Charles Gervasa,
Malik Dilshad Khana,
Chunyang Zhangb,
Chen Zhaob,
Ram K. Guptab,
Emanuela Carleschic,
Bryan P. Doylec and
Neerish Revaprasadu*a
aDepartment of Chemistry, University of Zululand, Private Bag X1001, Kwa-Dlangezwa 3880, South Africa. E-mail: RevaprasaduN@unizulu.ac.za
bDepartment of Chemistry, Pittsburg State University, Pittsburg, Kansas, USA
cDepartment of Physics, University of Johannesburg, P. O. Box 524, Auckland Park, 2006, Johannesburg, South Africa
First published on 2nd July 2018
Thiospinels show interesting catalytic and energy storage applications, however, the cationic disorder can have major influence on the energy generation and/or energy storage applications. In this study, the effect of stoichiometric variation of metals in a thiospinel i.e. NixCo3−xS4, is examined on energy generation and storage properties. Nickel- or cobalt-rich NixCo3−xS4 nanosheets were prepared by the hot injection method using single molecular precursors. The nanosheets were characterized by p-XRD, TEM, HR-TEM, EDX and XPS techniques. Nickel-rich and cobalt-rich nanosheets were tested for oxygen and hydrogen evolution reactions and for supercapacitance performance. It was observed that the nickel-rich NixCo3−xS4 nanosheets have superior energy storage and energy generation properties.
Recently oxide and sulfide based ternary compounds of cobalt and nickel have attracted the attention of researchers due to their high efficiency in energy generation and energy storage devices. Ternary metallic oxides, such as NiCo2O4, have been considered appropriate for the fabrication of supercapacitors, being a better-conducting material than the binary oxides of the parent metals i.e. nickel and cobalt.7–9 However, NiCo2O4 shows less electron conductivity when compared to its sulfide counterpart, NiCo2S4. Furthermore, the Ni–Co–S system provides the material with improved performance and stability as compared to the oxide counterpart and their binary counterparts, Co–S and Ni–S.10–12
Thiospinels are sulfur based ternary compounds that exist in the spinel structure, whose general formula is AB2S4 (where A = M2+ and B = M3+). Such a structure is said to be built on a closely packed array of S−2 ions, with A2+ and B3+ metallic ions occupying the tetrahedral and octahedral sites respectively.13 This means the ternary Ni–Co(S) system such as NiCo2S4 with a normal spinel structure has the synergetic effect from both Ni2+ and Co3+ in the presence of S2−, hence making the material better for both electrochemical activities as well as electrocatalytic activities for H2 and O2 evolution reactions.
Ternary Ni–Co(S) system has superior quality electrocatalytic activities for oxygen and hydrogen evolution studies.14,15 The bimetallic catalysts of Fe, Ni, Co with the chalcogen atom (S, Se, Te) are potentially efficient candidates because of their natural abundance (hence low cost),16 and high electrocatalytic activity, especially when attached to smaller organic molecules.17 Bimetallic chalcogenides fit in better than binary sulfides due to synergetic effects of the transition metal atoms that often exhibit collective higher catalytic activity in different chemical environments.
Various protocols have been used to fabricate NiCo2S4. They include hydrothermal methods,16 electrochemical depositions,18 anion exchange methods,19 controlled sulfurization methods,20 and chemical bath deposition.21 The hydrothermal method is the predominant protocol used in the preparation of NiCo2S4 as it is considered to be cost-effective and environment friendly as compared to other protocols. Lou and co-workers synthesized hollow nanofibers, ball-in-ball hollow spheres, and onion-like NiCo2S4 particles via the ion-exchange method.22–24 The controlled growth of NiCo2S4 nanosheets on a graphene matrix was shown to lead to superior supercapacitive performance.9 Peng et al. reported the hydrothermal synthesis of a NiCo2S4-RGO hybrid and investigated its electrochemical performance.25
In ternary and/or quaternary semiconductors, stoichiometric variations play an important role and this composition tunability is the key property to their characteristics and device performance.26 Non-stoichiometric compounds are not really desirable as, in II–V or III–V compounds, it results in low-performing device characteristics. However, for ternary materials, this is not necessarily always true. It was observed for Cu/(Ag)InS2 nanomaterials that the In-rich compounds are better emitters as compared to Ag- or Cu-rich compositions.27 Similarly, a slight cationic disorder in AgBiS2 results in a significant change in electronic properties.28 Although, there are some examples of cationic disorders of I–III–V compounds, the effect of cationic disorder on energy storage and/or energy generation performance of thiospinels is not very well explored.
The use of single molecular precursors is often known to give superior results in nanomaterials synthesis as it makes use of preformed bonds with the decomposition product easily predicted.29–33 Furthermore, they are equally suitable for the synthesis of nanomaterials and deposition of thin films.34–38
Herein, we have designed a facile one-pot synthesis of NixCo3−xS4 thiospinel using dithiocarbamate complexes of nickel and cobalt as single molecular precursors by thermolyzing the stoichiometric amounts of the complexes in a hot primary amine. This protocol has enabled us to tune the morphology and stoichiometry of NixCo3−xS4. It was observed that the nanoparticles synthesized at 200 °C were cobalt-rich, whereas the higher temperature (250 °C) results in the nickel-rich material. The effect of stoichiometric disorder of metals, on the energy storage and energy generation properties, has been examined in detail.
The electrocatalytic behavior of the NixCo3−xS4 samples for oxygen evolution reaction (OER) and hydrogen evolution reaction (HER) was also examined using standard three electrode system consisting of NiCo2S4 on nickel foam as a working electrode, platinum wire as a counter electrode and saturated calomel electrode (SCE) as a reference in 1 M KOH electrolyte. Electrocatalytic testing includes linear sweep voltammetry (LSV), cyclic voltammetry (CV), chronoamperometry and electrochemical impedance spectroscopy. LSV was performed at a scan rate of 2 mV s−1 in both OER and HER region. The potential was converted to RHE using the Nernst equation. All the EIS measurements were recorded in a frequency range of 0.05 Hz to 10 kHz with an applied 10 mV of AC amplitude.
The thermal stability of the complexes was observed by thermogravimetric analysis (TGA) and is shown in Fig. S1 (ESI†). Complex (1) decomposes in two steps, where a major mass loss of about (72%) was observed at 324 °C, followed by a minor weight loss of about (7.3%) around 385 °C, suggesting the progressive loss of the sulfur atoms. Complex (2) shows single step decomposition with a weight loss of 79.16% at 289 °C. The residue obtained from the complex (1) (21.09%) and complex (2) (20.84%) was found to be very close to the calculated values for NiS (20.75%) and CoS (20.82%) respectively.
Both complexes show high thermal stability, however, NixCo3−xS4 nanoparticles were synthesized with ease in OLA, at a temperature quite lower than the decomposition temperature as determined by TGA. The use of oleylamine (OLA) has two major advantages; the long alkyl chain helps in the effective capping of the nanoparticles,41 and the primary amine can catalyze the breakdown of the molecular precursors, initiating decomposition at low temperature.42 Both nickel and cobalt sulfides can exist in spinel structures i.e. Co3S4 and Ni3S4. Similarly, a combination of both in different ratios can also be used for generation of thiospinel lattice, such as NiCo2S4, CoNi2S4 and (Ni, Co)3S4. Most of the work has been done on energy applications of NiCo2S4 but little on nickel incorporated Co3S4 (i.e. NixCo3−xS4).
The complexes [Ni(Etpzdtc)2] and [Co(Etpzdtc)3], were dispersed in OLA in stoichiometric amounts (1:1), and injected in OLA at 200 °C and/or 250 °C, under similar reaction conditions. The effect of temperature was examined on the morphology, crystallinity, and composition of the synthesized NixCo3−xS4 nanoparticles. The phase of the synthesized OLA capped NixCo3−xS4 nanoparticles, was determined by p-XRD measurements (ESI Fig. S2†). The diffraction peaks can be indexed to the (220), (311), (400), (422), (511), (440), (533), (444) and (553) which matches well with the standard pattern (ICDD # 01-073-1704). The diffraction peaks are respectively positioned at 2θ = 26.84°, 31.58°, 38.32°, 47.40°, 50.47°, 55.31°, 65.10°, 69.29°, 78.76°. There is no detectable impurity, secondary phase or unassigned peaks. The intensity of the peaks was comparatively weaker at a lower temperature (200 °C) and more pronounced at a higher temperature of 250 °C. Coordinating solvents can reduce the decomposition temperature of both precursors, however, crystallinity can only be enhanced by increasing the temperature. The absence of binary phases, such as NiO, CoO, NiS, and CoS confirms the phase purity of nanomaterials prepared from single molecular precursors. The p-XRD displays typical behavior of cobalt-rich minerals i.e. the rising trend of the diffraction pattern from the baseline. It is due to the fact that the energy of the Kα radiation of copper is higher than the k-absorption edge of cobalt, hence, cobalt fluoresce strongly and give rise to such behaviour. The calculation of the lattice constant using diffraction peak (311) resulted in 9.319 Å which is in agreement with previous reports.21
The microscopic analysis of the morphology of the nanomaterials, synthesized at both temperatures, is shown in Fig. 1. The synthesis of NixCo3−xS4 at a temperature of 200 °C, indicates the formation of nanosheets. The sheets were stacked and showed a poly-dispersed nature whereas, the size of the nanosheets, was in the order of microns. The crystalline nature of the nanosheets was indicated by clear lattice fringes and the presence of well-defined spots in the SAED pattern (Fig. 1(b)). Similar sheet-like morphology was observed for NixCo3−xS4 synthesized at a higher temperature of 250 °C (Fig. 1(c)). However, the rate of nucleation increases many folds with an increase of temperature, hence, growth defects such as twinning and some stacking faults were also observed by TEM analysis (Fig. 1(d)).
Fig. 1 TEM images for OLA capped NixCo3−xS4 synthesized at (a and b) 200 °C (c and d) 250 °C and (e) shows HAADF image along with elemental mapping of nanosheets synthesized at 200 °C. |
Fig. 1(e) shows high-angle annular dark-field (HAADF) image and the elemental mapping indicates a uniform distribution of all elements in the sample. The composition by ICP analysis indicates that the sample prepared at 200 °C was cobalt-rich and the exact composition of the sample was found to be Ni1.22Co1.50S4.31 (henceforth referred to as NiCoS-1), whereas, the sample synthesized at 250 °C was nickel-rich, and a stoichiometric composition of Ni1.63Co1.18S4.17 (henceforth referred to as NiCoS-2) was observed.
Fig. 2 (a) XPS survey spectra (b) S 2p core level spectra (c) Co 2p core level spectra and (d) Ni 2p core level spectra for both samples synthesized at 200 °C and 250 °C. |
Fig. 2(c) shows the Co 2p core level for the investigated samples. The Co 2p binding energy region is dominated by two sharp peaks located at ∼779 eV and ∼794 eV binding energy, and two broad peaks at ∼780.5 eV and ∼796.5 eV, which correspond to the Co 2p3/2 and Co 2p1/2 spin–orbit components for Co3+ and Co2+ oxidation states, respectively. The identification of Co2+ and Co3+ oxidation states is consistent with what is published in the literature for similar NiCo-based systems.43–45 Together with the main peaks, two weak shake-up satellites are detected at about 6.5 eV towards higher binding energy from the Co3+ peaks. The line shape of this core level has been fitted using three spin–orbit doublets, as shown in Fig. 2(c). Each doublet is separated by a spin–orbit splitting of approximately 15.1 eV. The overall fit to the experimental data (thick line) shows good agreement with the raw data. Table S1 (ESI†) displays the Co 2p3/2 binding energy of the three doublets, as well as the area ratio between Co2+/Co3+ oxidation states. As one can see, the area changes quite drastically from one sample to the other, reflecting an intrinsic difference in the electronic structure within the series.
Fig. 2(d) shows the comparison of the fitted Ni 2p XPS core level spectra. The Ni 2p3/2 and 2p1/2 spectral regions are composed of three spin–orbit doublets corresponding to (in order from lower to higher binding energy) Ni3+, Ni2+ oxidation states and a broad shake-up satellite of the Ni3+ species, respectively. As for the Co 2p core level above, the fitted components (together with a Shirley-type background) appear in Fig. 2(d) together with the raw data. Table S2† (ESI†) displays the Ni 2p3/2 binding energy of the three doublets, as well as the area ratio between Ni2+/Ni3+ oxidation states. Interestingly, the ratios for Ni oxidation states have an opposite trend with respect to those for Co. This was to be expected as charge neutrality of the samples has to be ensured.
In general, the line shapes of the Co 2p and Ni 2p core levels reported here and the observation that both ions are present in divalent and trivalent oxidation states are consistent with previous studies for NiCo2S4.18,45 The elements nickel, cobalt, and sulfur have the tendency to exist in two mineralogical forms i.e. NiCo2S4 or CoNi2S4. In NiCo2S4, Ni is in the +2 oxidation state, whereas Co is in the +3 oxidation state. However, a small amount of Ni can convert to +3 oxidation state and Co can occupy the site with +2 oxidation state. The ratio can be increased if, comparatively, one of the element is in slight excess, and can occupy the position of the deficient element as well to stabilize the spinel framework.
The XPS results from the S 2p core level are reported in Fig. 2(b). The line shape is consistent with previous reports on these systems.44,45 Because the spin–orbit splitting between the 2p3/2 and the 2p1/2 peaks in sulfur is only 1.18 eV (i.e. of the same order of magnitude as the experimental resolution for these measurements), we have decided to report the 2p3/2 and the 2p1/2 peaks as separate peaks in the figure. This is to aid in visualization of the fits to the spectra. The main spin–orbit doublet at lower binding energy is attributed to S in the 2-oxidation state. This is typically what happens when sulfur ions bond with metallic ions in ternary metal sulfides.46 The S 2p doublets whose 2p3/2 peaks are observed at 164 eV and 169 eV are attributed to surface sulfur with high oxide state, such as sulfites and sulfates, respectively. The binding energies of the 2p3/2 core level for these three doublets are reported in Table S3† (ESI†).
Fig. 3 (a) Polarization curves and (b) Tafel slopes for NixCo3−xS4 samples in OER range and (c) polarization curves and (d) Tafel slopes for NixCo3−xS4 samples in HER range. |
The HER activity of the NixCo3−xS4 samples was also studied, the polarization curves for NixCo3−xS4 samples in HER region is given in Fig. 3(c). As seen in Fig. 3(c), NiCoS-2 required an overpotential 163 mV, while NiCoS-1 required an overpotential of 179 mV to achieve a current density of 10 mA cm−2. As seen in the polarization curves, NiCoS-2 showed significant improvement in the HER performance compared to NiCoS-1. It can be observed from Fig. 3(d), the Tafel slope for NiCoS-1 and NiCoS-2 was calculated to be 176 and 161 mV dec−1, respectively. The lower Tafel slope for NiCoS-2 showed faster HER kinetics for electrocatalysts, which confirmed the better HER performance. Sivanantham et al. have used a two-step hydrothermal method to complete the in situ growth of hierarchical NiCo2S4 nanowire arrays on a nickel foam substrate, which needed an overpotential of 210 mV to deliver a hydrogen production current density of 10 mA cm−2.52 The NiCo2S4 electrocatalyst prepared by Liu et al. required an overpotential of 305 mV to generate a current density of 10 mA cm−2 with a Tafel slope of 141 mV dec−1.50 The NiCo2S4 with porous nanosheets array topology showed an overpotential of 181 mV to reach 10 mA cm−2.51 The other recently reported HER catalytic properties using Ni, Co, and other non-precious metals in the alkaline medium are given in Table S5 (ESI†).
To further understand the reason for improved electrocatalytic properties of NiCoS-2 sample, the electrochemical active surface area of NiCoS-1 and NiCoS-2 was determined using cyclic voltammograms in the non-faradic region. All the CV measurements were performed in a potential range of 0.87–0.92 V (vs. RHE), where no faradic reactions were observed. The CV curves were recorded for NiCoS-1 and NiCoS-2 at various scan rates to determine the electrochemical double layer capacitance (Cdl), which is directly proportional to the electrochemical active surface area of the sample. Cdl was calculated by plotting half of the difference in positive and negative current densities at 0.895 V (vs. RHE) versus the scan rates (Fig. 4(a)). The Cdl value of the NiCoS-1 and NiCoS-2 was calculated to be 2.27 and 7.48 mC cm−2, respectively. The effective electrochemical surface area of the NiCoS-2 is more than 3 times higher than that of the NiCoS-1. The highest active electrochemical surface area provides NiCoS-2 the better OER and HER activity than NiCoS-1.53
Fig. 4 (a) Current density vs. scan rate plots for NixCo3−xS4 samples, (b) |Z| vs. frequency plots and (c) Zreal vs. Zimg plots for NiCo2S4 samples at 0.5 V (vs. SCE). |
The electrocatalytic properties of NixCo3−xS4 samples were further analyzed using electrochemical impedance spectroscopy tests. From Fig. 4(b), which is the plots of total impedance |Z| versus frequency, it is observed that NiCoS-2 has the lower total impedance, having only about half of the impedance of NiCoS-1. Moreover, NiCoS-2 has smaller semicircle in the Nyquist plots (Fig. 4(c)). The smaller radius of the semicircle at low-frequency region suggests a lower charge transfer resistance and a faster electron transfer during electrochemical reactions. Therefore, the results of EIS tests show that the NiCoS-2 possesses an excellent electrocatalytic performance which is consistent with the excellent OER and HER performance.
The total impedance |Z| and Nyquist plots of NiCoS-2 at various potentials are shown in Fig. S3 (ESI†). It is observed that the total impedance |Z| decreases with increasing potential, indicating improved electrocatalytic performance at higher potential. Similarly, the Nyquist plot starts to convert to semicircle from a straight line with the increase in the potential (Fig. S4(b)†). That is because semicircle region depends on overpotential as the increase in potential provides faster reaction which leads to the reduction of semicircle's diameter.
The electrocatalytic long-term durability was measured using chronoamperometry. It was observed that the current density of NiCoS-2 was maintained almost at a constant level of 10 mA cm−2 over the 16 h duration (Fig. S4, ESI†). The little negative effect on current density with increasing time confirms that the NiCoS-2 is very stable and could be used as a durable electrocatalyst for water splitting applications.
Specific capacitance of NiCoS-1 and NiCoS-2 were calculated from the CV data and GCD data and was plotted in Fig. 6(a) and (b) as a function of scan rates and current densities, respectively. Higher specific capacitance was observed at a lower scan rate and lower current density for both NixCo3−xS4 samples because of lower scan rate and lower current density provides more time for redox reactions. The maximum specific capacitance of NiCoS-2 was 1273 F g−1 and 494 F g−1 which were calculated from the CV curve at a scan rate of 1 mV s−1 and GCD curve at 1 A g−1, respectively. The supercapacitor performances of NiCoS-1 and NiCoS-2 are superior or comparable to other reported NiCo-based samples. The NiCo2O4 coral-like porous crystals synthesized using sol–gel approach by Wu et al. have a capacitance of 217 F g−1.54 Salunkhe et al. prepared NiCo2O4 films on indium tin oxide substrates by a chemical bath deposition method. The NiCo2O4 films showed a highest specific capacitance of 490 F g−1.55
The long-term performance of the NixCo3−xS4 samples was studied using galvanostatic charge–discharge measurements for 5000 cycles (Fig. 6(c and d)). As seen in the capacitance versus number of cycles, both samples exhibited a very stable performance. The NiCoS-1 and NiCoS-2 showed a capacitance retention of 95% and 85% after 5000 cycles, respectively. About 60% of capacitance was maintained after 1000 cycles for the ultrathin porous hierarchically textured NiCo2O4–graphene oxide nanosheets.9 The capacitance of caterpillar-like NiCo2S4 nanocrystal arrays on nickel foam prepared by Chen et al. kept 83% after 3000 cycles.56
To further evaluate the performance of NiCoS-2 as energy storage material and the effect of temperature on the electrochemical properties, a symmetrical supercapacitor device was fabricated and tested in 3 M KOH using CV, GCD, and EIS. Fig. 7(a) shows the CV curves of the device at various scan rates. The potential window was as large as 1 V. From the CV curves of the device at various temperatures (Fig. 7(b)), the area under the CV curves was observed to increase with increasing temperature, showing improvement in storage capacity of the device. Similar behavior was observed in the GCD curves at different temperatures (Fig. 7(c)). Charge and discharge time increased with temperature confirming the improvement in charge storage capacity. As seen in Fig. 7(d), there was a 180% increase in specific capacitance when the temperature was raised from 10 to 60 °C. The total impedance |Z| with different frequency and Nyquist plots at various temperatures are given in (ESI Fig. S5(a) and (b)†). As seen, total impedance |Z|, Zreal and Zimag decreases with increasing temperature, which could be due to increased mobility of the ions in the electrolyte.47 Similar phenomena was observed in other reports.8,57 A 150% improvement in the specific capacitance of the device using flower-structure NiCo2O4 by increasing temperature from 10 to 60 °C.8
Footnote |
† Electronic supplementary information (ESI) available: TGA, p-XRD and EDX, electrochemical plots and Tables. See DOI: 10.1039/c8ra03522a |
This journal is © The Royal Society of Chemistry 2018 |