Next Article in Journal
Electroless Cobalt Deposition on Dealloyed Nanoporous Gold Substrate: A Versatile Technique to Control Morphological and Magnetic Properties
Next Article in Special Issue
MOF-Derived Ultrathin NiCo-S Nanosheet Hybrid Array Electrodes Prepared on Nickel Foam for High-Performance Supercapacitors
Previous Article in Journal
Fano-Like Resonance of Heat-Reconfigurable Silicon Grating Metasurface Tuned by Laser-Induced Graphene
Previous Article in Special Issue
Facile Synthesis of NixCo3−xS4 Microspheres for High-Performance Supercapacitors and Alkaline Aqueous Rechargeable NiCo-Zn Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Ni3S2 and MOF-Derived Ni(OH)2 Composite Electrode Materials on Ni Foam for High-Performance Supercapacitors

School of Materials Science and Engineering, Shanghai University of Engineering Science, Shanghai 201620, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(3), 493; https://doi.org/10.3390/nano13030493
Submission received: 31 December 2022 / Revised: 20 January 2023 / Accepted: 22 January 2023 / Published: 26 January 2023
(This article belongs to the Special Issue Novel Nanoporous Materials for Energy Storage and Conversion)

Abstract

:
Honeycomb-like Ni(OH)2/Ni3S2/Ni foam (NF) was fabricated via a two-step hydrothermal process and subsequent alkalization. Ni3S2 with a honeycombed structure was in-situ synthesized on the NF surface by a hydrothermal process. MOF-derived Ni(OH)2 nanosheets were then successfully grown on the Ni3S2/NF surface by a second hydrothermal process and alkaline treatment, and a large number of nanosheets were interconnected to form a typical honeycomb-like structure with a large specific surface area and porosity. As a binder-free electrode, the prepared honeycomb-like Ni(OH)2/Ni3S2/NF exhibited a high specific capacitance (2207 F·g−1 at 1 A·g−1, 1929.7 F·g−1 at 5 mV·s−1) and a remarkable rate capability and cycling stability, with 62.3% of the initial value (1 A·g−1) retained at 10 A·g−1 and 90.4% of the initial value (first circle at 50 mV·s−1) retained after 5000 cycles. A hybrid supercapacitor (HSC) was assembled with Ni(OH)2/Ni3S2/NF as the positive electrode and activated carbon (AC) as the negative electrode and exhibited an outstanding energy density of 24.5 Wh·kg−1 at the power density of 375 W·kg−1. These encouraging results render the electrode a potential candidate for energy storage.

1. Introduction

Owing to economic development and population growth, the world’s demand for energy is growing. However, traditional energy sources present some disadvantages, such as limited resources and environmental unfriendliness [1,2,3,4]. Therefore, ecofriendly energy resources and sustainable electrical energy storage devices (especially lithium-ion batteries and supercapacitors) must be urgently developed. Compared with batteries, supercapacitors have the advantages of high charge energy storage, fast charge and discharge, long-term stability as well as environmental friendliness [5,6,7], and have become the research focus in recent years. Depending on their charge storage mechanism, supercapacitors are divided into electric double layer capacitors (EDLCs) and pseudo-capacitors [8]. Compared to EDLCs, pseudo-capacitors have a higher specific capacity owing to their quick and reversible Faraday reactions [9]. The electrode material is the most critical factor affecting the performance of supercapacitors [2,6,10] and is mainly classified as carbon materials, conductive polymers, and transition metal compounds [11]. Transition metal compounds show superior specific capacity due to their abundant oxidation–reduction states and small internal resistance [12]. Currently, the widely studied transition metal compounds mainly include oxides (NiO and MnO2), sulfides (Ni3S2 and MoS2), and hydroxides (Ni(OH)2 and Co(OH)2). An S atom with lower electronegativity than an O atom is beneficial to electron transportation and could further promote the redox reactions, resulting in transition metal sulfides exhibiting higher specific capacities than oxides [13].
Among transition metal sulfides, Ni3S2 is a promising electrode material due to its abundant resources and high theoretical specific capacitance and electrical conductivity [11,14,15]. However, one drawback of Ni3S2 is its poor cycling stability due to the fall-off of the active materials caused by drastic volume changes during charge–discharge [16]. Huo et al. [17] designed 3D Ni3S2 nanosheet arrays grown on Ni foam (NF) by a hydrothermal process. The electrode showed an excellent specific capacitance of 1370 F·g−1 at 2 A·g−1 and a remarkable rate capability (952.0 F·g−1 at 20 A·g−1); however, its specific capacitance only remained at 91.4% after 1000 cycles. Chen et al. [18] prepared Ni3S2 with a mixed structure of flakes and granules using the hydrothermal method. The product showed a specific capacitance of 736.64 F·g−1 at 0.8 A·g−1 and maintained 82% of its original specific capacitance after 1000 cycles. An effective solution is to combine Ni3S2 with other materials to form a composite. In the absence of a direct contact between Ni3S2 and electrolyte, the internal electrode material can be effectively prevented from falling off under the continuous influence of high current, and the cycling stability is consequently improved. At present, the electrode materials for Ni3S2 composite mainly include carbon materials [19], polymer compounds [20], MXenes [21], sulfides [22], oxides [23], and hydroxides [24,25,26]. Among these, hydroxides have attracted attention, especially for Ni(OH)2 with a high theoretical specific capacity and abundant resources [27]. Composite electrodes containing Ni(OH)2 and Ni3S2 can effectively relieve the detachment of the internal material (Ni3S2) from the current collector, resulting in an improvement in the cycling stability. They can also provide additional active sites from the external material (Ni(OH)2), leading to a high charge storage. Wang et al. [26] fabricated a NF-Ni3S2@Ni(OH)2-graphene sandwich structure by a two-step hydrothermal method. Ni(OH)2 was in the middle of the structure, with rGO and Ni3S2 attached above and below, respectively. Owing to the structural feature, the electrode exhibited a high specific capacitance of 2258 F·g−1 at 1 A·g−1 and excellent cycling stability with no capacitance reduction after 3000 cycles. Pan et al. [24] additionally fabricated Ni3S2 nanorods in situ on a NF and then loaded them with the ultrathin Ni(OH)2 nanosheets by a two-step hydrothermal process. The electrode exhibited remarkable specific capacitance of 826 C·g−1 at 1 A·g−1, excellent rate capability (50% retention at 20 A·g−1), and outstanding cycling stability (82% of its initial capacitance retained over 10,000 cycles). For further improvement in cycling stability and specific capacity, an effective method is to construct a Ni(OH)2/Ni3S2 composite electrode with multiple channels and a high specific surface area by changing the structure of Ni(OH)2. The abundant pores can provide a large number of diffusion channels, greatly shortening the diffusion path of ions and reducing the stability decline caused by volume strain [28]. Meanwhile, the high specific surface area can provide additional active sites and increase the charge storage of electrode materials [29]. The metal–organic framework (MOF), which has a high porosity and surface area, is often used as a precursor and template to derive Ni(OH)2 [29,30,31]. However, a suitable strategy to compound Ni(OH)2 derived from Ni-MOF with Ni3S2 has not been discovered.
In this work, honeycomb-like Ni(OH)2/Ni3S2/NF (Ni(OH)2 derived from Ni-MOF) grown on a NF was synthesized by a two-step hydrothermal method coupled with an alkalization treatment. Highly conductive Ni3S2 as a carrier was fabricated in situ on NF by the first hydrothermal method without using any organic binder, and Ni(OH)2 nanosheets derived from Ni-MOF were then incorporated as a support by the second hydrothermal method and alkalization treatment in 6 M KOH for 6 h at 75 °C. The change in the morphology structure from Ni3S2/NF to Ni(OH)2/Ni3S2/NF was investigated, and the relationship between this change and the electrochemical performance was analyzed. When tested in three- and two-electrode systems, the as-prepared honeycomb-like Ni(OH)2/Ni3S2/NF electrode demonstrated excellent electrochemical performance.

2. Experimental methods

2.1. Materials

NF (thickness of 1 mm, Guangjiayuan New Material Co., Kunshan, China), thiourea (≥99%, Shanghai Titan Scientific Co., Ltd., Shanghai, China), nickel nitrate hexahydrate (Ni(NO3)2·6H2O, 98.5%, Shanghai Titan Scientific Co., Ltd., Shanghai, China), 1,3,5-benzenetricarboxylic acid (PTA, 99%, Aladdin Industrial Corporation, Shanghai, China), and N,N-dimethylformamide (DMF, ≥99.9%, Aladdin Industrial Corporation) were used without further operation.

2.2. Synthesis of Ni3S2 and MOF-Derived Ni(OH)2

Prior to the experiments, the NF (1.5 × 2 cm) was successively cleaned with acetone, 1 M HCl aqueous solution, deionized water (DI), and absolute ethanol in an ultrasonic cleaner to remove the organic matter and oxide layer and then dried in a vacuum drying oven at 60 °C for 12 h. The cleaned grey NF was placed in a 50 mL Teflon-lined stainless autoclave containing a homogeneous solution of 35 mL of DI and 10 mmol thiourea. The Teflon stainless autoclave was kept at 120 °C for 5 h to synthesize Ni3S2, which was subsequently cooled to room temperature. The samples were taken out and gently cleaned in DI and absolute ethanol. In brief, 2 mmol Ni(NO3)2·6H2O and 0.5 mmol PTA were dissolved in 30 mL of DMF and stirred for 60 min. The NF with in-situ synthesized Ni3S2 and the above solution were transferred into a 50 mL Teflon stainless autoclave to heat at 120 °C for 9 h, after which Ni-MOF was deposited onto Ni3S2. The resultant sample was gently cleaned several times by DMF and DI and then immersed in KOH solution at 75 °C for 6 h to transform Ni-MOF into MOF-derived Ni(OH)2. The sample was washed in DI before vacuum drying 60 °C for 12 h to obtain NF coated with green materials. The above synthesis process is clearly and intuitively illustrated in Figure 1.

2.3. Material Characterization

X-ray diffraction patterns were obtained to analyze the phase constituents of the resultant products on an X-ray diffractometer (XRD, Bruker, Karlsruhe, Germany) with Cu-Kα radiation (γ = 0.1540560 nm). Their morphologies were observed by a field-emission scanning electron microscope (SEM, S-4800, Hitachi, Tokyo, Japan) and a transmission electron microscope (TEM, JEM-2100F, JEOL, Tokyo, Japan) equipped with an energy dispersive spectrometer (EDS, X-MAX 65T, OXFORD, UK). Their chemical compositions and chemical valences were analyzed with an X-ray photoelectron spectroscope (XPS, Thermo Fisher Scientific, Waltham, MA, USA).

2.4. Electrochemical Tests

2.4.1. Three-Electrode System

The electrochemical performance was tested by cyclic voltammetry (CV), galvanostatic charge-discharge tests (GCD), and electrochemical impedance spectroscopy (EIS) in a three-electrode system. CV and EIS were tested on a CS350H electrochemical workstation (CORRTEST, Wuhan, China), and the GCD was tested on a CHI 760E electrochemical workstation (CH Instrument Inc, Shanghai, China). The NF sheet loaded with Ni3S2/MOF-derived Ni(OH)2 and Ni3S2 and the naked NF sheet free of active substances were used as the working electrode, and a saturated calomel electrode and a graphite sheet acted as the reference and counter electrodes, respectively. A 6 M KOH aqueous solution was chosen as the electrolyte. Prior to the tests, the prepared working electrode was soaked in 6 M KOH solution for 12 h to ensure full contact between the active substance and the electrolyte. CV measurements were performed within a potential window of −0.1 V to 0.6 V at the scan rate ranging from 5 mV·s−1 to 50 mV·s−1. GCD tests were carried out in the potential window from 0 to 0.4 V under different current densities (1, 2, 4, 6, 8, and 10 A·g−1). CV tests to measure the cycling stability of the electrode (the NF sheet loading Ni3S2/MOF-derived Ni(OH)2) were conducted for 5000 cycles at 50 mV·s−1. An EIS test was carried out in the frequency range of 105 Hz to 0.01 Hz.
The specific capacity (Cα and Cm, F·g−1) of electrode materials was calculated from the CV and GCD curves using Formulas (1) and (2), respectively [28,29]:
For the CV tests:
C α = I d V m × v × Δ V
For the GCD tests:
C m = I   × Δ t m × Δ V
where I (A) is to the current, m (g) is the mass of the working electrodes, v (mV·s−1) is the scan rate, ∆V (V) is the potential range, and ∆t (s) is the discharge time.

2.4.2. Two-Electrode System

A hybrid supercapacitor (HSC) was assembled by using an Ni3S2/MOF-derived Ni(OH)2 electrode as the positive electrode, activated carbon as the negative electrode, filter paper as the separator, and 6 M KOH as the electrolyte. For the preparation of the negative electrode, 90 wt.% activated carbon (AC) and 10 wt.% poly vinylidene fluoride were mixed in an appropriate amount of N-methyl-2-pyrrolidone solution and then coated on a NF. The coated NF was then folded, kept at a pressure of 10 MPa for 10s, and dried at 120 °C for 6 h. The mass ratio of the NF sheet loaded with Ni3S2/MOF-derived Ni(OH)2 as the positive electrode and AC as the negative electrode was calculated by the following charge balance formula (3) [32]:
m m + = C α + × Δ V + C α × Δ V
where m (g) is the active material mass, Cα (F·g−1) is the specific capacity under the CV tests, and ∆V (V) is the potential window for positive electrodes (+) and negative electrodes (−). On the basis of the charge balance formula, the mass ratio of positive electrodes and negative electrodes was calculated to be approximately 0.2.
In the two-electrode system, the electrochemical workstation used to measure CV and GCD was the same as that in the three-electrode system. The specific capacity (Cdevice, F·g−1) of the HSC device was obtained by the following formula (4) [33]:
C d e v i c e = I × Δ t m t o t a l × Δ V
where I (A) is the constant discharge current, mtotal (g) is the sum of the masses of the positive and negative active materials, ∆V (V) is the potential window, and ∆t (s) is the discharge time.
The energy density (E, Wh·kg−1) and power density (P, W·kg−1) of the HSC device were computed according to Formulas (5) and (6), respectively [34]:
E = C d e v i c e × Δ V 2 2 × 3.6
p = E × 3600 Δ t
where Cdevice (F·g−1) is the specific capacitance of the HSC device, and ∆V (V) and ∆t (s) are the voltage window and discharge time, respectively.

3. Results and Discussion

3.1. Structural Characterization

The as-prepared electrode was characterized by XRD to investigate its phase composition (Figure 2). Strong diffraction peaks marked with “♧” were the characteristic peaks of the substrate (NF) (JCPDS, No.01-070-0989) found at 44.6°, 51.9°, and 76.6°, corresponding to the (111), (200), and (220) planes, respectively. The diffraction peaks (marked with “♢”) at 21.8°, 31.2°, 37.9°, 44.5°, 49.8°, 50.3° 54.7°, 55.3°, and 72.6° were assigned to the (100), (1−10), (111), (200), (210), (2−10), (211), (2−11), and (310) planes of Ni3S2, respectively (JCPDS, No:01-073-0698). The sulfidization of NF can be described as follows [11]:
( N H 2 ) 2 C S + 2 H 2 O = H 2 S + C O 2 + 2 N H 3
2 H 2 S + 3 N i = N i 3 S 2 + 2 H 2
In addition to the diffraction peaks of NF and Ni3S2, other diffraction peaks (marked with “♤”) were detected at 19.3°, 33.1°, 38.5°, 52.1°, 59.1°, 62.7°, and 72.6°, corresponding to the (001), (100), (101), (102), (110), (111), and (201) planes of Ni(OH)2, respectively (JCPDS, No:00-014-0117). Yang et al. [35] failed to observe the characteristic peaks of Ni-MOF (at 15.7°, 18.4°, and 23.7°), indicating that all Ni-MOFs were completely converted to Ni(OH)2. The diffraction peaks related to Ni3S2 and Ni(OH)2 were comparatively weak due to their lower content compared with that of the NF substrate. Overall, Ni3S2 and Ni(OH)2 were successfully synthesized on NF.
The XPS spectra of the Ni(OH)2/Ni3S2/NF electrode were tested to determine the element compositions and their valences. Ni, O, and S were found to be involved in the electrode (Figure 3a). Six peaks can be well fitted in the Ni 2p high-resolution spectrum (Figure 3b). Two peaks with the binding energies of 855.7 and 873.4 eV were related to Ni2+, two peaks with the binding energies of 857.6 and 875.2 eV were correlated with Ni3+ [36], and two peaks at 861.8 and 879.7 eV belonged to the satellite peaks [1]. These results confirmed the successful synthesis of Ni3S2. For the O 1s spectrum, the peak at 530.9 eV implied the existence of OH¯ in Ni(OH)2 derived from Ni-MOF [29]. The peak at 532.8 eV may be attributed to the oxygen atoms adsorbed on the surface of the sample (Figure 3c) [24,37]. As shown in Figure 3d, the S 2p spectrum was composed of two peaks of S 2p3/2 and S 2p1/2 at 162.4 and 163.6 eV, respectively [38]. The other peak at 168.0 eV was the satellite peak [14]. All these findings were consistent with the XRD analysis, further proving the existence of Ni(OH)2 and Ni3S2.

3.2. Morphology Characterization

The morphology of pure NF, Ni3S2/NF, and Ni(OH)2/Ni3S2/NF were characterized using a field-emission scanning electron microscope. A large number of equiaxed grains with a size of approximate were observed on the skeleton of NF (Figure 4a,b). The low-magnification images of the Ni3S2/NF electrode showed that the NF skeleton was uniformly covered with a layer of dense substances (Figure 4c,d). Therefore, Ni3S2 with a honeycombed structure was in-situ synthesized on the NF surface. Ni3S2 with a similar structure was also reported by Wei et al. [39]. When the electrode was subjected to the second hydrothermal process and subsequent alkaline treatment, its surface became rough due to the abundant nanosheets protruding from the Ni3S2/NF surface (Figure 4e). Some even clustered together to form a flower-like structure, which connected to the individual nanosheets to create a typical honeycombed structure (Figure 4f). This finding indicated that the Ni(OH)2 derived from Ni-MOF recombined on the Ni3S2/NF surface. The honeycomb structure of Ni(OH)2/Ni3S2 was formed through two processes. First, Ni-MOF recombined on the surface of Ni3S2/NF by ”selective attachment“ and ”self-assembly“ [40]. Ni2+ and carboxyl organic ligands selectively attached to the high energy sites (protrusion) of Ni3S2 nanosheets and then self-assembled to form Ni-MOF on Ni3S2. Second, the organic ligands of Ni-MOF were replaced by OH and then converted into Ni(OH)2. The evolution mechanism of Ni-MOF converted to Ni(OH)2 is related to a “crystal–crystal conversion” mechanism [29]. When Ni-MOF was immersed in an alkaline solution, Ni(OH)2 with a hexagonal structure was formed due to the carboxyl organic ligands being replaced by OH. The Ni(OH)2 molecules lined up along the crystal plane parallel to the Z-axis and grew along the X–Y crystal plane, eventually forming the sheets. A large number of ultrathin nanosheets joined to one another to form a honeycomb structure. Owing to the high porosity of Ni-MOF retained in the Ni(OH)2 derived from Ni-MOF and the large specific surface area involved in the honeycombed structure, the diffusion path of OH- in the electrolyte was greatly shortened and additional active sites were provided. Hence, the electrochemical performance of the electrode material was improved.
Figure 5 displays the microstructure characteristics of the as-prepared samples. The sheet-like morphology of Ni3S2 can be vaguely distinguished due to discontinuous distribution and unclear boundaries (Figure 5a). The lattice fringe space was confirmed as 0.41 nm in the HRTEM image (Figure 5b), corresponding to the (100) plane of Ni3S2. After the second hydrothermal process and alkaline treatment, the sheet-like structure became evident, indicating that the Ni(OH)2 nanosheets derived from Ni-MOF had tightly attached to the surfaces of Ni3S2/NF nanosheets (Figure 5c). The dark part was the Ni3S2 nanosheets, around which a bright outer layer of Ni(OH)2 nanosheets can be detected (Figure 5d). The lattice stripes with d-values of 0.41 and 0.23 nm appeared in the inner and outer layers, respectively, which were related to the (100) plane of Ni3S2 and (101) plane of Ni(OH)2, respectively.
In Figure 6, the EDS element mapping analysis of Ni(OH)2/Ni3S2/NF revealed the co-existence of Ni, S, and O elements. This finding was consistent with the XPS results. The elements were uniformly distributed on the nanosheets. The content of O and Ni elements was higher than that of the S element, possibly because Ni3S2 was coated by Ni(OH)2. This result further proved the successful preparation of Ni3S2/Ni(OH)2/NF.

3.3. Electrochemical Tests

The electrochemical performance of all electrodes was investigated in a 6 M KOH aqueous solution with a standard three-electrode system. Figure 7a shows the CV curves of NF, Ni3S2/NF, and Ni(OH)2/Ni3S2/NF at the scanning rate of 10 mV·s−1 in a potential of −0.1 V to 0.6 V. For the NF electrode, the curve during the positive sweeping almost coincided with that during the negative sweeping and no redox peaks were observed, implying that NF as an inert substance hardly has the ability to store electrical charges. When Ni3S2 was in-situ synthesized on the NF, the CV curve seriously deviated from the rectangle of EDLCs and showed distinct redox peaks at 0.29 and 0.10 V, demonstrating the typical pseudo-capacitance behavior of the electrode [23,39]. The Faradic redox reaction can be described as follows [7,41]:
N i 3 S 2 + 3   O H     N i 3 S 2 ( O H ) 2 + 3   e
After Ni3S2 was covered with the Ni(OH)2 nanosheets, the area surrounding the CV curve became further enlarged. Clear inspection revealed that the oxidation and reduction peaks widened. Other redox peaks were discerned at 0.32 and 0.08 V, corresponding to the following reactions [7,41]:
N i ( O H ) 2 + O H     N i O O H + H 2 O + e
The area surrounding the CV curves represents the number of charges involved in the Faradic redox reaction [42]. Ni(OH)2/Ni3S2/NF had the largest surrounding area, followed by Ni3S2/NF. This finding indicated that the integration of Ni(OH)2 into Ni3S2 contributes to improving the charge storage capacity. As calculated by Formula (1), the specific capacitances of NF, Ni3S2/NF, and Ni(OH)2/Ni3S2/NF were 3.3, 623.8, and 1929.7 F·g−1, respectively. The Ni(OH)2 derived from Ni-MOF with a specific surface area showed an increase in specific capacitance by approximately 209%. The high charge storage of Ni(OH)2/Ni3S2/NF is closely related to the high specific capacitance (2082 F·g−1) of Ni(OH)2 and its unique morphology. On the one hand, XRD and XPS results showed that the Ni-MOF-derived Ni(OH)2 was successfully synthesized after the second hydrothermal reaction and subsequent alkalization treatment. The attached Ni(OH)2 with a layered structure is conducive to ion transport and therefore has a high specific capacity. As shown in the SEM image, the Ni(OH)2/Ni3S2/NF surface became rough after recombination, forming a honeycomb structure with a large contact area with electrolyte and thus providing additional active sites for the Faradic redox reaction. On the other hand, the Ni(OH)2 derived from Ni-MOF with a high porosity allowed OH ions to readily penetrate the interior of the electrode material and interact with Ni3S2/NF. The active substance of Ni3S2 was developed directly onto the NF substrate, eliminating the interface resistance between the two and enhancing the electron transport rate. These factors led to a synergistic effect, which can remarkably increase the specific capacitance of electrode materials.
Figure 7b displays the CV curves of Ni(OH)2/Ni3S2/NF samples at the scanning rates ranging from 5 mV·s−1 to 50 mV·s−1. All CV curves showed good symmetry, indicating that the samples have excellent electrochemical reaction reversibility. The specific capacitance of Ni(OH)2/Ni3S2/NF at different scanning rates was 1977.1 (5 mV·s−1), 1929.7 (10 mV·s−1), 1858.0 (15 mV·s−1), 1582.3 (25 mV·s−1), and 1063.5 F·g−1 (50 mV·s−1). When the scanning rate was increased from 5 mV·s−1 to 50 mV·s−1, the specific capacitance still remained at 53.8% of that obtained at 5 mV·s−1, demonstrating the remarkable rate capability of Ni(OH)2/Ni3S2/NF. With the gradual increase in the scanning rate, the specific capacitance showed a downward trend (Figure 7c) because the reaction was controlled by diffusion. At a high scanning rate, OH ions cannot diffuse to the electrode surface in time, resulting in a decrease in specific capacity. Moreover, the position of oxidation/reduction peaks of the CV curve is also affected by the diffusion rate of OH ions, manifesting as oxidation and reduction peaks moving in positive and negative directions, respectively. When the scanning rate is low, a large number of OH ions on the electrode surface participate in the Faradic redox reactions, that is, a quasi-equilibrium state. With the gradual increase in the scanning rate, the OH ions on the electrode surface drop sharply and those in the electrolyte cannot rapidly diffuse to the electrode surface, resulting in a disturbed equilibrium state. In conclusion, the position of the oxidation/reduction peaks and the surrounding area of the CV curve are affected by the diffusion rate of OH ions, thus confirming the pseudo-capacitance characteristics of the electrode material.
GCD tests were performed with a potential window ranging from 0 to 0.4 V at the current density of 2 A·g−1 to further explore the electrochemical performance of the electrode material (Figure 8a). The GCD curve of NF was approximately a straight vertical line, indicating that it virtually has no ability to store charge. This finding was consistent with the CV curve. In addition, the GCD curves of Ni3S2/NF and Ni(OH)2/Ni3S2/NF electrodes showed charge–discharge platforms related to redox reactions, and the discharge time of the latter (353.2 s) was longer than that of the former (93.2 s). This result indicated that the Ni(OH)2/Ni3S2/NF derived from Ni-MOF shows the highest capacitance. As calculated in Formula (2), the specific capacitances of NF, Ni3S2/NF, and Ni(OH)2/Ni3S2/NF were 2.6, 466.0, and 1766.0 F·g−1 at 2 A·g−1, respectively. The specific capacitance of Ni(OH)2/Ni3S2/NF was greatly improved at about 3.8 times that of Ni3S2/NF. GCD curves were obtained at different current densities ranging from 1 A·g−1 to 10 A·g−1 to further determine the specific capacitance of Ni(OH)2/Ni3S2/NF (Figure 8b). When the current density gradually increased, the Faradic redox reaction between the electrode and electrolyte interface decreased and the impedance value increased, resulting in an increase in voltage drop. From 1 A·g−1 to 10 A·g−1, the IR was calculated to be 0.002, 0.013, 0.029, 0.039 and 0.051 V. The relation of the specific capacitance of Ni(OH)2/Ni3S2/NF with different current densities is shown in Figure 8c. When the current density was 1, 2, 4, 6, 8, and 10 A·g−1, the specific capacitance was 2207, 1766, 1620, 1515, 1440, and 1375 F·g−1, respectively. When the current density was increased to 10 A·g−1, the specific capacitance remained at a relatively large value (62.3% of initial value), indicating the remarkable rate capability of Ni(OH)2/Ni3S2/NF. The specific capacitance decreased correspondingly with the increase in current density due to the insufficient utilization of active material caused by the rapid Faradic redox reactions at a high current density. According to the curves, the maximum specific capacitance of Ni(OH)2/Ni3S2/NF can reach 2207 F·g−1, which is significantly higher than those in the literature (Table 1). In addition, the GCD curves can reflect the rate capacity of electrode materials.
Figure 9 shows the capacitance retention and Coulombic efficiency of Ni(OH)2/Ni3S2/NF after 5000 cycles at a scanning rate of 50 mV·s−1. The Coulombic efficiency of Ni(OH)2/Ni3S2/NF was 81.4%. After 500 cycles of CV tests, the electrode was able to maintain 90.4% of its initial specific capacitance and showed virtually no change afterward. The reduction in the specific capacitance may be due to the active materials on the electrode surface slightly falling off from the current collectors during the early cycles, a common phenomenon caused by volume expansion [32]. After 500 cycles, the extremely high capacitance retention rate may be attributed to the unique morphology of the electrode. During this process, OH ions can completely penetrate the interior (Ni3S2) of the electrode material and undergo redox reactions. However, under the protection of the external material Ni(OH)2, the volume change of Ni3S2 will be limited, thus greatly reducing the shedding of Ni3S2. Meanwhile, the porous honeycomb structure can alleviate the volume stress of the material to a certain extent, thus leading to excellent cycling stability [28]. The above factors lead to the synergistic effect and promote the cycling stability of Ni(OH)2/Ni3S2/NF. Compared with those in the literature [46,47,48,49,50,51], the sample showed better cycling capability (Table 2).
Impedance affects the ability of an electrode material to store charge to some extent by hindering the transport of electrons/ions. EIS was tested in the frequency range of 105 to 0.01 Hz to investigate the impedance of Ni3S2/NF and Ni(OH)2/Ni3S2/NF (Figure 10) (inset displays the fitting circuit). The components of the fitting circuit are closely related to the Nyquist plots, which are composed of a small semicircle in the high frequency region and a straight line in the low frequency. Rs is the equivalent series resistance (ESR), and its value corresponds to the intercept value on the real axis (Z’), reflecting the resistance of the electrolyte, electrode material inherent resistance, and active material/current collector interface contact resistance [20,26]. The smaller the Rs, the higher the conductivity and the more easily the electrons are transported. The Rs of Ni3S2/NF and Ni(OH)2/Ni3S2/NF were about 0.48 and 0.46 Ω, respectively, which may attributed to the binder-free structure without the obstruction of electron transportation. Rct, which stands for charge transfer resistance, is related to the diameter of the semicircle [52]. Its value reflects the difficulty level of the Faradic redox reactions. The lower the Rct, the easier the charge transfer and thus the Faradic redox reactions. Compared with that of Ni3S2/NF (0.56 Ω), the Rct of Ni(OH)2/Ni3S2/NF was significantly reduced (0.22 Ω). This result indicated that charges are easily transferred in the electrode material and electrolyte and further promote Faradic redox reactions. This phenomenon is due to the large specific surface area and high porosity of Ni-MOF-derived Ni(OH)2, which greatly reduces the transport path of active ions. Warburg impedance (W0), which reflects the diffusion rate of active ion inside the electrode materials, is related to the slope of a straight line in the low frequency region [11]. The higher the slope, the lower the diffusion resistance and the better the electrochemical performance of the electrode material. The slope of Ni(OH)2/Ni3S2/NF was greater than that of Ni3S2/NF, suggesting that the W0 of Ni(OH)2/Ni3S2/NF (0.32 Ω) is smaller than Ni3S2/NF (0.65 Ω) due to the synergistic effect of a large contact area, a high specific area, and a high porosity. Meanwhile, a constant phase element was used to form a fitting circuit. The results demonstrated the advantage of low impedance of the Ni(OH)2/Ni3S2/NF electrode for an excellent electrochemical performance.

3.4. Electrochemical Measurements of the Ni(OH)2/Ni3S2/NF//AC HSC Device

A HSC device was assembled to further evaluate the practical application of the Ni(OH)2/Ni3S2/NF electrode (Figure 11). An Ni(OH)2/Ni3S2/NF electrode was used as the positive electrode, AC as the negative electrode, filter paper as the separator, and 6 M KOH as the electrolyte. As a key factor for evaluation, the mass ratio of Ni(OH)2/Ni3S2/NF and AC (apart from the NF current collector) was calculated using Formulas (1) and (3). Figure 12a shows the CV curves of Ni(OH)2/Ni3S2/NF and AC with voltage ranges of −0.1 V to 0.6 V and −1 V to 0 V, respectively, in a three-electrode system at a scanning rate of 50 mV·s−1. From the CV curves, the specific capacitances of Ni(OH)2/Ni3S2/NF and AC were calculated to be 1063.5 and 143.1 F·g−1, respectively (Formula 1). Therefore, the mass ratio of Ni(OH)2/Ni3S2/NF and AC was calculated to be approximately 0.2 by substituting the above values of the specific capacitance into Equation (3). Finally, the calculated masses of Ni(OH)2/Ni3S2/NF and AC were 4.6 and 23.0 mg, respectively, and both electrodes were assembled into the HSC device. The mass of the entire HSC device was the sum of the two active materials.
Figure 12b displays the CV curves of the device with the scanning rate ranging from 5 mV·s−1 to 100 mV·s−1 at a voltage window of 0–1.5 V. The shapes of all CV curves remained unchanged, indicating that 1.5 V is a suitable working potential range for the HSC device. GCD curves (Figure 12c) were also tested to identity the electrochemical performance of the HSC device. The curves almost maintained their symmetry at different current densities, demonstrating that the device has good electrochemical reversibility. As calculated in Equation (4), the specific capacitances of the device were 78.4, 50.3, 39.5, 29.9, 23.2, 22.4, and 18.7 F·g−1 at the current density of 0.5, 1, 2, 4, 6, 8, and 10 A·g−1, respectively. From 0.5 A·g−1 to 10 A·g−1, the decrease in specific capacity can be attributed to an incomplete reaction at a large current density. Ragone plots with the energy and power density of the HSC device were obtained at different current densities using Equations (5) and (6) (Figure 12d). The device showed a maximum energy density of 24.5 Wh·kg−1, accompanied by a power density of 375.0 W·kg−1. Its energy density gradually reduced to 4.2 Wh·kg−1 as the current density increased to 10 A·g−1, accompanied by an increase in power density to 7.6 KW·kg−1. This value is higher than previous results [53,54,55,56,57].

4. Conclusions

(1) A honeycomb-like Ni(OH)2/Ni3S2/NF grown on an NF was prepared successfully via the two-step hydrothermal approach, followed by an alkaline treatment.
(2) The MOF-derived Ni(OH)2 nanosheets were grown on the Ni3S2/NF surface, resulting in the formation of a honeycomb-like Ni(OH)2/Ni3S2/NF that has a large specific area and porosity and can increase the active sites and shorten the ion transport pathway.
(3) The as-prepared Ni(OH)2/Ni3S2/NF electrode showed an outstanding specific capacitance (2207 F·g−1 at 1 A·g−1, 1929.7 F·g−1 at 5 mV·s−1) and retained 62.3% of the initial value at 10 A·g−1. Moreover, 90.4% of its initial specific capacitance was retained after 5000 cycles at 50 mV·s−1 in a three-electrode system.
(4) A HSC was assembled and tested in a two-electrode system with an Ni(OH)2/Ni3S2/NF electrode as the positive electrode, AC as the negative electrode, filter paper as the separator, and 6 M KOH as the electrolyte. The device displayed a superior electrochemical performance with a high energy density (24.5 Wh·kg−1 at a power density of 375 W·kg−1).
All the results demonstrated that honeycomb-like Ni(OH)2/Ni3S2/NF has great potential for application in high-performance supercapacitors. However, given that the electrode material determines the type of electrolyte, the working voltage of the aqueous electrolyte (KOH) will be limited by the voltage decomposition of water. According to Equations (5) and (6), despite the high specific capacity of this electrode material, its energy density is extremely low and greatly restricts its commercialization. Appropriate composites of carbon and MOF-derived Ni(OH)2 can be compounded on the surface of Ni3S2/NF to improve the electrochemical performance of Ni(OH)2/Ni3S2/NF. In addition, an appropriate amount of active material must be attached to Ni3S2/NF by controlling the time or concentration of the hydrothermal process. If the attached active materials are excessive, then Ni3S2/NF cannot fully react with the ions in the electrolyte and the impedance will increase. As a result, the electrode material will have a low specific capacity and a poor rate capability and cycling stability. On the contrary, if the attached active materials are not enough to completely cover the underlying materials, then the active sites will be significantly reduced and the electrochemical performance will be poor. Carbon materials (graphene and carbon nanotube) with excellent electrical conductivity, a large specific surface area, and high mechanical flexibility can promote the ion/electron transport rate while further reducing the volume change caused by OH¯ intercalation/deintercalation in Ni(OH)2/Ni3S2/NF, thus providing the electrode material with a high specific capacity and an excellent rate capability and cycling stability.

Author Contributions

Formal analysis, M.S.; Investigation, M.S.; Resources, J.L. (Jun Li); Writing –original draft, M.S.; Supervision, R.L., Y.Y. and J.L. (Jing Li); Project administration, J.L. (Jun Li). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Shanghai, China (20ZR1422200) and Class III Peak Discipline of Shanghai—Materials Science and Engineering (High-Energy Beam Intelligent Processing and Green Manufacturing).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fu, S.; Ma, L.; Gan, M.; Shen, J.; Li, T.; Zhang, X.; Zhan, W.; Xie, F.; Yang, J. Sn-doped nickel sulfide (Ni3S2) derived from bimetallic MOF with ultra high capacitance. J. Alloys Compd. 2020, 859, 157798. [Google Scholar] [CrossRef]
  2. Jiang, S.; Li, S.; Xu, Y.; Liu, Z.; Weng, S.; Lin, M.; Xu, Y.; Jiao, Y.; Chen, J. An iron based organic framework coated with nickel hydroxide for energy storage, conversion and detection. J. Colloid Interface Sci. 2021, 600, 150–160. [Google Scholar] [CrossRef] [PubMed]
  3. Shahi, M.; Hekmat, F.; Shahrokhian, S. 3D flower-like nickel cobalt sulfide directly decorated grassy nickel sulfide and encapsulated iron in carbon sphere hosts as hybrid energy storage device. Appl. Surf. Sci. 2021, 558, 149869. [Google Scholar] [CrossRef]
  4. Yan, Z.; Guo, C.; Yang, F.; Zhang, C.; Mao, Y.; Cui, S.; Wei, Y.; Hou, L.; Xu, L. Cliff-like NiO/Ni3S2 Directly Grown on Ni Foam for Battery-type Electrode with High Area Capacity and Long Cycle Stability. Electrochim. Acta 2017, 251, 235–243. [Google Scholar] [CrossRef]
  5. Chang, Y.; Sui, Y.W.; Qi, Q.J.; Jiang, Y.L.; He, Z.Y.; Wei, X.F.; Meng, K.Q. Hierarchical Ni3S2 nanosheets coated on mesoporous NiCo2O4 nanoneedle arrays as high-performance electrode for supercapacitor. Mater. Lett. 2016, 176, 274–277. [Google Scholar] [CrossRef]
  6. Liu, Y.; Ma, Z.; Xin, N.; Ying, Y.; Shi, W. High-performance supercapacitor based on highly active P-doped one-dimension/two-dimension hierarchical NiCo2O4/NiMoO4 for efficient energy storage. J. Colloid Interface Sci. 2021, 601, 793–802. [Google Scholar] [CrossRef]
  7. Yang, P.; Feng, L.; Wang, S.; Shi, J.; Xing, H. Construction of core-shell Ni@Ni3S2@NiCo2O4 nanoflakes as advanced electrodes for high-performance hybrid supercapacitors. J. Phys. Chem. Solids 2021, 155, 110110. [Google Scholar] [CrossRef]
  8. Xu, Z.; Gao, R.; Yang, T.; Hou, X.; Cao, L. Three-dimensional NiCo2O4 nanosheets arrays on carbon nanofibers for high-performance asymmetric solid-state supercapacitor. Diam. Relat. Mater. 2021, 119, 108584. [Google Scholar] [CrossRef]
  9. Shi, Z.; Wei, S.; Zuo, H.; Huang, M.; Shi, J.; Wang, H. Boosting capacitance and energy density by construction NiCoO2/CoS2 nanocomposites arrays as pseudocapacitor. J. Alloys Compd. 2021, 881, 160627. [Google Scholar] [CrossRef]
  10. Cui, P.; Zhang, Y.; Cao, Z.; Liu, Y.; Sun, Z.; Cheng, S.; Wu, Y.; Fu, J.; Xie, E. Plasma-assisted lattice oxygen vacancies engineering recipe for high-performing supercapacitors in a model of birnessite-MnO2. Chem. Eng. J. 2021, 412, 128676. [Google Scholar] [CrossRef]
  11. Xie, Z.; Liu, L.; Li, Y.; Yu, D.; Wei, L.; Han, L.; Hua, Y.; Wang, C.; Zhao, X.; Liu, X. Synthesis of core-shell structured Ni3S2@MnMoO4 nanosheet arrays on Ni foam for asymmetric supercapacitors with superior performance. J. Alloys Compd. 2021, 874, 159860. [Google Scholar] [CrossRef]
  12. Sahoo, S.; Dhakal, G.; Kim, W.K.; Shim, J.J. Ternary Nanohybrid of Ni3S2/CoMoS4/MnO2 on Nickel Foam for Aqueous and Solid-State High-Performance Supercapacitors. Nanomaterials 2022, 12, 1945. [Google Scholar] [CrossRef]
  13. Vijayakumar, S.; Dhakal, G.; Kim, S.H.; Lee, J.; Lee, Y.R.; Shim, J.J. Facile Synthesis of Zn-Co-S Nanostrip Cluster Arrays on Ni Foam for High-Performance Hybrid Supercapacitors. Nanomaterials 2021, 11, 3209. [Google Scholar] [CrossRef]
  14. Ma, J.; Li, W.; Zhang, X.; Cheng, Y.; Zhang, F. Free-standing Ni3S2 nanowire derived from in-situ synthetized coordination supramolecular as electrode materials for high performance asymmetric supercapacitors. Appl. Surf. Sci. 2019, 507, 145074. [Google Scholar] [CrossRef]
  15. Kumar, M.; Jeong, D.I.; Sarwar, N.; Yoon, H.D. Heazlewoodite, Ni3S2: An electroactive material for supercapacitor application. Ceram. Int. 2021, 47, 16852–16860. [Google Scholar] [CrossRef]
  16. Ma, X.; Kang, Z. A facile electrodeposition technique for synthesis of nickel sulfides/carbon nanotubes nanocomposites as high performance electrodes for supercapacitor. Mater. Lett. 2019, 236, 468–471. [Google Scholar] [CrossRef]
  17. Huo, H.; Zhao, Y.; Xu, C. 3D Ni3S2 nanosheet arrays supported on Ni foam for high-performance supercapacitor and non-enzymatic glucose detection. J. Mater. Chem. A 2014, 2, 15111–15117. [Google Scholar] [CrossRef]
  18. Chen, L.; Guan, L.; Tao, J. Morphology control of Ni3S2 multiple structures and their effect on supercapacitor performances. J Mater. Sci. 2019, 54, 12737–12746. [Google Scholar] [CrossRef]
  19. Miao, J.; Wu, D.; Gao, X.; Wang, W.; Wu, W.; Tao, S.; Fang, Y.; Han, Z.; Qian, B.; Jiang, X.; et al. Ni3S2/rGO nanoparticles ensemble by an in-situ microwave irradiation route for supercapacitors. J. Alloys Compd. 2022, 890, 161435. [Google Scholar] [CrossRef]
  20. Cheng, Q.; Yang, C.; Han, L.; Tao, K. Construction of Hierarchical 2D PANI/Ni3S2 Nanosheet Arrays on Ni Foam for High-Performance Asymmetric Supercapacitors. Batter. Supercaps 2020, 3, 370–375. [Google Scholar] [CrossRef]
  21. Zhao, Y.; Guo, J.; Liu, A.; Ma, T. 2D heterostructure comprised of Ni3S2/d-Ti3C2 supported on Ni foam as binder-free electrode for hybrid supercapacitor. J. Alloys Compd. 2019, 814, 152271. [Google Scholar] [CrossRef]
  22. Li, X.; Sun, J.; Feng, L.; Zhao, L.; Ye, L.; Zhang, W.; Duan, L. Cactus-like ZnS/Ni3S2 hybrid with high electrochemical performance for supercapacitors. J. Alloys Compd. 2018, 753, 508–516. [Google Scholar] [CrossRef]
  23. Deng, X.; Fan, Y.; Zhou, Q.; Huang, H.; Zhou, W.; Lan, Z.; Liang, X.; Li, G.; Guo, J.; Tang, S. Self-supported Ni3S2/NiCo2O4 core-shell flakes-arrays on Ni foam for enhanced charge storage properties. Electrochim. Acta 2019, 319, 783–790. [Google Scholar] [CrossRef]
  24. Pan, X.; Zhao, L.; Liu, H.; Guo, M.; Han, C.; Wang, W. Hierarchical structure Ni3S2/Ni(OH)2 nanoarrays towards high-performance supercapacitors. J. Solid State Chem. 2022, 309, 122974. [Google Scholar] [CrossRef]
  25. Appiagyei, A.B.; Bonsu, J.O.; Han, J.I. Construction of NiCo-OH/Ni3S2 core-shell heterostructure wrapped in rGO nanosheets as efficient supercapacitor electrode enabling high stability up to 20,000 cycles. J. Electroanal. Chem. 2021, 889, 115226. [Google Scholar] [CrossRef]
  26. Wang, X.; Hu, J.; Su, Y.; Hao, J.; Liu, F.; Han, S.; An, J.; Lian, J. Ni Foam-Ni3S2@Ni(OH)2-Graphene Sandwich Structure Electrode Materials: Facile Synthesis and High Supercapacitor Performance. Chemistry 2017, 23, 4128–4136. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, G.; Qi, K.; Yan, Z.; Yue, L.; Ding, Y.; Li, W.; Xu, Z. Microwave hydrothermal synthesis of La decorated Ni(OH)2 nanosheets for performance-enhanced hybrid supercapacitor. Appl. Surf. Sci. 2022, 592, 153293. [Google Scholar] [CrossRef]
  28. Peng, Z.; Zou, H.; Yang, W.; Feng, Z.; Chen, S. Hierarchical porous Co9S8 nanowire arrays derived from zeolitic imidazolate framework on Ni foam for button-type asymmetric supercapacitor. J. Energy Storage 2021, 40, 102697. [Google Scholar] [CrossRef]
  29. Li, X.; Li, J.; Zhang, Y.; Zhao, P.; Lei, R.; Yuan, B.; Xia, M. The Evolution in Electrochemical Performance of Honeycomb-Like Ni(OH)2 Derived from MOF Template with Morphology as a High-Performance Electrode Material for Supercapacitors. Materials 2020, 13, 4870. [Google Scholar] [CrossRef] [PubMed]
  30. Xiao, Z.; Mei, Y.; Yuan, S.; Mei, H.; Xu, B.; Bao, Y.; Fan, L.; Kang, W.; Dai, F.; Wang, R.; et al. Controlled Hydrolysis of Metal-Organic Frameworks: Hierarchical Ni/Co-Layered Double Hydroxide Microspheres for High-Performance Supercapacitors. ACS Nano 2019, 13, 7024–7030. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, S.; Yang, Z.; Gong, K.; Xu, B.; Mei, H.; Zhang, H.; Zhang, J.; Kang, Z.; Yan, Y.; Sun, D. Temperature controlled diffusion of hydroxide ions in 1D channels of Ni-MOF-74 for its complete conformal hydrolysis to hierarchical Ni(OH)2 supercapacitor electrodes. Nanoscale 2019, 11, 9598–9607. [Google Scholar] [CrossRef]
  32. Jia, W.L.; Li, J.; Lu, Z.J.; Juan, Y.F.; Jiang, Q.Y. Synthesis of porous Co3O4/Reduced graphene oxide by a two-step method for supercapacitors with excellent electrochemical performance. J. Alloys Compd. 2019, 815, 152373. [Google Scholar] [CrossRef]
  33. Yue, L.; Wang, X.; Sun, T.; Liu, H.; Li, Q.; Wu, N.; Guo, H.; Yang, W. Ni-MOF coating MoS2 structures by hydrothermal intercalation as high-performance electrodes for asymmetric supercapacitors. Chem. Eng. J. 2019, 375, 121959. [Google Scholar] [CrossRef]
  34. Zhang, Z.; Huo, H.; Gao, J.; Yu, Z.; Ran, F.; Guo, L.; Lou, S.; Mu, T.; Yin, X.; Wang, Q.; et al. Ni-MOF derived NiO/C nanospheres grown in situ on reduced graphene oxide towards high performance hybrid supercapacitor. J. Alloys Compd. 2019, 801, 158–165. [Google Scholar] [CrossRef]
  35. Yang, W.; Guo, H.; Fan, T.; Zhao, X.; Zhang, L.; Guan, Q.; Wu, N.; Cao, Y.; Yang, W. MoS2/Ni(OH)2 composites derived from in situ grown Ni-MOF coating MoS2 as electrode materials for supercapacitor and electrochemical sensor. Colloids Surf. A Physicochem. Eng. Asp. 2021, 615, 126178. [Google Scholar] [CrossRef]
  36. Yang, T.; Huang, L.; Hou, W.; Guo, W.; Wang, S. The preparation of 3D Ni3S2/MnS2 composite by in-situ vulcanization for a hybrid supercapacitor. Mater. Lett. 2022, 319, 132274. [Google Scholar] [CrossRef]
  37. Nath, A.R.; Sandhyarani, N. SILAR deposited nickel sulphide-nickel hydroxide nanocomposite for high performance asymmetric supercapacitor. Electrochim. Acta 2020, 356, 136844. [Google Scholar] [CrossRef]
  38. Chang, Y.; Sui, Y.; Qi, J.; Jiang, L.; He, Y.; Wei, F.; Meng, Q.; Jin, Y. Facile synthesis of Ni3S2 and Co9S8 double-size nanoparticles decorated on rGO for high-performance supercapacitor electrode materials. Electrochim. Acta 2017, 226, 69–78. [Google Scholar] [CrossRef]
  39. Wei, J.; Han, D.; Pan, Y.; Gao, D.; Mao, L.; Wei, Y. Carbon decorated dendritic nickel sulfide coordinating with situ growth enabling hybrid supercapacitors with high specific energy and rate performance. J. Alloys Compd. 2021, 884, 161160. [Google Scholar] [CrossRef]
  40. Li, Q.; Guo, H.; Xue, R.; Wang, M.; Xu, M.; Yang, W.; Zhang, J.; Yang, W. Self-assembled Mo doped Ni-MOF nanosheets based electrode material for high performance battery-supercapacitor hybrid device. Int. J. Hydrogen Energy 2020, 45, 20820–20831. [Google Scholar] [CrossRef]
  41. Xiao, C.-Y.; Chen, T.-Y.; Chung, R.-J.; Yougbaré, S.; Lin, L.-Y.; Wu, Y.-F. Binder-free synthesis of metal organic framework derived cobalt sulfide on carbon cloth for efficient energy storage devices. J. Energy Storage 2022, 55, 105622. [Google Scholar] [CrossRef]
  42. Nandhini, S.; Muralidharan, G. Graphene encapsulated NiS/Ni3S4 mesoporous nanostructure: A superlative high energy supercapacitor device with excellent cycling performance. Electrochim. Acta 2020, 365, 137367. [Google Scholar] [CrossRef]
  43. Li, G.; Cong, Y.; Zhang, C.; Tao, H.; Sun, Y.; Wang, Y. Hierarchical nanosheet-based Ni3S2 microspheres grown on Ni foam for high-performance all-solid-state asymmetric supercapacitors. Nanotechnology 2017, 28, 425401. [Google Scholar] [CrossRef]
  44. Yang, Y.; Zhu, H.; Meng, H.; Ma, W.; Wang, C.; Ma, F.; Hu, Z. Metal-organic framework derived Co9S8/Ni3S2 composites on Ni foam with enhanced electrochemical performance by one-step sulfuration strategy for supercapacitors electrode. Colloids Surf. A Physicochem. Eng. Asp. 2021, 623, 126695. [Google Scholar] [CrossRef]
  45. Gao, X.; Zhang, X.; Lu, Q.; Guo, E.; Si, C.; Wei, M.; Pang, Y. High-performance hybrid supercapacitors based on hierarchical NiC2O4/Ni(OH)2 nanospheres and biomass-derived carbon. Appl. Surf. Sci. 2022, 595, 153571. [Google Scholar] [CrossRef]
  46. Chen, H.; Zhou, J.; Li, Q.; Zhao, S.; Yu, X.; Tao, K.; Hu, Y.; Han, L. MOF-assisted construction of a Co9S8@Ni3S2/ZnS microplate array with ultrahigh areal specific capacity for advanced supercapattery. Dalton Trans. 2020, 49, 10535–10544. [Google Scholar] [CrossRef] [PubMed]
  47. He, Y.; Liu, D.; Zhao, H.; Wang, J.; Sui, Y.; Qi, J.; Chen, Z.; Zhang, P.; Chen, C.; Zhuang, D. Carbon-coated NiMn layered double hydroxides/Ni3S2 nanocomposite for high performance supercapacitors. J. Energy Storage 2021, 41, 103003. [Google Scholar] [CrossRef]
  48. Zhao, J.; Song, L.; Liang, X.; Zhao, Z. Ion diffusion-assisted preparation of Ni3S2/NiO nanocomposites for electrochemical capacitors. Inorg. Chem. Commun. 2019, 107, 107469. [Google Scholar] [CrossRef]
  49. Zhang, T.F.; Lee, W.-J.; Kwon, S.-H.; Wan, Z. Facile syntheses and electrochemical properties of Ni(OH)2 nanosheets/porous Ni foam for supercapacitor application. Mater. Lett. 2019, 256, 126656. [Google Scholar] [CrossRef]
  50. Fan, J.; Chen, A.; Xie, X.; Gu, L. Co3O4-induced area-selective growth of Ni(OH)2 cellular arrays for high-capacity supercapacitor electrode. J. Energy Storage 2022, 48, 103964. [Google Scholar] [CrossRef]
  51. Xu, L.; Zhang, L.; Cheng, B.; Yu, J. Rationally designed hierarchical NiCo2O4–C@Ni(OH)2 core-shell nanofibers for high performance supercapacitors. Carbon 2019, 152, 652–660. [Google Scholar] [CrossRef]
  52. Li, X.; Li, J.; Zhang, Y.; Zhao, P. Synthesis of Ni-MOF derived NiO/rGO composites as novel electrode materials for high performance supercapacitors. Colloids Surf. A Physicochem. Eng. Asp. 2021, 622, 126653. [Google Scholar] [CrossRef]
  53. Mao, Y.; Li, T.; Guo, C.; Zhu, F.; Zhang, C.; Wei, Y.; Hou, L. Cycling stability of ultrafine β-Ni(OH)2 nanosheets for high capacity energy storage device via a multilayer nickel foam electrode. Electrochim. Acta 2016, 211, 44–51. [Google Scholar] [CrossRef] [Green Version]
  54. Xu, S.; Su, C.; Wang, T.; Ma, Y.; Hu, J.; Hu, J.; Hu, N.; Su, Y.; Zhang, Y.; Yang, Z. One-step electrodeposition of nickel cobalt sulfide nanosheets on Ni nanowire film for hybrid supercapacitor. Electrochim. Acta 2018, 259, 617–625. [Google Scholar] [CrossRef]
  55. Li, Z.; Yu, X.; Gu, A.; Tang, H.; Wang, L.; Lou, Z. Anion exchange strategy to synthesis of porous NiS hexagonal nanoplates for supercapacitors. Nanotechnology 2017, 28, 065406. [Google Scholar] [CrossRef] [PubMed]
  56. Bao, Y.; Deng, Y.; Wang, M.; Xiao, Z.; Wang, M.; Fu, Y.; Guo, Z.; Yang, Y.; Wang, L. A controllable top-down etching and in-situ oxidizing strategy: Metal-organic frameworks derived α-Co/Ni(OH)2@Co3O4 hollow nanocages for enhanced supercapacitor performance. Appl. Surf. Sci. 2019, 504, 144395. [Google Scholar] [CrossRef]
  57. He, Q.-Q.; Wang, H.-Y.; Lun, N.; Qi, Y.-X.; Liu, J.-R.; Feng, J.-K.; Qiu, J.; Bai, Y.-J. Fabricating a Mn3O4/Ni(OH)2 Nanocomposite by Water-Boiling Treatment for Use in Asymmetric Supercapacitors as an Electrode Material. ACS Sustain. Chem. Eng. 2018, 6, 15688–15696. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the preparation process for green Ni(OH)2/Ni3S2/NF.
Figure 1. Schematic illustration of the preparation process for green Ni(OH)2/Ni3S2/NF.
Nanomaterials 13 00493 g001
Figure 2. XRD patterns of the Ni(OH)2/Ni3S2/NF electrode.
Figure 2. XRD patterns of the Ni(OH)2/Ni3S2/NF electrode.
Nanomaterials 13 00493 g002
Figure 3. XPS spectra of the Ni(OH)2/Ni3S2/NF electrode. (a) Survey; (b) Ni 2p; (c) O 1s and (d) S 2p.
Figure 3. XPS spectra of the Ni(OH)2/Ni3S2/NF electrode. (a) Survey; (b) Ni 2p; (c) O 1s and (d) S 2p.
Nanomaterials 13 00493 g003
Figure 4. FE-SEM images of the electrodes. (a,b) pure NF; (c,d) Ni3S2/NF and (e,f) Ni(OH)2/Ni3S2/NF.
Figure 4. FE-SEM images of the electrodes. (a,b) pure NF; (c,d) Ni3S2/NF and (e,f) Ni(OH)2/Ni3S2/NF.
Nanomaterials 13 00493 g004
Figure 5. TEM and HRTEM images of (a,b) Ni3S2/NF; (c,d) Ni(OH)2/Ni3S2/NF.
Figure 5. TEM and HRTEM images of (a,b) Ni3S2/NF; (c,d) Ni(OH)2/Ni3S2/NF.
Nanomaterials 13 00493 g005
Figure 6. (ad) EDS mapping of Ni, S and O elements in Ni(OH)2/Ni3S2/NF.
Figure 6. (ad) EDS mapping of Ni, S and O elements in Ni(OH)2/Ni3S2/NF.
Nanomaterials 13 00493 g006
Figure 7. (a) CV curves of NF, Ni3S2/NF and Ni(OH)2/Ni3S2/NF at the scanning rate of 10 mV·s−1; (b) CV curves of Ni(OH)2/Ni3S2/NF at the scanning rate ranging from 5 mV·s−1 to 50 mV·s−1; (c) Specific capacitance of Ni(OH)2/Ni3S2/NF at different scanning rates.
Figure 7. (a) CV curves of NF, Ni3S2/NF and Ni(OH)2/Ni3S2/NF at the scanning rate of 10 mV·s−1; (b) CV curves of Ni(OH)2/Ni3S2/NF at the scanning rate ranging from 5 mV·s−1 to 50 mV·s−1; (c) Specific capacitance of Ni(OH)2/Ni3S2/NF at different scanning rates.
Nanomaterials 13 00493 g007
Figure 8. (a) GCD curves of NF, Ni3S2/NF and Ni(OH)2/Ni3S2/NF at the current density of 2 A·g−1; (b) GCD curves of Ni(OH)2/Ni3S2/NF at the current density from 1 to 10 A·g−1; (c) Specific capacitance of Ni(OH)2/Ni3S2/NF at different current densities.
Figure 8. (a) GCD curves of NF, Ni3S2/NF and Ni(OH)2/Ni3S2/NF at the current density of 2 A·g−1; (b) GCD curves of Ni(OH)2/Ni3S2/NF at the current density from 1 to 10 A·g−1; (c) Specific capacitance of Ni(OH)2/Ni3S2/NF at different current densities.
Nanomaterials 13 00493 g008
Figure 9. Cycling tests for Ni(OH)2/Ni3S2/NF at the scanning rate of 50 mV·s−1 up to 5000 cycles.
Figure 9. Cycling tests for Ni(OH)2/Ni3S2/NF at the scanning rate of 50 mV·s−1 up to 5000 cycles.
Nanomaterials 13 00493 g009
Figure 10. Nyquist plots of Ni3S2/NF and Ni(OH)2/Ni3S2/NF.
Figure 10. Nyquist plots of Ni3S2/NF and Ni(OH)2/Ni3S2/NF.
Nanomaterials 13 00493 g010
Figure 11. Schematic diagram of the Ni(OH)2/Ni3S2/NF//AC hybrid supercapacitor (HSC).
Figure 11. Schematic diagram of the Ni(OH)2/Ni3S2/NF//AC hybrid supercapacitor (HSC).
Nanomaterials 13 00493 g011
Figure 12. (a) CV curves of the AC and Ni(OH)2/Ni3S2/NF electrode at 50 mV·s−1; (b) CV curves of the HSC device at different scanning rate ranges from 5 to 100 mV·s−1; (c) GCD curves of the HSC device at different current density from 0.5 to 10 A·g−1; (d) Ragone plot of the Ni(OH)2/Ni3S2/NF//AC HSC.
Figure 12. (a) CV curves of the AC and Ni(OH)2/Ni3S2/NF electrode at 50 mV·s−1; (b) CV curves of the HSC device at different scanning rate ranges from 5 to 100 mV·s−1; (c) GCD curves of the HSC device at different current density from 0.5 to 10 A·g−1; (d) Ragone plot of the Ni(OH)2/Ni3S2/NF//AC HSC.
Nanomaterials 13 00493 g012
Table 1. Comparison of electrochemical performance of Ni(OH)2/Ni3S2/NF with those reported in the other related literatures.
Table 1. Comparison of electrochemical performance of Ni(OH)2/Ni3S2/NF with those reported in the other related literatures.
Refs.MaterialsElectrolytePerformanceCyclic stability
This workNi(OH)2/Ni3S2/NF6 M KOH2207 F·g−1, 1 A·g−15000, 90.4%
[24]Ni3S2/Ni(OH)23 M KOH2065 F·g−1, 1 A·g−13000, 82.1%
[35]MoS2/Ni(OH)21 M KOH2192 F·g−1, 1 A·g−110,000, 90.6%
[43]Ni3S2PVA/KOH981.8 F·g−1, 2 A·g−11000, 96.9%
[44]Co9S8/Ni3S2/NF2 M KOH1356 F·g−1, 1 A·g−12500, 80.8%
[45]NiCo2O4/Ni(OH)21 M KOH1336 F·g−1, 1 A·g−1/
Table 2. Comparison of capacitance retention of electrode materials with those reported in the other related literatures after different cycle number of charge–discharge tests.
Table 2. Comparison of capacitance retention of electrode materials with those reported in the other related literatures after different cycle number of charge–discharge tests.
Refs.MaterialsCycle numberCapacitance retention
This workNi(OH)2/Ni3S2/NF500090.4%
[46]Co9S8@Ni3S2/ZnS400079.7%
[47]C/NiMn-LDH/Ni3S2500084.25%
[48]Ni3S2/NiO400090.2%
[49]Ni(OH)2/NF300090%
[50]Co3O4/Ni(OH)2500080.1%
[51]NiCo2O4-C@Ni(OH)2500079%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shao, M.; Li, J.; Li, J.; Yan, Y.; Li, R. Synthesis of Ni3S2 and MOF-Derived Ni(OH)2 Composite Electrode Materials on Ni Foam for High-Performance Supercapacitors. Nanomaterials 2023, 13, 493. https://doi.org/10.3390/nano13030493

AMA Style

Shao M, Li J, Li J, Yan Y, Li R. Synthesis of Ni3S2 and MOF-Derived Ni(OH)2 Composite Electrode Materials on Ni Foam for High-Performance Supercapacitors. Nanomaterials. 2023; 13(3):493. https://doi.org/10.3390/nano13030493

Chicago/Turabian Style

Shao, Meng, Jun Li, Jing Li, Yanan Yan, and Ruoliu Li. 2023. "Synthesis of Ni3S2 and MOF-Derived Ni(OH)2 Composite Electrode Materials on Ni Foam for High-Performance Supercapacitors" Nanomaterials 13, no. 3: 493. https://doi.org/10.3390/nano13030493

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