† Electronic supplementary information (ESI) available: Additional characterization data including the powder XRD pattern of uncalcined precursor, TGA, BET, FESEM, CV CP and analysis of GO, XPS survey spectra and EDS spectrum of NiCo2O4–rGO nanocomposite, comparison of CV curves for Ni foil, pristine NiCo2O4 and NiCo2O4–rGO, and comparison of CP curves for GO, NiCo2O4 and NiCo2O4–rGO. See DOI: 10.1039/c5ra11239g |
This journal is © The Royal Society of Chemistry 2015 |