Next Article in Journal
Single Crystal Diamond Deposited by Dual Radio-Frequency Plasma Jet CVD with High Growth Rate
Previous Article in Journal
Bis(triphenylphosphine)iminium Salts of Dioxothiadiazole Radical Anions: Preparation, Crystal Structures, and Magnetic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of NiMoO4/3D-rGO Nanocomposite in Alkaline Environments for Supercapacitor Electrodes

by
Shahrzad Arshadi Rastabi
1,
Rasoul Sarraf Mamoory
1,*,
Fatemeh Dabir
2,
Nicklas Blomquist
3,
Manisha Phadatare
3 and
Håkan Olin
3
1
Department of Materials Engineering, Tarbiat Modares University, 14115111 Tehran, Iran
2
Non-Metallic Materials Research Group, Niroo Research Institute, 15119-43943 Tehran, Iran
3
Department of Natural Sciences, Mid Sweden University, 85170 Sundsvall, Sweden
*
Author to whom correspondence should be addressed.
Submission received: 29 November 2018 / Revised: 23 December 2018 / Accepted: 3 January 2019 / Published: 9 January 2019

Abstract

:
Although Graphene oxide (GO)-based materials is known as a favorable candidate for supercapacitors, its conductivity needs to be increased. Therefore, this study aimed to investigate the performance of GO-based supercapicitor with new methods. In this work, an ammonia solution has been used to remove the oxygen functional groups of GO. In addition, a facile precipitation method was performed to synthesis a NiMoO4/3D-rGO electrode with purpose of using synergistic effects of rGO conductivity properties as well as NiMoO4 pseudocapacitive behavior. The phase structure, chemical bands and morphology of the synthesized powders were investigated by X-ray diffraction (XRD), Raman spectroscopy, and field emission secondary electron microscopy (FE-SEM). The electrochemical results showed that the NiMoO4/3D-rGO(II) electrode, where ammonia has been used during the synthesis, has a capacitive performance of 932 Fg−1. This is higher capacitance than NiMoO4/3D-rGO(I) without using ammonia. Furthermore, the NiMoO4/3D-rGO(II) electrode exhibited a power density of up to 17.5 kW kg−1 and an energy density of 32.36 Wh kg−1. These results showed that ammonia addition has increased the conductivity of rGO sheets, and thus it can be suggested as a new technique to improve the capacitance.

1. Introduction

Growing concerns about the serious environmental pollution and fossil-fuel energy crisis have attracted considerable attention to renewable and clean energy storage technologies [1]. Lithium-ion batteries, sodium-ion batteries, and supercapacitors (SCs) are currently potential energy storage devices [2]. Among them, SCs have competitive characteristics, such as high power density, long cycle life, fast charging/discharging rate [3], high safety, simple operating principles, and low maintenance costs [4,5]. Various kinds of materials have been selected to improve electrode properties. Although carbonaceous materials such as graphene and reduced GO can deliver high power density and possess high specific-area, good electrical conductivity, high chemical stability, and biocompatibility [6], faradaic electrode materials like metal oxides can provide higher capacitance and larger energy densities due to their redox reactions [7]. Recently, binary nickel-based oxides have attracted tremendous research interest because of their high theoretical capacitance and excellent electrochemical activity [8], arising from their improved electronic conductivity and surface redox reactions. Among various pseudo-active materials, NiMoO4 is an especially attractive candidate for pseudocapacitors, offering good stability in alkaline electrolytes, high specific capacitance and accessibility [9]. Therefore, the motivation to design and fabricate an NiMoO4 and rGO electrode with a novel hybrid nanostructure, which might combine the merits of both high specific capacitance originating from NiMoO4 and excellent rate capability provided by rGO, is well founded.
Intensive studies have been carried out on the synthesis of NiMoO4 with different morphologies such as nanorods [10], nanowires [11,12,13], nano-spheres [14], and nanosheets [15,16,17] for various application. Most of these studies used the hydrothermal method [8,10,11,12,13,14,15,18,19,20,21,22,23,24,25] that requires autoclave equipped with high-pressure and high-temperature, which finally resulted to the particles with large sizes and some of the papers used sonochemical [26], microwave [27], Sol-Gel [28], electrodeposition [17], and co-precipitation method [29,30]. Obviously, NiMoO4 synthesis is much less achieved by chemical precipitation than the hydrothermal method. We also synthesized NiMoO4 and graphene simultaneously for supercapacitor application, while those papers on chemical coprecipitation were exclusively for the synthesis of just NiMoO4 and its photocatalyst properties. In the current work, a simple and low-cost chemical precipitation method was used to synthesize NiMoO4/3D-rGO nanocomposites with no need of any specific facilities. In addition, this work studies the mechanism of formation of NiMoO4 nanoparticles on the surface of the graphene oxide and its reduction, which was not discussed in previous papers. Furthermore, there are several ways to reduce GO to rGO, one of which is the chemical reduction method. In previous works, various reduction agents such as hydrazine monohydrate [31], sodium borohydrate [32], ascorbic acid [33], ammonia borane [34], and heavy metal ions [35] have been used, while in the present research an ammonia solution has been applied.
In this work, we aim to synthesis the NiMoO4/3D-rGO electrode through a facile precipitation method for supercapacitors with particular emphasis on reduction of GO to rGO. Therefore, synergistic effects of ammonia, metal ions and low temperature have been used to eliminate large amount of GO oxygenated groups.

2. Experimental

2.1. Materials

Graphene oxide (GO), nickel nitrate hexahydrate (Ni(NO3)2·6H2O), sodium molybdate dihydrate (Na2MoO4·2H2O), and potassium hydroxide (KOH) were of reagent quality and were obtained from VWR, Sweden. Ammonia (28 wt. %-NH3) was purchased from VWR, Sweden. Polyvinylidene fluoride (PVDF, –(C2H2F2)n–), its solvent dimethylformamide (DMF, C3H7NO), and starch ((C6H10O5)n) were obtained from Sigma, Iran.

2.2. Preparation of the NiMoO4/3D-rGO(I) and NiMoO4/3D-rGO(II) Nanocomposites

Figure 1 briefly illustrates the fabrication of the NiMoO4/3D-rGO(II) nanocomposite through a simple precipitation process. The NiMoO4/3D-rGO composite was synthesized via a precipitation method in a starch environment as follows: First, 1 g of starch was added to 20 mL of distilled water at 80 °C to form a gelation suspension. Then, 1 mg of GO was dispersed in 20 mL distilled water under 5 min ultrasonication before being added to the starch solution to form a brown suspension. To prepare an alkaline environment, 3 mL ammonia (28 wt. %) was quickly added to the above suspension to adjust the pH about 10. Then, 5 mM Ni(NO3)2.6H2O and 5 mM Na2MoO4·2H2O were added to the above mixture in a dropwise manner and heated at 80 °C for 1 h, followed by 10 min of stirring. Subsequently, the powder was washed with distilled water to remove the impurities, and then dried at 60 °C for 24 h in a vacuum oven. Eventually, the final powder was calcinated in a tube furnace under Ar atmosphere at 350 °C for 2 h and then cooled to room temperature naturally to obtain the NiMoO4/3D-rGO(II) nanocomposite. The reduction of GO and the in situ formation of NiMoO4 were achieved simultaneously, causing the generation of NiMoO4/rGO. The same process has been done without the addition of ammonia (NH3) and the sample is called NiMoO4/3D-rGO(I).

2.3. Electrode Fabrication of NiMoO4, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) Nanocomposites

The electrodes were prepared by mixing active material, carbon black, and polyvinylidene fluoride (PVDF) with a mass ratio of 80:10:10 in DMF, followed by ultrasonication of the slurry for 10 min to form a uniform solution. The Ni foam was cleaned with HCl (10 wt. %) and ethanol during ultrasonication. Then, 2 mg·cm−2 of the mixture was pasted onto the Ni foam (1 × 1 cm2) as the working electrode before being dried in an oven at 70 °C for 24 h.

2.4. Material Characterization

In this research, the following analytical systems were used to characterize the prepared powder. Structural characterization of the samples were characterized using an X-ray powder diffraction (XRD, Bruker with CuKα radiation (λ = 0.1542 nm), Sundsvall, Sweden). Field-emission scanning electron microscopy (FE-SEM, MAIA3, TESCAN, Sundsvall, Sweden) was used to investigate the morphology and microstructure of the powder. Raman spectroscopy (Raman microscope, Horiba Xplora plus, laser excitation at 532 nm, Sundsvall, Sweden) in the frequency range of 50–3000 cm−1 was used to study the molecular bonds. A potentiostat system (VersaStat 4 with VersaStudio, Sundsvall, Sweden) was used to record the supercapacitor measurement.

3. Results and Discussion

X-ray diffraction analysis was used for studying the phase and structure of the samples. Figure 2 shows the XRD patterns of GO, rGO, NiMoO4 NPs, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) before and after calcination at 350 °C. X-ray diffraction spectra of the samples were carried out at 2θ = 10°–80° with a scanning rate of 0.02°/min. As can be seen, GO has a sharp peak at 2θ = 10.6° corresponded with the (002), which indicates the presence of the oxygen functional groups between the layers. This peak is disappeared in rGO pattern, implying the reduction of oxygen groups after heating at 350 °C. The XRD pattern of the NiMoO4 NPs shows high intensity peaks, which implies the crystallinity of NiMoO4 nanoparticles. This pattern is in accordance with Card No 00-045-0142 of JCPDS (JCPDS or ICDD -International Centre for Diffraction Data) data and related to the monoclinic phase of NiMoO4. The peaks at 2θ = 27.21°, 30.36°, and 43.9° which are the main characteristic peaks of the formation of the α-NiMoO4 phase, show that the sample after calcination is well crystallized and NiMoO4·nH2O has transformed to α-NiMoO4 [36]. Although the results indicate that most of the phases formed are related to the α-NiMoO4, a number of peaks at 2θ = 21.12°, 22.67°, and 32.07° signifies the presence of the orthorhombic phase of Mo17O47 (01-071-0566). These peaks are respectively attributed to the (510), (001), and (521) planes of Mo17O47 phase. In fact, Mo17O47 phase consists of seventeen MoO3 phases, which has lost some of its oxygen and formed a number of vacancies, and eventually produced Mo17O47 instead of Mo17O51. XRD pattern of the NiMoO4/3D-rGO(I) shows the exact pattern of α-NiMoO4, with no diffraction peak of GO at 2θ = 10.6°. The peak at 2θ = 22.67° has been broaden rather than α-NiMoO4 pattern, which is the same as rGO pattern. It can be concluded that in this case, the oxygenated groups of GO has been eliminated. The ammonia molecule is consisted of one nitrogen and three hydrogen atoms and due to the presence of lone pair on the nitrogen atom, this molecule is a strong source of electrons which is expected that it can eliminate oxygen groups of GO sheets. In order to investigate the role of the ammonia solution in the reduction process of GO, ammonia solution was added to GO suspension to adjust the pH about 10. After drying the sample at 60 °C without further heat treatment, its color changed from brown to black. According to some papers, changing in color can be the evidence of converting GO into rGO [37]. This sample is characterized before and after calcination at 350 °C. According to the XRD pattern of the NiMoO4/3D-rGO(II) nanocomposite before calcination, there is no peak of GO at 2θ = 10.6°, which indicates that before heat treatment, ammonia caused reduction of GO sheets. Furthermore, the XRD pattern of the NiMoO4/3D-rGO(II) nanocomposite after calcination reveals a peak at 43.3°, which corresponds to N-rGO, based on data from [38]. It reveals that nitrogen of ammonia has been doped into the rGO network at 350 °C. In addition, the peak at 2θ = 23.1° (d ~ 3.4 Å), while the GO has a peak at 2θ = 10.6° (d ~ 8.3 Å). The large space between the layers in GO is attributed to the presence of hydrocarbons, the oxygen-containing groups, and also the H2O molecules at the edges and the distance between the sheets. In other words, decrease in the space between layers of graphene sheets in the NiMoO4/3D-rGO(II) sample is a reason for the successful reduction of the GO sheets [39]. Results show that oxygen-containing groups can easily accept free electrons of ammonia to separate the carbon-oxygen bond, repaired π bands and form carbon radicals on GO sheets, and lead to the formation of rGO [37].
To compare the reduction rate of GO in the NiMoO4/3D-rGO(II) nanocomposite with that of the NiMoO4/3D-rGO(I) nanocomposite, the Raman spectra were measured for GO, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) nanocomposites (shown in Figure 2). The Raman spectra of the pure GO (Figure 3a) shows that the D and G peaks are located at 1347.64 cm−1 and 1574.98 cm−1, respectively. The relative intensity ratio of the peaks (ID/IG) is a measure of the disorder degree, and its value is approximately 0.98 for GO [23]. In both spectra of the NiMoO4/3D-rGO(I) and NiMoO4/3D-rGO(II) nanocomposites, the characteristic peaks at 969 cm−1 (Mo=O), 928 cm−1 (Mo=O), and 697 cm−1 (Ni–O–Mo) and some medium-intensity peaks at 394 cm−1 (Mo–O) and 325 cm−1 (Mo–O–Mo) belong to the α-phase NiMoO4 [10,40,41] according to NiMoO4 spectra (Figure 3a). The Raman spectra of the NiMoO4/3D-rGO(I) nanocomposite (Figure 3b) shows that after chemical reduction of GO to rGO, the ID/IG ratio slightly increased from 0.98 (GO) to 1.02 (rGO), due to the increase in the amount of defects in rGO. In comparison to the NiMoO4/3D-rGO(I), the ID/IG ratio of NiMoO4/3D-rGO(II) nanocomposite (Figure 3c) was calculated to be approximately 1.049, which is higher than that of the NiMoO4/3D-rGO(I) nanocomposite (1.02). This finding shows that ammonia reduced more oxygenated groups in the NiMoO4/3D-rGO(II) sample. In addition, the Raman spectra of the NiMoO4/3D-rGO(II) nanocomposite (Figure 3c) shows that the D and G peaks redshifted to the wavenumbers at approximately 1344.89 cm−1 and 1574.98 cm−1, respectively, compared to 1339.39 cm−1 and 1577.65 cm−1 in NiMoO4/3D-rGO(I). The Raman results confirm that although the GO has been reduced in the presence of NiMoO4 NPs, using an ammonia solution during synthesis results in more reduction of GO. In addition, the higher intensity of ID/IG in the NiMoO4/3D-rGO(II) nanocomposite shows more defects and disorders in the rGO structure due to the removal of more functional groups on the surface of GO, which results in higher conductivity and provides more active sites for electron storage [23,42].
The morphologies and microstructure of the prepared NiMoO4/3D-rGO(I) and NiMoO4/rGO(II) nanocomposites were investigated by SEM (Figure 4). The SEM images of the both nanocomposites show a 3D interconnected network consisting of NiMoO4 nanorods and rGO sheets. According to the SEM images of NiMoO4/3D-rGO(I) nanocomposite, it seems that graphene oxide sheets were the nucleation centers for growing NiMoO4 nanorods and they have been anchored the surface of reduced graphene sheets. This probably happened due to the strong electrostatic interaction between the nanorods and the reduced graphene oxide. Whereas the SEM images of NiMoO4/3D-rGO(II) nanocomposite consisted of two different phases. It shows thick rGO sheets which can be interpret as graphite sheets. This phenomena can be attributed to the ammonia effect on reduction of the sheets. Another phase is NiMoO4 nanorods which due to magnetic interaction has gathered together and made a separate porous structure. The results showed that starch as an environmentally friendly surfactant prevented the agglomeration of particles and eventually turned into gases (CO-CO2) at 350 °C and resulted in a porous structure nanocomposite. Such a nanostructured morphology can provide more voids and a short diffusion path between the electrode material and the electrolyte which increases electrolyte diffusion and accessibility to reaction sites for more energy storage [23].
A possible formation mechanism for the synthesis of the NiMoO4/3D-rGO(I) nanocomposite is schematically shown in Figure 5. It is well known that, –COOH or –OH oxygen-containing functional groups of GO sheets are the nucleation centers for combining Ni2+ and MoO42− ions to form NiMoO4. It has been reported that during the adsorption process, the release of the proton (H+) of –COOH or –OH functional groups to the solution results in a decrease of the equilibrium pH of the solution [43]. It can be concluded that the exchange of Ni2+ and H+ on –COOH or –OH oxygenated groups can be considered to be a mechanism for the adsorption of Ni2+ onto GO, and the formation of a bridge between different GO sheets. The relevant reactions in the current study can be expressed as follows:
—COOH + Ni2+ → —COONi2+ + H+
—OH + Ni2+ → —ONi2+ + H+
—2COOH + Ni2+ → —COONiOOC— + 2H+
—2OH + Ni2+ → —ONiO— + 2H+
—COOH + Ni2+ + —OH → —COONiO— + 2H+.

Electrochemical Studies

The electrochemical measurements of the samples were performed in a three-electrode cell configuration at room temperature with 3 M KOH electrolyte. In this system, a platinum sheet and a saturated calomel Hg/Hg2Cl2 electrode were used as the counter and the reference electrodes, respectively. Cyclic voltammetry (CV) analysis was evaluated within a potential range of 0–0.7 V at different scan rates (10–100 mVs−1). Figure 6a shows the CV curves of the Ni foam, NiMoO4 NPs, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) nanocomposites. The redox peaks confirm that the charge-storage mechanism is mainly Faradaic (battery-type) and related to reversible chemical reactions of Ni(II) ↔ Ni(III) + e [44] which can be expressed as follows [3,45]:
Ni(OH)2 + OH ↔ NiOOH + H2O + e
NiOOH + OH ↔ NiO2 + H2O + e.
Mo atoms do not take part in the redox reaction. The main function of Mo element was to improve the conductivity of Nickel molybdate, thereby contributing in increasing specific capacitance and it does not show any redox peak in CV curves [46,47]. As seen, the area surrounded by the CV curve of the NiMoO4/3D-rGO(II) hybrid electrode is larger and the redox current is higher than that of the NiMoO4/3D-rGO(I) electrode, indicating much more specific capacitance of the NiMoO4/3D-rGO(II) electrode. More capacitance of NiMoO4/3D-rGO(II) supports the fact that removing more oxygenated surface functional groups of GO can provide more active sites for storing energy and increasing conductivity. Furthermore, the potential difference (ΔE) between the reduction and oxidation peaks of the NiMoO4/3D-rGO(II) composite is 160 mV, which is approximately 81 mV smaller than that of the NiMoO4/3D-rGO(I) electrode (241 mV, Figure 6a), indicating better electrochemical reversibility of NiMoO4/3D-rGO(II) due to more reduction of GO sheets in the presence of ammonia [48].
Figure 6b illustrates the CV curves of the NiMoO4/3D-rGO(II) electrode at the scan rates of 10–100 mVs−1 in the potential window of 0–0.7 V. With increased scan rates, the anodic and cathodic currents increase, suggesting fast redox reactions at the electrode/electrolyte contact surface at high scan rates, which results in lower specific capacitance [22]. Due to the presence of graphene oxide sheets, in addition to the Faradaic and electrochemical reactions of NiMoO4, formation of the electrical double-layer can lead to more charge storage in the system [49]. According to the resulting curves, it is noteworthy that by increasing the scan rate, even at the scan rate of 100 mVs−1, redox peaks are still observed, indicating good chemical stability and conductivity in the hybrid structure of the electroactive materials [50].
Galvanostatic charge–discharge (GCD) tests were further performed to evaluate the capacitance of the electrodes at current densities of 1–50 Ag−1. Figure 6c illustrates the GCD curves of the NiMoO4/3D-rGO(II) electrode between 0 V and 0.5 V. A pair of voltage plateaus appear at ~0.29 and ~0.23 V and nonlinear behavior of the GCD curves, revealing the pseudocapacitive features of Ni2+ ions of the NiMoO4/3D-rGO(II) electrode [18]. According to the three-electrode system, the specific capacitance of the single electrode can be estimated from the cyclic voltammetry (CV) curves and discharge curves with the Equations (8) and (9), respectively [51,52].
C s = 1 m ν   Δ V IdV
C s = I Δ t m   Δ V
where ν is the scan rate, I is the discharge current, Δt the discharge time, m is the mass of the active material loaded on the electrode surface, and ΔV refers to the potential range [51,52].
Stability behavior of the electrode was performed over repeated charge and discharge cycling. Therefore, GCD measurement for the NiMoO4/3D-rGO(II) electrode were carried out at a current density of 2 Ag−1 for approximately 500 cycles in 3 M KOH electrolyte (Figure 6d). Initially, the NiMoO4/3D-rGO(II) electrode showed a significant decrease in the capacitance up to first 200 cycles. However, after 200 cycles the capacitance decreased slowly and at the end of 500 cycles, the capacitance dropped to 76% of its initial value. Good cycle stability of the NiMoO4/3D-rGO(II) is due to the porous structure with NiMoO4 rod morphology and low amount of oxygenated groups of GO after reduction in an alkaline environment.
Figure 6e illustrates the electrochemical impedance spectra of the NiMoO4/3D-rGO(II) electrode before and after 500 cycles over the frequency range of 0.1 Hz to 100 kHz by applying a 10 mV-AC voltage. The Nyquist plots are usually consisted of two segments; the small semicircle in the high frequency region and the approximately straight line in the low frequency region. The diameter of the semicircle corresponds to the charge transfer resistance (Rct) during Faradaic redox reaction. As can be seen, the Nyquist plots of both spectra show a negligible semi-circle in the high frequency region because of fast faradaic reactions, implying low Rct and good charge transfer [53]. The intercept on the real axis at the beginning of the semicircle part determine the ESR (Rs) which is combination of electrode material resistance, ionic resistance of electrolyte and contact resistance between current collector and electrode material [54]. ESR for NiMoO4/3D-rGO(II) electrode before cycling is 0.82 and after cycling the ESR value has increased to 0.9. Moreover, the linear line in low frequency region implies Warburg impedance (ZW) attributed to OH ions diffusion of electrolyte. The more vertical the line, the closer to ideal capacitive behavior of the nanocomposite [53]. In the EIS plot of NiMoO4/3D-rGO(II) electrode before 500 cycles, the straight line leans more towards the vertical Z″ axis, indicating a better capacitive performance and lower ion diffusion resistance of this electrode in contrast to NiMoO4/3D-rGO(II) electrode after 500 cycles.

4. Conclusions

A simple precipitation approach is demonstrated to synthesize the NiMoO4 NPs, NiMoO4/3D-rGO(I) and NiMoO4/3D-rGO(II) nanocomposite electrodes. The results showed that GO sheets were successfully reduced due to the disappearance of 2θ = 10.6° peak in XRD patterns and increase of ID/IG ratio in Raman spectrum. It can be concluded that the synergistic effect of the ammonia, metal ions and temperature benefited the deoxygenation of GO sheets. In general, addition of ammonia provided the possibility of GO deoxygenating and subsequent thermal treatment will restore the C=C bond to convert GO to reduced graphene oxide. The NiMoO4/3D-rGO(I) and NiMoO4/3D-rGO(II) electrodes demonstrated specific capacitances of 790 Fg−1 and 932 Fg−1, at a current density of 1 Ag−1. An improved capacitive performance of the NiMoO4/3D-rGO(II) electrode is attributed to the pseudocapacitance behavior of NiMoO4 and the elimination of oxygen-containing groups which will be beneficial to the increase in the electronic conductivity of resulting rGO. In addition, the NiMoO4/3D-rGO(II) electrode showed a good cycle life with 76% retention of the initial specific capacitance after 500 cycles.

Author Contributions

S.A.R.; Investigation, Synthesize and Writing—original draft, R.S.M.; Supervision, F.D.; Software, N.B.; Editing and Formal analysis, M.P.; Software, and H.O.; Supervision and editing.

Funding

This research received no external funding.

Acknowledgments

The authors acknowledge Tarbiat Modares University and Mid Sweden University for providing the facilities and technical assistance for this research. We thank the honorable supervisors, Magnus Hummelgård, and all of the personnel who work in the research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zuo, W.; Li, R.; Zhou, C.; Li, Y.; Xia, J.; Liu, J. Battery-Supercapacitor Hybrid Devices: Recent Progress and Future Prospects. Adv. Sci. 2017, 4, 1600539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sun, F.; Gao, J.; Zhu, Y.; Pi, X.; Wang, L.; Liu, X.; Qin, Y. A high performance lithium ion capacitor achieved by the integration of a Sn-C anode and a biomass-derived microporous activated carbon cathode. Sci. Rep. 2017, 7, 40990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Yan, A.L.; Wang, X.C.; Cheng, J.P. Research Progress of NiMn Layered Double Hydroxides for Supercapacitors: A Review. Nanomaterials 2018, 8, 747. [Google Scholar] [CrossRef]
  4. Chiam, S.L.; Lim, H.N.; Hafiz, S.M.; Pandikumar, A.; Huang, N.M. Electrochemical Performance of Supercapacitor with Stacked Copper Foils Coated with Graphene Nanoplatelets. Sci. Rep. 2018, 8, 3093. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, L.; Zhang, W.; Xiang, J.; Xu, H.; Li, G.; Huang, Y. Hierarchical core-shell NiCo2O4@NiMoO4 nanowires grown on carbon cloth as integrated electrode for high-performance supercapacitors. Sci. Rep. 2016, 6, 31465. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, C.; Zhou, E.; He, W.; Deng, X.; Huang, J.; Ding, M.; Wei, X.; Liu, X.; Xu, X. NiCo2O4-Based Supercapacitor Nanomaterials. Nanomaterials 2017, 7, 41. [Google Scholar] [CrossRef] [PubMed]
  7. Owusu, K.A.; Qu, L.; Li, J.; Wang, Z.; Zhao, K.; Yang, C.; Zhou, L. Low-crystalline iron oxide hydroxide nanoparticle anode for high-performance supercapacitors. Nat. Commun. 2017, 8, 14264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Xue, W.D.; Wang, W.J.; Fu, Y.F.; He, D.X.; Zeng, F.Y.; Zhao, R. Rational synthesis of honeycomb-like NiCo2O4@NiMoO4 core/shell nanofilm arrays on Ni foam for high-performance supercapacitors. Mater. Lett. 2017, 186, 34–37. [Google Scholar] [CrossRef]
  9. Xiong, X.; Ding, D.; Chen, D.; Waller, G.; Bu, Y.; Wang, Z.; Liu, M. Three-dimensional ultrathin Ni(OH)2 nanosheets grown on nickel foam for high- performance supercapacitors. Nano Energy 2015, 11, 154–161. [Google Scholar] [CrossRef]
  10. Jothi, P.R.; Kannan, S.; Velayutham, G. Enhanced methanol electro-oxidation over in-situ carbon and graphene supported one dimensional NiMoO4 nanorod. Power Sources 2015, 277, 350–359. [Google Scholar] [CrossRef]
  11. Guo, D.; Zhang, P.; Zhang, H.; Yu, X.; Zhu, J.; Li, Q.; Wang, T. NiMoO4 nanowires supported on Ni foam as novel advanced electrodes for supercapacitors. J. Mater. Chem. A 2013, 1, 9024–9027. [Google Scholar] [CrossRef]
  12. Guo, D.; Luo, Y.; Yu, X.; Li, Q.; Wang, T. High performance NiMoO4 nanowires supported on carbon cloth as advanced electrodes for symmetric supercapacitors. Nano Energy 2014, 8, 174–182. [Google Scholar] [CrossRef]
  13. Liu, X.; Zhang, K.; Yang, B.; Song, W.; Liu, Q.; Jia, F.; Qin, S.; Chen, W.; Li, J.; Zhang, Z. Three-Dimensional Graphene Supported Nickel Molybdate Nanowires as Novel Ultralight and Flexible Electrode for Supercapacitors. Mater. Lett. 2016, 164, 401–404. [Google Scholar] [CrossRef]
  14. Cai, D.; Wang, D.; Liu, B.; Wang, Y.; Liu, Y.; Wang, L.; Li, H.; Huang, Ch.; Li, Q.; Wang, T. Comparison of the Electrochemical Performance of NiMoO4 Nanorods and Hierarchical Nanospheres for Supercapacitor Applications. ACS Appl. Mater. Interfaces 2013, 5, 12905–12910. [Google Scholar] [CrossRef] [PubMed]
  15. Cai, D.; Liu, B.; Wang, D.; Liu, Y.; Wang, L.; Li, H.; Wang, Y.; Wang, C.; Li, Q.; Wang, T. Enhanced performance of supercapacitors with ultrathin mesoporous NiMoO4 nanosheets. Electrochim. Acta 2014, 125, 294–301. [Google Scholar] [CrossRef]
  16. Wu, F.; Liao, Q.; Cao, F.; Li, L.; Zhang, Y. Non-noble bimetallic NiMoO4 nanosheets integrated Si photoanodes for highly efficient and stable solar water splitting. Nano Energy 2017, 34, 8–14. [Google Scholar] [CrossRef]
  17. Xiao, K.; Xia, L.; Liu, G.; Wang, S.; Ding, L.X.; Wang, H. Honeycomb-like NiMoO4 Ultrathin Nanosheet Arrays for High-performance Electrochemical Energy Storage. J. Mater. Chem. A 2015, 3, 6128–6135. [Google Scholar] [CrossRef]
  18. Jinlong, L.; Miura, H.; Meng, Y. A novel mesoporous NiMoO4@rGO nanostructure for supercapacitor applications. Mater. Lett. 2017, 194, 94–97. [Google Scholar] [CrossRef]
  19. Zhang, P.; Zhou, J.; Chen, W.; Zhao, Y.; Mu, X.; Zhang, Z.; Pan, X.; Xie, E. Constructing highly-efficient electron transport channels in the 3D electrode materials for high-rate supercapacitors: The case of NiCo2O4@NiMoO4 hierarchical nanostructures. Chem. Eng. Sci. 2017, 307, 687–695. [Google Scholar] [CrossRef]
  20. Huang, L.; Xiang, J.; Zhang, W.; Chen, C.; Xu, H.; Huang, Y. 3D interconnected porous NiMoO4 nanoplate arrays on Ni foam as high-performance binder-free electrode for supercapacitors. J. Mater. Chem. A 2015, 3, 22081–22087. [Google Scholar] [CrossRef]
  21. Nti, F.; Anang, D.A.; Han, J.I. Facilely synthesized NiMoO4/CoMoO4 nanorods as electrode material for high performance supercapacitor. Alloy Compd. 2018, 742, 342–350. [Google Scholar] [CrossRef]
  22. Lin, L.; Liu, T.; Liu, J.; Sun, R.; Hao, J.; Ji, K.; Wang, Z. Facile synthesis of groove-like NiMoO4 hollow nanorods for high-performance supercapacitors. Appl. Surf. Sci. 2016, 360, 234–239. [Google Scholar] [CrossRef]
  23. Li, Y.; Jian, J.; Fan, Y.; Wang, H.; Yu, L.; Cheng, G.; Sun, M. Facile one-pot synthesis of NiMoO4/reduced graphene oxide composite as pseudocapacitors with superior performance. RSC Adv. 2016, 6, 69627–69633. [Google Scholar] [CrossRef]
  24. Bakhshali-Dehkordy, R.; Aghajani, Z. Hydrothermal-Assisted Synthesis of TiO2@NiMoO4 Nanocomposites and Evaluation of Their Photocatalysis Properties. J. Electron. Mater. 2018, 48, 278–285. [Google Scholar] [CrossRef]
  25. Kannan, V.; Kim, H.J.; Park, H.C.; Kim, H.S. Single-Step Direct Hydrothermal Growth of NiMoO4 Nanostructured Thin Film on Stainless Steel for Supercapacitor Electrodes. Nanomaterials 2018, 8, 563. [Google Scholar] [CrossRef] [PubMed]
  26. Kianpour, G.; Salavati-Niasari, M.; Emadi, H. Sonochemical synthesis and characterization of NiMoO4 nanorods. Ultrason. Sonochem. 2013, 20, 418–424. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, T.; Chai, H.; Jia, D.; Su, Y.; Wang, T.; Zhou, W. Rapid microwave-assisted synthesis of mesoporous NiMoO4 nanorod/reduced graphene oxide composites for high-performance supercapacitors. Electrochim. Acta 2015, 180, 998–1006. [Google Scholar] [CrossRef]
  28. Umapathy, V.; Neeraja, P.; Manikandan, A.; Ramu, P. Synthesis of NiMoO4 nanoparticles by sol–gel method and their structural, morphological, optical, magnetic and photocatlytic properties. Trans. Nonferr. Met. Soc. China 2017, 27, 1785–1793. [Google Scholar] [CrossRef]
  29. Liu, M.C.; Kong, L.B.; Lu, C.; Ma, X.J.; Li, X.M.; Luo, Y.C.; Kang, L. Design and synthesis of CoMoO4–NiMoO4·xH2O bundles with improved electrochemical properties for supercapacitors. J. Mater. Chem. A 2013, 1, 1380–1387. [Google Scholar] [CrossRef]
  30. Kianpour, G.; Soofivand, F.; Badiei, M.; Salavati-Niasari, M.; Hamadanian, M. Facile synthesis and characterization of nickel molybdate nanorods as an effective photocatalyst by co-precipitation method. J. Mater. Sci. Mater. Electron. 2016, 27, 10244–10252. [Google Scholar] [CrossRef]
  31. Chua, C.K.; Pumera, M. The reduction of graphene oxide with hydrazine: Elucidating its reductive capability based on a reaction-model approach. Chem. Commun. 2016, 52, 72–75. [Google Scholar] [CrossRef] [PubMed]
  32. Yang, Z.Z.; Zheng, Q.B.; Qiu, H.X.; Jing, L.I.; Yang, J.H. A simple method for the reduction of graphene oxide by sodium borohydride with CaCl2 as a catalyst. New Carbon Mater. 2015, 30, 41–47. [Google Scholar] [CrossRef]
  33. De Silva, K.K.H.; Huang, H.H.; Yoshimura, M. Progress of reduction of graphene oxide by ascorbic acid. Appl. Surf. Sci. 2018, 447, 338–346. [Google Scholar] [CrossRef]
  34. Zhuo, Q.; Zhang, Y.; Du, Q.; Yan, C. Facile reduction of graphene oxide at room temperature by ammonia borane via salting out effect. J. Colloid Interface Sci. 2015, 457, 243–247. [Google Scholar] [CrossRef] [PubMed]
  35. Devi, M.M.; Sahu, S.R.; Mukherjee, P.; Sen, P.; Biswas, K. Graphene: A self-reducing template for synthesis of graphene–nanoparticles hybrids. RSC Adv. 2015, 5, 62284–62289. [Google Scholar] [CrossRef]
  36. Bankar, P.K.; Ratha, S.; More, M.A.; Late, D.J.; Rout, C.S. Enhanced field emission performance of NiMoO4 nanosheets by tuning the phase. Appl. Surf. Sci. 2017, 418, 270–274. [Google Scholar] [CrossRef]
  37. Feng, H.; Cheng, R.; Zhao, X.; Duan, X.; Li, J. A low-temperature method to produce highly reduced graphene oxide. Nat. Commun. 2013, 4, 1539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Du, M.; Sun, J.; Chang, J.; Yang, F.; Shi, L.; Gao, L. Synthesis of nitrogen-doped reduced graphene oxide directly from nitrogen-doped graphene oxide as a high-performance lithium ion battery anode. RSC Adv. 2014, 4, 42412–42417. [Google Scholar] [CrossRef]
  39. Soni, M.; Kumar, P.; Kumar, R.; Sharma, S.K.; Soni, A. Photo-catalytic reduction of oxygenated graphene dispersions for supercapacitor applications. J. Phys. D Appl. Phys. 2017, 50, 124003–124011. [Google Scholar] [CrossRef]
  40. Ezeigwe, E.R.; Khiew, P.S.; Siong, C.W.; Tan, M.T. Synthesis of NiMoO4 nanorods on graphene and superior electrochemical performance of the resulting ternary based composites. Ceram Int. 2017, 43, 13772–13780. [Google Scholar] [CrossRef]
  41. Ghosh, D.; Giri, S.; Das, C.K. Synthesis, characterization and electrochemical performance of graphene decorated with 1D NiMoO4.nH2O nanorods. Nanoscale 2013, 5, 10428–10437. [Google Scholar] [CrossRef]
  42. Liu, H.; Zhang, G.; Zhou, Y.; Gao, M.; Yang, F. One-step potentiodynamic synthesis of poly (1,5-diaminoanthraquinone)/reduced graphene oxide nanohybrid with improved electrocatalytic activity. Mater. Chem. A 2013, 1, 13902–13913. [Google Scholar] [CrossRef]
  43. Peng, W.; Li, H.; Liu, Y.; Song, S. A review on heavy metal ions adsorption from water by graphene oxide and its composites. Mol. Liq. 2017, 230, 496–504. [Google Scholar] [CrossRef]
  44. Li, Y.; Jian, J.; Xiao, L.; Wang, H.; Yu, L.; Cheng, G.; Sun, M. Synthesis of NiMoO4 nanosheets on graphene sheets as advanced supercapacitor electrode materials. Mater. Lett. 2016, 184, 21–24. [Google Scholar] [CrossRef]
  45. Rajkumar, M.; Hsu, C.T.; Wu, T.H.; Chen, M.G.; Hu, C.C. Advanced materials for aqueous supercapacitors in the asymmetric design. Prog. Nat. Sci. Mater. 2015, 25, 527–544. [Google Scholar] [CrossRef] [Green Version]
  46. Jiang, G.; Zhang, M.; Li, X.; Gao, H. NiMoO4@ Ni(OH)2 core/shell nanorods supported on Ni foam for high-performance supercapacitors. RSC Adv. 2015, 85, 69365–69370. [Google Scholar] [CrossRef]
  47. Budhiraju, V.S.; Kumar, R.; Sharma, A.; Sivakumar, S. Structurally stable hollow mesoporous graphitized carbon nanofibers embedded with NiMoO4 nanoparticles for high performance asymmetric supercapacitors. Electrochim. Acta 2017, 238, 337–348. [Google Scholar] [CrossRef]
  48. Yang, J.; Liu, W.; Niu, H.; Cheng, K.; Ye, K.; Zhu, K.; Yan, J. Ultrahigh energy density battery-type asymmetric supercapacitors: NiMoO4 nanorod-decorated graphene and graphene/Fe2O3 quantum dots. Nano Res. 2018, 11, 4744–4758. [Google Scholar] [CrossRef]
  49. Wei, C.; Huang, Y.; Yan, J.; Chen, X.; Zhang, X. Synthesis of hierarchical carbon sphere@NiMoO4 composite materials for supercapacitor electrodes. Ceram Int. 2016, 42, 15694–15700. [Google Scholar] [CrossRef]
  50. Zhang, Z.; Liu, Y.; Huang, Z.; Ren, L.; Qi, X.; Wei, X.; Zhong, J. Facile hydrothermal synthesis of NiMoO4@CoMoO4 hierarchical nanospheres for supercapacitor applications. Phys. Chem. Chem. Phys. 2015, 17, 20795–20804. [Google Scholar] [CrossRef]
  51. Yang, S.; Zhang, K. Converting Corncob to Activated Porous Carbon for Supercapacitor Application. Nanomaterials 2018, 8, 181. [Google Scholar] [CrossRef] [PubMed]
  52. Krishnan, S.G.; Ab Rahim, M.H.; Jose, R. Synthesis and characterization of MnCo2O4 cuboidal microcrystals as a high performance psuedocapacitor electrode. Alloy. Compd. 2016, 656, 707–713. [Google Scholar] [CrossRef]
  53. Fang, L.; Wang, F.; Zhai, T.; Qiu, Y.; Lan, M.; Huang, K.; Jing, Q. Hierarchical CoMoO4 nanoneedle electrodes for advanced supercapacitors and electrocatalytic oxygen evolution. Electrochim. Acta 2018, 259, 552–558. [Google Scholar] [CrossRef]
  54. Miankushki, H.N.; Sedghi, A.; Baghshahi, S. Facile and scalable fabrication of graphene/polypyrrole/MnOx/Cu(OH)2 composite for high-performance supercapacitors. J. Solid State Chem. 2018, 22, 3317–3329. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the synthesis process of NiMoO4/3D-rGO(I).
Figure 1. Schematic diagram of the synthesis process of NiMoO4/3D-rGO(I).
Crystals 09 00031 g001
Figure 2. The XRD pattern of GO, rGO, NiMoO4 NPs, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) before and after heating at 350 °C.
Figure 2. The XRD pattern of GO, rGO, NiMoO4 NPs, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) before and after heating at 350 °C.
Crystals 09 00031 g002
Figure 3. Raman spectra of the NiMoO4, pure GO, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) nanocomposites.
Figure 3. Raman spectra of the NiMoO4, pure GO, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) nanocomposites.
Crystals 09 00031 g003
Figure 4. SEM images of (a,b) NiMoO4/3D-rGO(I) and (c,d) NiMoO4/rGO(II) nanocomposites with two different magnifications.
Figure 4. SEM images of (a,b) NiMoO4/3D-rGO(I) and (c,d) NiMoO4/rGO(II) nanocomposites with two different magnifications.
Crystals 09 00031 g004
Figure 5. Schematic illustration of the fabrication process of hybrid NiMoO4/3D-rGO(I) nanocomposite.
Figure 5. Schematic illustration of the fabrication process of hybrid NiMoO4/3D-rGO(I) nanocomposite.
Crystals 09 00031 g005
Figure 6. (a) Cyclic voltammetry (CV) curves of the Ni foam, NiMoO4 NPs, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) nanocomposites at scan rates of 10 mVs−1; (b) CV curves of the NiMoO4/3D-rGO(II) nanocomposite at different scan rates of 10–100 mVs−1; (c) GCD curves of the NiMoO4/3D-rGO(II) nanocomposite at various current densities (1–50 Ag−1); and (d) cyclic stability of the NiMoO4/3D-rGO(II) electrode after 1000 cycles; and (e) The Nyquist plots of the NiMoO4/3D-rGO(II) electrode before and after 500 cycles.
Figure 6. (a) Cyclic voltammetry (CV) curves of the Ni foam, NiMoO4 NPs, NiMoO4/3D-rGO(I), and NiMoO4/3D-rGO(II) nanocomposites at scan rates of 10 mVs−1; (b) CV curves of the NiMoO4/3D-rGO(II) nanocomposite at different scan rates of 10–100 mVs−1; (c) GCD curves of the NiMoO4/3D-rGO(II) nanocomposite at various current densities (1–50 Ag−1); and (d) cyclic stability of the NiMoO4/3D-rGO(II) electrode after 1000 cycles; and (e) The Nyquist plots of the NiMoO4/3D-rGO(II) electrode before and after 500 cycles.
Crystals 09 00031 g006

Share and Cite

MDPI and ACS Style

Arshadi Rastabi, S.; Sarraf Mamoory, R.; Dabir, F.; Blomquist, N.; Phadatare, M.; Olin, H. Synthesis of NiMoO4/3D-rGO Nanocomposite in Alkaline Environments for Supercapacitor Electrodes. Crystals 2019, 9, 31. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9010031

AMA Style

Arshadi Rastabi S, Sarraf Mamoory R, Dabir F, Blomquist N, Phadatare M, Olin H. Synthesis of NiMoO4/3D-rGO Nanocomposite in Alkaline Environments for Supercapacitor Electrodes. Crystals. 2019; 9(1):31. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9010031

Chicago/Turabian Style

Arshadi Rastabi, Shahrzad, Rasoul Sarraf Mamoory, Fatemeh Dabir, Nicklas Blomquist, Manisha Phadatare, and Håkan Olin. 2019. "Synthesis of NiMoO4/3D-rGO Nanocomposite in Alkaline Environments for Supercapacitor Electrodes" Crystals 9, no. 1: 31. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9010031

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop