Next Article in Journal
The Effect of Slag on the Mechanical Properties of Coralline-Activated Materials and the Formation and Transformation of Mineral Crystals
Next Article in Special Issue
Zeolitic Imidazolate Framework 67-Derived Ce-Doped CoP@N-Doped Carbon Hollow Polyhedron as High-Performance Anodes for Lithium-Ion Batteries
Previous Article in Journal
Structural Study of the Complex of cblC Methylmalonic Aciduria and Homocystinuria-Related Protein MMACHC with Cyanocobalamin
Previous Article in Special Issue
Layer-by-Layer Assembly of Polyelectrolytes on Urchin-like MnO2 for Extraction of Zn2+, Cu2+ and Pb2+ from Alkaline Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Preparation and Electrochemical Pseudocapacitive Performance of Mutual Nickel Phosphide Heterostructures

College of Materials and Chemistry & Chemical Engineering, Chengdu University of Technology, Chengdu 610059, China
*
Authors to whom correspondence should be addressed.
Crystals 2022, 12(4), 469; https://doi.org/10.3390/cryst12040469
Submission received: 17 January 2022 / Revised: 11 February 2022 / Accepted: 11 February 2022 / Published: 28 March 2022
(This article belongs to the Special Issue Emerging Low-Dimensional Materials)

Abstract

:
Transition metal phosphide composite materials have become an excellent choice for use in supercapacitor electrodes due to their excellent conductivity and good catalytic activity. In our study, a series of nickel phosphide heterostructure composites was prepared using a temperature-programmed phosphating method, and their electrochemical performance was tested in 2 mol L−1 KOH electrolyte. Because the interface effect can increase the catalytic active sites and improve the ion transmission, the prepared Ni2P/Ni3P/Ni (Ni/P = 7:3) had a specific capacity of 321 mAh g−1 under 1 A g−1 and the prepared Ni2P/Ni5P4 (Ni/P = 5:4) had a specific capacity of 218 mAh g−1 under 1 A g−1. After the current density was increased from 0.5 A g−1 to 5 A g−1, 76% of the specific capacity was maintained. After 7000 cycles, the capacity retention rate was above 82%. Due to the phase recombination effect, the electrochemical performance of Ni2P/Ni3P/Ni and Ni2P/Ni5P4 was much better than that of single-phase N2P. After assembling the prepared composite and activated carbon into a supercapacitor, the Ni2P/Ni3P/Ni//AC had an energy density of 22 W h kg−1 and a power density of 800 W kg−1 and the Ni2P/Ni5P4//AC had an energy density of 27 W h kg−1 and a power density of 800 W kg−1.

1. Introduction

With the rapid development of human society, the demand for energy has increased substantially. At the same time, the energy crisis is becoming increasingly more serious. In the future, exploring and developing clean and renewable energy such as wind and hydropower will become the mainstream [1,2,3,4,5,6]. However, due to the instability of power generation, this type of energy has not yet fully met the demand [7,8]. Through electrochemical energy storage devices, the electricity generated from these renewable energy sources can be stored for effective use [9,10]. Presently, there are two main types of commercial electrochemical energy storage devices: batteries, such as lithium-ion batteries [11,12], and supercapacitors, such as pseudocapacitance supercapacitors [13,14]. Supercapacitors have many advantages, such as higher power density, long cycle life, and fast charging and discharging, so they have attracted much attention [15,16,17]. However, there are still some challenges associated with supercapacitors, including how to increase their energy density while maintaining the above advantages [18,19].
Electrode material is the core factor of the supercapacitor system and affects the performance of the supercapacitor directly [20]. According to their characteristics, electrode materials can be divided into two categories [21]. The first is a variety of carbon materials including activated carbon, graphene, and carbon nanotubes; the energy storage mechanism is electrostatic adsorption and desorption [9]. These electrode materials provide good stability, porous characteristics, and good electrical conductivity [22], but the relatively low energy density associated with this group limits its wide application [23]. The second category is pseudocapacitive materials including transition metal oxides and transition metal phosphides; the energy storage mechanism is rapid reversible redox or intercalation reactions on or near the electrode surface [24,25]. Compared to transition metal oxides, transition metal phosphides have better electrical conductivity and a lower cost. They also have a specific capacity comparable to or even beyond that of transition metal oxides [26,27]. This is due to some of the characteristics of transition metal phosphides.
Phosphorus atoms can enter the transition metal crystals to form intermetallics [28]. The presence of phosphorus atoms pulls the electron delocalization of metal phosphide, which improves its catalytic activity. The higher the metal content, the more free electrons it contains and the better it is for conducting electricity [29]. Therefore, this compound with metallic properties exhibits high electrical conductivity and specific capacity [30,31]. Nickel phosphide has been a good choice for electrode materials for supercapacitors because of its high electrical conductivity, fast charge transfer ability, good reaction kinetics, and abundant earth reserves [32,33,34]. Single-phase nickel phosphide is difficult to prepare, and its electrochemical performance is unsatisfactory [35]. Therefore, many researchers have shown great interest in preparing nickel phosphide composites and applying them to supercapacitors [36,37]. Nickel phosphide compounds can be prepared using available nickel chemical plating methods on a nickel phosphide surface coated with a layer of amorphous nickel or by mechanically mixing graphene and nickel phosphide. It is also possible to compound nickel phosphide with other compounds through chemical precipitation to make composite materials [38,39,40]. Other types of composites used in the study of the electrochemical performance of supercapacitors include nickel-cobalt oxide modified with reduced graphene oxide, ZnFe2O4 nanorods on reduced graphene oxide, NiCo2O4/Ni2P, nitrogen-doped Ni2P/Ni12P5/Ni3S2, and MoS2–ReS2/rGO [41,42,43,44,45]. Due to the synergistic effect between different components and the interfacial effect, the electrochemical performance of such materials is satisfactory. The preparation of composite nickel phosphide shows good electrochemical performance, which provides inspiration for exploring such electrode materials.
In our study, nickel phosphide composite was prepared in one step by temperature programmed phosphating. Since no polymer binder was added during the preparation of the electrode, the electrode had maximum conductivity and catalytic activity. We used 2 mol L−1 KOH aqueous solution as the electrolyte for electrochemical performance testing. The best comprehensive electrochemical performance was observed when the stoichiometric ratio was 5:4. At this stoichiometric ratio, the nickel phosphide composite formed a Ni2P/Ni5P4 heterostructure with a specific capacity of 218 mAh g−1 and a rate performance of 76%. After 7000 cycles, the capacity retention rate was above 82%. After combining it with activated carbon to form an asymmetric supercapacitor, the energy density was 27 W h kg−1 while the power density was 800 W kg−1, demonstrating good electrochemical performance. At the stoichiometric ratio of 7:3, it formed a Ni2P/Ni3P/Ni heterostructure with a specific capacity of 321 mAh g−1 and a rate performance of 59%. After combining it with activated carbon to form an asymmetric supercapacitor, the energy density was 22 W h kg−1 while the power density was 800 W kg−1.

2. Experimental Section

2.1. Chemicals

All chemicals used were of analytical grade. C2H5OH, KOH, H2SO4, and acetone were purchased from Kelon Chemicals Co. Ltd., Chengdu, China. Nickel foam was purchased from Shanghai (China) Hesen Electric Co. Ltd. Nickel powder and red phosphorus were from Aladdin.

2.2. Electrode Material Synthesis

The nickel foam was cut into small pieces, each with a length and width of 1 cm, and was then pretreated with dilute hydrochloric acid solution and acetone solution (Vacid:Vacetone = 6:1) to remove oxides and organic pollutants. The surface was then rinsed with anhydrous ethanol and deionized water. The metal nickel powder was pretreated with dilute sulfuric acid to remove oxides and organic pollutants, and then rinsed with anhydrous ethanol and deionized water several times. It was put into the oven at a constant temperature of 60 °C until dried. About 5 g of treated metal nickel powder for separated for use. Metal nickel powder and red phosphorus were mixed according to the stoichiometric amounts; a 1.5% excess of red phosphorous was required.
To make electrodes, slightly more than 4 mg of mixed powder was pressed on the treated nickel foam; the applied pressure was 10 MPa. The foam nickel electrode pads loaded with mixed powder were put into a porcelain boat along with the remaining mixed powder. The porcelain boat was then put into a tubular furnace, washed to vacuum with nitrogen, gradually heated to 700 °C at 4 °C min−1, and held for 6 h. The intermediate temperatures were 350 °C, 450 °C, and 550 °C. Each intermediate temperature was held for 1 h. According to the stoichiometric amount, we recorded them as (3:1), (5:2), (12:5), (7:3), (2:1), (5:4), and (1:1). Single-phase Ni2P can be made from stoichiometry (5:2).

2.3. Material Characterization

The crystalline structures were confirmed by an X-ray diffraction (XRD, Dandong DX-2700, Dandong, China) with Cu-Kα radiation (2θ = 5~80°) operating at 40 kV. The morphologies and microstructures of the samples were characterized by field-emission scanning electron microscope (SEM, JEOL JSM-6701 F, Tokyo, Japan) and transmission electron microscope (TEM, JEOL JEM2010, Tokyo, Japan). The pore properties and Brunauer-Emmett-Teller specific surface area were investigated via N2 adsorption–desorption (NAD, ASAPR 2020, Atlanta, GA, USA) test at −196 °C. The surface bonding state was tested by X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha, New York, NY, USA).

2.4. Electrochemical Evaluation

All electrochemical measurements were executed on the electrochemical station, and the electrolyte was 2 mol L−1 aqueous KOH. Before the test, we soaked the undertested materials in 2 mol L−1 of KOH electrolyte for 24 h to activate them. The mass of the active substance was obtained by weighing the nickel foam before and after the reaction. The mass of the active material on the nickel foam electrode was observed as 4 mg. All electrochemical measurements for the single electrode were executed in a three-electrode system with a platinum plate electrode as a counter electrode and a Hg/HgO electrode as a reference electrode.
The formula of mass specific capacitance is as follows:
C s = I d Δ t 3.6
Calculated by integrating the slope of the discharge curve for the asymmetric device, the formula for energy density (W h kg−1) and power density (W kg−1) is as follows [46]:
E = I d 3.6 E   d t
P = 3600 E Δ t
The supercapacitor was assembled with activated carbon as a cathode and the prepared electrode as an anode. The mass balancing of cathode and anode will follow the equation:
m + m = C s Δ V C s + Δ V +
where Cs (mAh g−1) is the specific capacitance, Id is the specific current, and Δt is the discharge time.

3. Results and Discussion

3.1. Characterization of Materials

The crystal structure of each Ni/P ratio and single-phase Ni2P are shown in Figure 1. In Figure 1d, some diffraction peaks appear and correspond to the crystal planes of standard PDF card Ni3P, Ni2P, and Ni. This XRD pattern shows the formation of Ni2P/Ni3P/Ni heterostructures with a Ni/P ratio of 7:3. Due to the interaction between the different phases, the positions of the diffraction peaks are slightly shifted, which indicates the successful synthesis of the heterostructure and has a positive effect on the electrochemical performance of the electrode. Due to the presence of metallic nickel, the catalytic activity of this electrode was greatly improved. In Figure 1f, some diffraction peaks appear and correspond to the crystal planes of standard PDF card Ni2P, and Ni5P4. This XRD pattern shows that Ni2P/Ni5P4 heterostructures are formed at a Ni/P ratio of 5:4. The combination of different phases produces a certain lattice distortion, so the position of the corresponding diffraction peak is slightly shifted compared to the standard card. Although only two nickel phosphide phases exist, the interfacial effect between the phases can still improve the electrochemical performance of the electrode, and at the same time will improve its catalytic stability. It can be observed from Figure 1h that diffraction peaks appear at 17.463°, 26.330°, 30.489°, 31.771°, 35.350°, 40.714°, 44.611°, 47.362°, 54.196°, 54.998°, 66.371°, 72.719°, and 74.790°, corresponding to (100), (001), (110), (101), (200), (111), (201), (210), (300), (211), (310), (311), and (400) crystal planes of standard PDF card Ni2P, respectively. This XRD pattern shows the formation of single-phase Ni2P. It can be observed from Figure 1a–c,e,g that Ni3P/Ni2P/Ni, Ni5P4/Ni2P, Ni5P4/Ni2P/Ni, Ni5P4/NiP2/Ni, and Ni5P4/Ni2P heterostructures are formed at a Ni/P ratio of 3:1, 5:2, 12:5, 2:1, and 1:1, respectively. Such heterostructures will contribute to electrochemical performance.
The SEM images of each Ni/P ratio and single-phase Ni2P are shown in Figure 2. In addition to the single-phase nickel phosphide, two distinct structures can be seen in each image, which confirms the successful preparation of the composite phase. Figure 2d is the SEM image of a Ni/P ratio of 7:3. It can be observed from this image that the linear structure is wrapped around the granular structure, increasing the contact area between the two. Figure 2f is the SEM image of a Ni/P ratio of 5:4. It can be observed from this image that the villous structure grows on the spherical particles, and there are small particles interspersed between them. Such a combination of different phases will produce lattice distortion at the interface, thereby improving the charge transfer efficiency and enhancing the catalytic activity.
Figure 2a is the SEM image of a Ni/P ratio of 3:1. It can be observed from this image that braided linear structures and granular structures intersected with each other are generated under such condition. Figure 2b is the SEM image of a Ni/P ratio of 5:2. It can be observed from this image that under this condition, acicular and granular structures are formed, and most of the acicular structures are embedded between large and small particles. Figure 2c is the SEM image of a Ni/P ratio of 12:5. It can be observed from this image that prismatic and granular structures form under this condition, with granular structures surrounding the prism. It can also be observed that the prism is formed by a combination of small acicular structures. Figure 2e is the SEM image of a Ni/P ratio of 2:1. It can be observed from this image that the acicular and granular structures are bound together. Figure 2g is the SEM image of a Ni/P ratio of 1:1. It can be observed from this image that the bean-sprout-shaped structure is mixed with the lamellar structure. Figure 2h is the SEM image of single-phase Ni2P. It can be observed from this image that large and small particles are not evenly distributed.
In order to explore the structure of the prepared nickel phosphide composite, we performed the TEM test on Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4). Figure 3a,d are the TEM images of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4). Both images display different shapes of substances. This is the result of different phase recombinations of nickel phosphide, corresponding to the SEM image. Figure 3b is the HRTEM image of Ni2P/Ni3P/Ni (Ni/P = 7:3). We can observe the (111) crystal plane of Ni2P, (112) crystal planes of Ni3P and (111) crystal planes of Ni, which prove the formation of the nickel phosphide composite phase. Figure 3e is the HRTEM image of Ni2P/Ni5P4 (Ni/P = 5:4). We can observe the (111) crystal plane of Ni2P and (103) crystal planes of Ni5P4, which further proves the formation of the nickel phosphide composite phase. The formation of the composite phase can effectively improve the catalytic activity of electrode materials. Different phases have different crystal lattices. The interface at the grain boundary when recombined is an inhomogeneous interface, and this causes lattice distortion due to the Jahn-Teller effect. The electronic state at the grain boundary is changed, and unpaired electrons may appear, which facilitates the transfer of charge. At the same time, the accumulated electrons can also generate a built-in electric field to promote the transmission of ions. Some randomly arranged atoms (marked by green dots) were also observed in image 3b,e; these disordered atoms rearrange in order to balance when the external environment changes. This generates more interfaces, greatly increases the efficiency of ion transport, and improves electrochemical performance [47]. We can observe the (311), (300) crystal plane of Ni2P and (301) crystal planes of Ni3P in Figure 3c, (102), (213) crystal plane of Ni5P4 and (311) crystal planes of Ni2P in Figure 3f. These crystal planes correspond to XRD test results, which proves the formation of the nickel phosphide composite phase.
Figure 4 displays the BET test and pore size distribution information. Figure 4a,c show the BET test information of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4), respectively. From that information we can conclude that the BET surface area of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4) are 1.32 m2 g−1 and 0.54 m2 g−1, respectively. Figure 4b,d show the pore size distribution information of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4), and their pore size distribution is between 2–50 nm, belonging to mesoporous materials. Such a porous structure can provide more ion channels for the catalytic reaction, thus having high-efficiency energy storage effects and good electrochemical performance.
X-ray photoelectron spectroscopy (XPS) was used to test and analyze the electronic structure of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4). Figure 5a is the XPS spectrum of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4) on the Ni 2p orbital. From the figure, we can observe that single-phase Ni2P has peaks at the positions where the binding energy is 870.62 eV and 853.41 eV, which correspond to the Ni2+ 2p1/2 and Ni2+ 2p3/2 orbitals, respectively. Ni2P/Ni5P4 (Ni/P = 5:4) has peaks at the positions where the binding energy is 870.22 eV and 853.09 eV, which correspond to the Ni2+ 2p1/2 and Ni2+ 2p3/2 orbitals, respectively. Compared with single-phase Ni2P, the binding energy of Ni2P/Ni5P4 (Ni/P = 5:4) has a negative offset of 0.4 eV and 0.32 eV in these two orbitals, respectively. This shows that the formation of the Ni2P/Ni5P4 (Ni/P = 5:4) led to the occurrence of electronic remodeling, and the electrons flow from Ni5P4 to Ni2P. This is due to the difference in Fermi energy levels. The electron density around the Ni2P increases, and the electronic interaction between the Ni2P and the Ni5P4 may increase the catalytic activity [48]. At the same time, it can be observed that single-phase Ni2P has peaks at the positions where the binding energy is 875.4 eV and 857.46 eV and Ni2P/Ni5P4 (Ni/P = 5:4) has peaks at the positions where the binding energy is 875 eV and 857.16 eV, corresponding to Ni3+ 2p1/2 and Ni3+ 2p3/2 orbitals, respectively. This proves that Ni has multiple valence states, so the electrons can transition and cause Debye relaxation [49], which facilitates the occurrence of redox reactions. In addition, studies have shown that the redistribution of interface electrons has a positive effect on conductivity [50,51]. The other peaks in this image are satellite peaks. Figure 5b is the XPS spectrum of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4) on the P 2p orbital. Single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4) have peaks at binding energy 129.92 eV and 129.39 eV, respectively. Compared with single-phase Ni2P, the binding energy of Ni2P/Ni5P4 (Ni/P = 5:4) has a negative offset of 0.53 eV; this can once again prove the existence of charge transfer. As for the presence of P-O bonds, this is related to the surface reaction that occurs when the sample comes in contact with air during the transfer process [11,12].

3.2. Electrochemical Performance

The electrochemical test graphs of different nickel–phosphorus ratios are shown in Figure 6. Figure 6a shows the galvanostatic charge–discharge curves of electrodes prepared with different Ni/P ratios at 1 A g−1. The curve of each electrode material has the appearance of a platform, which indicates that the redox reaction occurs during the charging and discharging process and corresponds to the redox peak in the cyclic voltammetry curve [30,31]. From this figure, it can be clearly seen that when the ratio of nickel to phosphorus is 7:3, it has the highest specific capacity of 321 mAh g−1. Secondly, when the ratio of nickel to phosphorus is 5:4, the specific capacity is 218 mAh g−1. The specific capacitance of each electrode material can be calculated according to Equation (1). The composites prepared with different raw material ratios and their electrochemical performances are shown in Table 1. The faradaic reaction corresponding to the redox peak is as follows [40]:
Ni2+ + 2OH → Ni(OH)2
Ni(OH)2 + OH ↔ NiOOH + H2O + e
Figure 6b is the cyclic voltammetry curve of various Ni/P ratios at 3 mV s−1. The area is the largest when the ratio of nickel to phosphorus is 7:3, followed by a ratio of 5:4, which corresponds to its specific capacity. From the platform of the galvanostatic charge–discharge curve and the obvious redox peak of the cyclic voltammetry curve, it can be judged that the electrode material is a typical pseudocapacitive material. Figure 6c is the electrochemical impedance spectroscopy of different nickel–phosphorus ratios, and Figure 6d is an enlarged view of their high frequency region. Each curve has an intersection point in the high frequency area with the real axis, representing the equivalent series resistance (Res). The resistance of this part is related to the interface resistance, internal resistance, and electrolyte ion resistance. It can be seen from Figure 6d that the high-frequency region has a shape similar to a semicircle, and the diameter of the semicircle can essentially represent the charge transfer resistance (Rct). The linear slope in the low-frequency region is related to the diffusion resistance, which is related to the characteristics of the electrolyte and the material itself. With a larger slope comes a faster ion diffusion and stronger capacitance. It can be seen from the curve that when the ratio of nickel to phosphorus is 7:3, its equivalent series resistance and charge transfer resistance are the smallest, and its slope is the largest; therefore, it has good electrochemical performance. Figure 6e is the Bode graph of different ratios of nickel to phosphorus. At the phase angle of −45°, the nickel–phosphorus ratio of 7:3 corresponds to the largest frequency; therefore, its time constant is the smallest, followed by the time constant of the nickel–phosphorus ratio of 5:4. This proves that these two ratios need less time to reach the steady state. There are two main reasons for the relatively high electrochemical performance of the composites: electron transfer can occur between different components and the presence of heterojunctions improves the catalytic activity.
The electrochemical test graphs of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P are shown in Figure 7. Figure 7a displays the galvanostatic charge–discharge curves of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P at 1 A g−1. According to Equation (1), their specific capacities are 321 mAh g−1, 218 mAh g−1, and 58 mAh g−1, respectively. This shows that compared with single-phase Ni2P, the preparation of the composite greatly improves its specific capacity. Figure 7b shows the cyclic voltammetry curves of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P at sweep rate of 3 mV s−1. The areas of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4) are much larger than that of the single-phase nickel phosphide, which corresponds to the specific capacity. Figure 7c shows the Nyquist plots of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P. The equivalent series resistances (Res) of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P are 0.838 Ω, 1.473 Ω, and 0.862 Ω, respectively. The charge transfer resistances (Rct) of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P are 0.056 Ω, 0.111 Ω, and 0.178 Ω, respectively. Figure 7d is the bode plot of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P. The time constants of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P are 0.87 s, 4.01 s, and 43 s, respectively. Comprehensive electrochemical test results showed that the performance of nickel phosphide composite was better than that of single-phase nickel phosphide. The electrochemical performances of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase nickel phosphide are listed in Table 2.
In order to further understand the electrochemical performance of nickel phosphide composites, further tests were done on materials with better performance. Figure 8 is the electrochemical test graph of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4). Figure 8a,b are the galvanostatic charge–discharge curves of Ni2P/Ni3P/Ni (Ni/P = 7:3) under different current densities and cyclic voltammetry curves of Ni2P/Ni3P/Ni (Ni/P = 7:3) at different scan rates. As the current density increases, the specific capacitance shows a downward trend, which is caused by the penetration of electrons and ions into the electrode surface [35]. The current density increased from 0.5 A g−1 to 5 A g−1 and its rate capability is 59%. The appearance of the redox peak is related to the surface Faraday reaction. As the sweep speed increases, the area of the cyclic voltammetry curve also increases, and the redox peaks move to higher and lower windows, respectively. This is related to the diffusion kinetics and internal resistance [30,31]. After 2000 cycles, the capacity retention rate was above 66% as shown in the Figure 8f. Figure 8d,e are the galvanostatic charge–discharge curves of Ni2P/Ni5P4 (Ni/P = 5:4) under various current densities and the cyclic voltammetry curves of Ni2P/Ni5P4 (Ni/P = 5:4) at various scan rates. The current density increased from 0.5 A g−1 to 5 A g−1, and its rate capability was 76%. As the sweep speed increased, the area of the cyclic voltammetry curve also increased, and the oxidation peak gradually disappeared. This may be because the sweep speed was too fast, and the changes caused by the redox reaction were not recorded in time. After 7000 cycles, the capacity retention rate was above 82% as shown in the Figure 8f. For scientific analysis, we compared the electrochemical properties of our prepared electrodes with those of the electrode materials reported in the literature, as shown in Table 3. The electrochemical performance of our electrode material was relatively good.
In order to deeply evaluate the electrochemical performance of the prepared materials in practical applications, we assembled a supercapacitor with Ni2P/Ni3P/Ni (Ni/P = 7:3) as the cathode and activated carbon as the anode. We also assembled a supercapacitor with Ni2P/Ni5P4 (Ni/P = 5:4) as the cathode and activated carbon as the anode. Figure 9a shows the galvanostatic charge–discharge curves of the Ni2P/Ni3P/Ni//AC and Ni2P/Ni5P4//AC supercapacitors at a current density of 1 A g−1. According to Equation (1), the specific capacity of supercapacitor Ni2P/Ni3P/Ni//AC at 1 A g−1 is 27.72 mAh g−1, and according to Equations (2) and (3), it has an energy density of 22 W h kg−1 while its power density is 800 W kg−1. The specific capacity of supercapacitor Ni2P/Ni5P4//AC at 1 A g−1 is 33.39 mAh g−1, and it has an energy density of 27 W h kg−1 while its power density is 800 W kg−1. Figure 9b displays the cyclic voltammetry curves of the Ni2P/Ni3P/Ni//AC and Ni2P/Ni5P4//AC supercapacitors at a scan rate of 100 mV s−1. The area of Ni2P/Ni5P4//AC is larger than that of Ni2P/Ni3P/Ni//AC, which corresponds to the specific capacity of the two supercapacitors. From the comprehensive test results, the electrochemical performance of Ni2P/Ni5P4//AC was better than that of Ni2P/Ni3P/Ni//AC; therefore, we further explored Ni2P/Ni5P4//AC. Figure 9c is the galvanostatic charge–discharge curve of the Ni2P/Ni5P4//AC supercapacitor at various current densities. We noticed that there are no obvious charging or discharging platforms in the galvanostatic charge–discharge curves, which indicates that the assembled device has good capacitance characteristics and a good electron transfer rate. Figure 9d shows the cyclic voltammetry curves of the Ni2P/Ni5P4//AC supercapacitor at various scan rates. We noticed that the CV curve has no obvious redox peak, indicating that it is suitable for high-power output, and the shape of the curve does not change with the scanning speed, indicating that the assembled supercapacitor has a rate capability. The CV and GCD curves correspond to each other. With a larger specific capacitance comes a longer charge and discharge time in the GCD and a larger CV area. The relatively large specific capacitance of Ni2P/Ni5P4//AC may be mainly due to its better stability.
In order to further study the internal reasons for the difference in performance of the different electrode materials prepared, we performed the work shown in Figure 10 including the linear sweep volt–ampere curve, schematic diagram of heterogeneous interface dangling bonds (unpaired electronics) generation, and energy band diagrams of metal and semiconductors. Figure 10a is the current–voltage curve of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single phase Ni2P tested by linear sweep voltammetry. Clearly, the conductivity of electrode Ni2P/Ni3P/Ni (Ni/P = 7:3) is the largest, followed by electrode Ni2P/Ni5P4 (Ni/P = 5:4), and the conductivity of electrode single-phase Ni2P is the worst. We recognize that there is an interface between the two-phase composite. Due to the difference in the phase structure, the lattice mismatch at the interface is inevitable. At this time, a part of the unsaturated bond will appear in the semiconductor material with a smaller crystal lattice at the interface as shown in the Figure 10b. These unsaturated bonds are dangling bonds, which form unpaired electrons. The bond density is affected by different semiconductor lattice constants and the crystal plane as the interface, which is determined by the following formula:
Ns = Ns1Ns2
Ns1, Ns2 are the bond density of the two semiconductor materials at the interface, respectively.
These dangling bonds, i.e., the unpaired electrons, are in a relatively free state and are easily excited to become free electrons, which provides excellent conditions for high-efficiency ion transmission and more active sites, thereby improving the electrochemical performance of the electrode. The prepared working electrodes with heterostructures all have such an interface as Ni2P/Ni5P4 (Ni/P = 5:4). As shown in Figure 10c, the work function of the metal is greater than the work function of the semiconductor. The Mott-Schottky interface shown in Figure 10d is formed after the metal is in contact with the semiconductor. Due to the difference in work function, electrons flow from the metal to the semiconductor, carriers flow from the semiconductor to the metal, and the accumulation of electrons in the semiconductor forms an internal potential field, which provides a high-speed channel for the continuous transmission of free electrons [52]. This improves the utilization of active sites, promotes the occurrence of redox reactions, and makes the working electrode have better capacitance characteristics. The prepared working electrodes with metallic nickel all have such an effect as Ni2P/Ni3P/Ni (Ni/P = 7:3).

4. Conclusions

In summary, the nickel phosphide composites were prepared in one step using the temperature-programmed phosphating method, and the prepared nickel phosphide composite showed good electrochemical performance. The prepared compound Ni2P/Ni3P/Ni (Ni/P = 7:3) had a specific capacity of 321 mAh g−1 under 1 A g−1, and the prepared compound Ni2P/Ni5P4 (Ni/P = 5:4) had a specific capacity of 218 mAh g−1 under 1 A g−1 and a 76% rate performance. After 7000 cycles, the capacity retention rate was above 82%. After assembling the prepared composite and activated carbon into a supercapacitor, Ni2P/Ni3P/Ni//AC was found to have an energy density of 22 W h kg−1 and a power density of 800 W kg−1, while Ni2P/Ni5P4 //AC had an energy density of 27 W h kg−1 and a power density of 800 W kg−1. These results provide ideas for further research.

Author Contributions

Conceptualization, Methodology, Software, Investigation, Writing-original draft. S.-B.G.; Supervision, Resources, Writing—review & editing. W.-B.Z. and J.L.; Visualization, Formal analysis, Validation. X.B., L.Z. and Y.-W.G.; Software, Validation. Z.-Q.Y. and X.-W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was financially supported by the Sichuan Science and Technology Project (No.2020YJ0163) and the Research Foundation for Teacher Development of Chengdu University of Technology (No.10912-2019KYQD-06847).

Conflicts of Interest

The authors declare no conflict of interest.

Highlights

-
A relatively simple method was used to prepare energy storage electrode materials with good electrochemical properties.
-
The Ni2P/Ni5P4 electrode possessed satisfactory specific capacitance and excellent electrochemical kinetics.
-
The assembled Ni2P/Ni5P4//AC based on nickel phosphide heterostructure showed remarkable performance.

References

  1. Winter, M.; Brodd, R. What Are Batteries, Fuel Cells, and Supercapacitors? Chem. Rev. 2004, 104, 4245–4269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Zhang, L.; Zhao, X.S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38, 2520–2531. [Google Scholar] [CrossRef] [PubMed]
  3. Li, S.; Yu, C.; Yang, J.; Zhao, C.; Zhang, M.; Huang, H.; Liu, Z.; Guo, W.; Qiu, J. A superhydrophilic “nanoglue” for stabilizing metal hydroxides onto carbon materials for high-energy and ultralong-life asymmetric supercapacitors. Energy Environ. Sci. 2017, 10, 1958–1965. [Google Scholar] [CrossRef]
  4. Bao, X.; Zhang, W.B.; Zhang, Q.; Zhang, L.; Ma, X.J.; Long, J. Interlayer material technology of manganese phosphate toward and beyond electrochemical pseudocapacitance over energy storage application. J. Mater. Sci. Technol. 2021, 71, 109–128. [Google Scholar] [CrossRef]
  5. Han, X.-W.; Zhang, W.-B.; Ma, X.-J.; Zhou, X.; Zhang, Q.; Bao, X.; Guo, Y.-W.; Zhang, L.; Long, J. Review—Technologies and Materials for Water Salinity Gradient Energy Harvesting. J. Electrochem. Soc. 2021, 168, 90505. [Google Scholar] [CrossRef]
  6. Yu, X.; Ren, X.; Zhang, Y.; Yuan, Z.; Sui, Z.; Wang, M. Self-supporting hierarchically micro/nano-porous Ni3P-Co2P-based film with high hydrophilicity for efficient hydrogen production. J. Mater. Sci. Technol. 2021, 65, 118–125. [Google Scholar] [CrossRef]
  7. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [Green Version]
  8. Shao, Y.; El-Kady, M.F.; Sun, J.; Li, Y.; Zhang, Q.; Zhu, M.; Wang, H.; Dunn, B.; Kaner, R.B. Design and Mechanisms of Asymmetric Supercapacitors. Chem. Rev. 2018, 118, 9233–9280. [Google Scholar] [CrossRef]
  9. Dou, Q.; Park, H.S. Perspective on High-Energy Carbon-Based Supercapacitors. Energy Environ. Mater. 2020, 3, 286–305. [Google Scholar] [CrossRef]
  10. Zhang, L.; Zhang, W.B.; Chai, S.S.; Han, X.W.; Zhang, Q.; Bao, X.; Guo, Y.W.; Zhang, X.L.; Zhou, X.; Guo, S.B.; et al. Review—Clay Mineral Materials for Electrochemical Capacitance Application. J. Electrochem. Soc. 2021, 168, 70558. [Google Scholar] [CrossRef]
  11. Ran, Z.; Shu, C.; Hou, Z.; Cao, L.; Liang, R.; Li, J.; Hei, P.; Yang, T.; Long, J. Ni3Se2/NiSe2 heterostructure nanoforests as an efficient bifunctional electrocatalyst for high-capacity and long-life Li–O2 batteries. J. Power Sources 2020, 468, 228308. [Google Scholar] [CrossRef]
  12. Ran, Z.; Shu, C.; Hou, Z.; Zhang, W.; Yan, Y.; He, M.; Long, J. Modulating electronic structure of honeycomb-like Ni2P/Ni12P5 heterostructure with phosphorus vacancies for highly efficient lithium-oxygen batteries. Chem. Eng. J. 2021, 413, 127404. [Google Scholar] [CrossRef]
  13. Abdollahifar, M.; Liu, H.W.; Lin, C.H.; Weng, Y.T.; Sheu, H.S.; Lee, J.F.; Lu, M.L.; Liao, Y.F.; Wu, N.L. Enabling Extraordinary Rate Performance for Poorly Conductive Oxide Pseudocapacitors. Energy Environ. Mater. 2020, 3, 405–413. [Google Scholar] [CrossRef]
  14. Brezesinski, T.; Wang, J.; Tolbert, S.H.; Dunn, B. Ordered mesoporous alpha-MoO3 with iso-oriented nanocrystalline walls for thin-film pseudocapacitors. Nat. Mater. 2010, 9, 146–151. [Google Scholar] [CrossRef] [PubMed]
  15. He, X.; Li, Z.; Hu, Y.; Li, F.; Huang, P.; Wang, Z.; Jiang, J.; Wang, C. Three-dimensional coral-like Ni2P-ACC nanostructure as binder-free electrode for greatly improved supercapacitor. Electrochim. Acta 2020, 349, 136259. [Google Scholar] [CrossRef]
  16. Wang, D.; Kong, L.B.; Liu, M.C.; Zhang, W.B.; Luo, Y.C.; Kang, L. Amorphous Ni–P materials for high performance pseudocapacitors. J. Power Sources 2015, 274, 1107–1113. [Google Scholar] [CrossRef]
  17. Gou, J. Ni2P/NiS2 Composite with Phase Boundaries as High-Performance Electrode Material for Supercapacitor. J. Electrochem. Soc. 2017, 164, A2956–A2961. [Google Scholar] [CrossRef]
  18. Zhang, L.; Zhang, W.B.; Zhang, Q.; Bao, X.; Ma, X.J. Electrochemical capacitance of intermetallic vanadium carbide. Intermetallics 2020, 127, 106976. [Google Scholar] [CrossRef]
  19. Zong, Q.; Yang, H.; Wang, Q.; Zhang, Q.; Xu, J.; Zhu, Y.; Wang, H.; Wang, H.; Zhang, F.; Shen, Q. NiCo2O4/NiCoP nanoflake-nanowire arrays: A homogeneous hetero-structure for high performance asymmetric hybrid supercapacitors. Dalton Trans. 2018, 47, 16320–16328. [Google Scholar] [CrossRef]
  20. Zhang, Q.; Zhang, W.B.; Hei, P.; Hou, Z.; Yang, T.; Long, J. CoP nanoprism arrays: Pseudocapacitive behavior on the electrode-electrolyte interface and electrochemical application as an anode material for supercapacitors. Appl. Surf. Sci. 2020, 527, 146682. [Google Scholar] [CrossRef]
  21. Elshahawy, A.M.; Guan, C.; Li, X.; Zhang, H.; Hu, Y.; Wu, H.; Pennycook, S.J.; Wang, J. Sulfur-doped cobalt phosphide nanotube arrays for highly stable hybrid supercapacitor. Nano Energy 2017, 39, 162–171. [Google Scholar] [CrossRef]
  22. Shang, M.; Zhang, J.; Liu, X.; Liu, Y.; Guo, S.; Yu, S.; Filatov, S.; Yi, X. N, S self-doped hollow-sphere porous carbon derived from puffball spores for high performance supercapacitors. Appl. Surf. Sci. 2021, 542, 148697. [Google Scholar] [CrossRef]
  23. Shao, Y.; Zhao, Y.; Li, H.; Xu, C. Three-Dimensional Hierarchical NixCo1-xO/NiyCo2-yP@C Hybrids on Nickel Foam for Excellent Supercapacitors. ACS Appl. Mater. Interfaces 2016, 8, 35368–35376. [Google Scholar] [CrossRef] [PubMed]
  24. Fleischmann, S.; Mitchell, J.B.; Wang, R.; Zhan, C.; Jiang, D.E.; Presser, V.; Augustyn, V. Pseudocapacitance: From Fundamental Understanding to High Power Energy Storage Materials. Chem Rev. 2020, 120, 6738–6782. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, S.; Huang, Z.; Li, R.; Zheng, X.; Lu, F.; He, T. Template-assisted synthesis of NiP@CoAl-LDH nanotube arrays with superior electrochemical performance for supercapacitors. Electrochim. Acta 2016, 204, 160–168. [Google Scholar] [CrossRef]
  26. Ruan, Y.; Wang, C.; Jiang, J. Nanostructured Ni compounds as electrode materials towards high-performance electrochemical capacitors. J. Mater. Chem. A. 2016, 4, 14509–14538. [Google Scholar] [CrossRef]
  27. Liu, Y.; Zhang, X.; Matras-Postolek, K.; Yang, P. Ni2P nanosheets modified N-doped hollow carbon spheres towards enhanced supercapacitor performance. J. Alloys Compd. 2021, 854, 157111. [Google Scholar] [CrossRef]
  28. Li, X.; Elshahawy, A.M.; Guan, C.; Wang, J. Metal Phosphides and Phosphates-based Electrodes for Electrochemical Supercapacitors. Small 2017, 13, 1701530. [Google Scholar] [CrossRef]
  29. Aziz, S.T.; Kumar, S.; Riyajuddin, S.; Ghosh, K.; Nessim, G.D.; Dubal, D.P. Bimetallic Phosphides for Hybrid Supercapacitors. J. Phys. Chem. Lett. 2021, 12, 5138–5149. [Google Scholar] [CrossRef]
  30. Zhang, W.B.; Zhang, Q.; Bao, X.; Ma, X.J.; Han, X.W.; Zhang, L.; Guo, Y.; Long, J.; Nian, C. Enhancing pseudocapacitive performance of CoP coating on nickel foam via surface Ni2P modification and Ni (II) doping for supercapacitor energy storage application. Surf. Coat. Technol. 2021, 421, 127469. [Google Scholar] [CrossRef]
  31. Zhang, Q.; Zhang, W.B.; Ma, X.J.; Zhang, L.; Bao, X.; Guo, Y.W.; Long, J. Boosting pseudocapacitive energy storage performance via both phosphorus vacancy defect and charge injection technique over the CoP electrode. J. Alloys Compd. 2021, 864, 158106. [Google Scholar] [CrossRef]
  32. Zhao, H.; Yuan, Z.Y. Transition metal–phosphorus-based materials for electrocatalytic energy conversion reactions. Catal. Sci. Technol. 2017, 7, 330–347. [Google Scholar] [CrossRef]
  33. Liu, M.C.; Hu, Y.M.; An, W.Y.; Hu, Y.X.; Niu, L.Y.; Kong, L.B.; Kang, L. Construction of high electrical conductive nickel phosphide alloys with controllable crystalline phase for advanced energy storage. Electrochim. Acta 2017, 232, 387–395. [Google Scholar] [CrossRef]
  34. Zhou, K.; Zhou, W.; Yang, L.; Lu, J.; Cheng, S.; Mai, W.; Tang, Z.; Li, L.; Chen, S. Ultrahigh-Performance Pseudocapacitor Electrodes Based on Transition Metal Phosphide Nanosheets Array via Phosphorization: A General and Effective Approach. Adv. Funct. Mater. 2015, 25, 7530–7538. [Google Scholar] [CrossRef]
  35. Ayom, G.E.; Khan, M.D.; Choi, J.; Gupta, R.K.; van Zyl, W.E.; Revaprasadu, N. Synergistically enhanced performance of transition-metal doped Ni2P for supercapacitance and overall water splitting. Dalton Trans 2021, 50, 11821–11833. [Google Scholar] [CrossRef]
  36. An, C.; Wang, Y.; Wang, Y.; Liu, G.; Li, L.; Qiu, F.; Xu, Y.; Jiao, L.; Yuan, H. Facile synthesis and superior supercapacitor performances of Ni2P/rGO nanoparticles. RSC Adv. 2013, 3, 4628–4633. [Google Scholar] [CrossRef]
  37. Dang, T.; Wei, D.; Zhang, G.; Wang, L.; Li, Q.; Liu, H.; Cao, Z.; Zhang, G.; Duan, H. Homologous NiCoP/CoP hetero-nanosheets supported on N-doped carbon nanotubes for high-rate hybrid supercapacitors. Electrochim. Acta 2020, 341, 135988. [Google Scholar] [CrossRef]
  38. Du, W.; Wei, S.; Zhou, K.; Guo, J.; Pang, H.; Qian, X. One-step synthesis and graphene-modification to achieve nickel phosphide nanoparticles with electrochemical properties suitable for supercapacitors. Mater. Res. Bull. 2015, 61, 333–339. [Google Scholar] [CrossRef]
  39. Lu, Y.; Liu, J.K.; Liu, X.Y.; Huang, S.; Wang, T.Q.; Wang, X.L.; Gu, C.D.; Tu, J.P.; Mao, S.X. Facile synthesis of Ni-coated Ni2P for supercapacitor applications. Cryst. Eng. Comm. 2013, 15, 7071–7079. [Google Scholar] [CrossRef]
  40. Hu, Y.M.; Liu, M.C.; Hu, Y.X.; Yang, Q.Q.; Kong, L.B.; Han, W.; Li, J.J.; Kang, L. Design and synthesis of Ni2P/Co3V2O8 nanocomposite with enhanced electrochemical capacitive properties. Electrochim. Acta 2016, 190, 1041–1049. [Google Scholar] [CrossRef]
  41. Jia, H.; Li, Q.; Li, C.; Song, Y.; Zheng, H.; Zhao, J.; Zhang, W.; Liu, X.; Liu, Z.; Liu, Y. A novel three-dimensional hierarchical NiCo2O4/Ni2P electrode for high energy asymmetric supercapacitor. Chem. Eng. J. 2018, 354, 254–260. [Google Scholar] [CrossRef]
  42. Adán-Más, A.; Silva, T.M.; Guerlou-Demourgues, L.; Bourgeois, L.; Labrugere-Sarroste, C.; Montemor, M.F. Nickel-cobalt oxide modified with reduced graphene oxide: Performance and degradation for energy storage applications. J. Power Sources 2019, 419, 12–26. [Google Scholar] [CrossRef]
  43. Askari, M.B.; Salarizadeh, P.; Seifi, M.; Ramezan zadeh, M.H.; Di Bartolomeo, A. ZnFe2O4 nanorods on reduced graphene oxide as advanced supercapacitor electrodes. J. Alloys Compd. 2021, 860, 158497. [Google Scholar] [CrossRef]
  44. Qin, Y.; Lyu, Y.; Chen, M.; Lu, Y.; Qi, P.; Wu, H.; Sheng, Z.; Gan, X.; Chen, Z.; Tang, Y. Nitrogen-doped Ni2P/Ni12P5/Ni3S2 three-phase heterostructure arrays with ultrahigh areal capacitance for high-performance asymmetric supercapacitor. Electrochim. Acta 2021, 393, 139059. [Google Scholar] [CrossRef]
  45. Salarizadeh, P.; Askari, M.B. MoS2–ReS2/rGO: A novel ternary hybrid nanostructure as a pseudocapacitive energy storage material. J. Alloys Compd. 2021, 874, 159886. [Google Scholar] [CrossRef]
  46. Oyedotun, K.O.; Mirghni, A.A.; Fasakin, O.; Tarimo, D.J.; Kitenge, V.N.; Manyala, N. High-Energy Asymmetric Supercapacitor Based on the Nickel Cobalt Oxide (NiCo2O4) Nanostructure Material and Activated Carbon Derived from Cocoa Pods. Energy Fuels 2021, 35, 20309–20319. [Google Scholar] [CrossRef]
  47. Liang, R.; Shu, C.; Hu, A.; Li, M.; Ran, Z.; Zheng, R.; Long, J. Interface engineering induced selenide lattice distortion boosting catalytic activity of heterogeneous CoSe2@NiSe2 for lithium-oxygen battery. Chem. Eng. J. 2020, 393, 124592. [Google Scholar] [CrossRef]
  48. Hu, A.; Lv, W.; Lei, T.; Chen, W.; Hu, Y.; Shu, C.; Wang, X.; Xue, L.; Huang, J.; Du, X.; et al. Heterostructured NiS2/ZnIn2S4 Realizing Toroid-like Li2O2 Deposition in Lithium–Oxygen Batteries with Low-Donor-Number Solvents. ACS Nano 2020, 14, 3490–3499. [Google Scholar] [CrossRef]
  49. Liang, L.L.; Song, G.; Liu, Z.; Chen, J.P.; Xie, L.J.; Jia, H.; Kong, Q.Q.; Sun, G.H.; Chen, C.M. Constructing Ni12P5/Ni2P Heterostructures to Boost Interfacial Polarization for Enhanced Microwave Absorption Performance. ACS Appl. Mater. Interfaces 2020, 12, 52208–52220. [Google Scholar] [CrossRef]
  50. An, L.; Li, Y.; Luo, M.; Yin, J.; Zhao, Y.Q.; Xu, C.; Cheng, F.; Yang, Y.; Xi, P.; Guo, S. Atomic-Level Coupled Interfaces and Lattice Distortion on CuS/NiS2.Nanocrystals Boost Oxygen Catalysis for Flexible Zn-Air Batteries. Adv. Funct. Mater. 2017, 27, 1703779. [Google Scholar] [CrossRef]
  51. Liu, Y.; Hua, X.; Xiao, C.; Zhou, T.; Huang, P.; Guo, Z.; Pan, B.; Xie, Y. Heterogeneous Spin States in Ultrathin Nanosheets Induce Subtle Lattice Distortion to Trigger Efficient Hydrogen Evolution. J. Am. Chem. Soc. 2016, 138, 5087–5092. [Google Scholar] [CrossRef] [PubMed]
  52. Sun, Z.; Wang, Y.; Zhang, L.; Wu, H.; Jin, Y.; Li, Y.; Shi, Y.; Zhu, T.; Mao, H.; Liu, J.; et al. Simultaneously Realizing Rapid Electron Transfer and Mass Transport in Jellyfish-Like Mott–Schottky Nanoreactors for Oxygen Reduction Reaction. Adv. Funct. Mater. 2020, 30, 1910482. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of the samples in various Ni/P ratios. (ad) XRD pattern of Ni/P ratio 3:1, 5:2, 12:5, and 7:3. (eh) XRD pattern of Ni/P ratio 2:1, 5:4, 1:1 and single-phase Ni2P.
Figure 1. X-ray diffraction patterns of the samples in various Ni/P ratios. (ad) XRD pattern of Ni/P ratio 3:1, 5:2, 12:5, and 7:3. (eh) XRD pattern of Ni/P ratio 2:1, 5:4, 1:1 and single-phase Ni2P.
Crystals 12 00469 g001
Figure 2. SEM images of various Ni/P ratios and single-phase Ni2P. (ag) SEM images of Ni/P ratio 3:1, 5:2, 12:5, 7:3, 2:1, 5:4, and 1:1. (h) SEM images of single-phase Ni2P.
Figure 2. SEM images of various Ni/P ratios and single-phase Ni2P. (ag) SEM images of Ni/P ratio 3:1, 5:2, 12:5, 7:3, 2:1, 5:4, and 1:1. (h) SEM images of single-phase Ni2P.
Crystals 12 00469 g002
Figure 3. TEM images of the prepared sample. (ac) The TEM, HRTEM, and SAED images of Ni2P/Ni3P/Ni (Ni/P = 7:3). (df) The TEM, HRTEM, and SAED images of Ni2P/Ni5P4 (Ni/P = 5:4).
Figure 3. TEM images of the prepared sample. (ac) The TEM, HRTEM, and SAED images of Ni2P/Ni3P/Ni (Ni/P = 7:3). (df) The TEM, HRTEM, and SAED images of Ni2P/Ni5P4 (Ni/P = 5:4).
Crystals 12 00469 g003
Figure 4. The BET test and pore size distribution information. (a,c) The BET test information of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4). (b,d) The pore size distribution information of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4).
Figure 4. The BET test and pore size distribution information. (a,c) The BET test information of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4). (b,d) The pore size distribution information of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4).
Crystals 12 00469 g004
Figure 5. The XPS test and analysis of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4). (a) The XPS spectrum of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4) on the Ni 2p orbital. (b) The XPS spectra of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4) on the P 2p orbital.
Figure 5. The XPS test and analysis of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4). (a) The XPS spectrum of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4) on the Ni 2p orbital. (b) The XPS spectra of single-phase Ni2P and Ni2P/Ni5P4 (Ni/P = 5:4) on the P 2p orbital.
Crystals 12 00469 g005
Figure 6. The electrochemical test graphs of various Ni/P ratios. (a) Galvanostatic charge-discharge curves of electrodes prepared with various Ni/P ratios at 1 A g−1. (b) Cyclic voltammetry curves of electrode prepared with various Ni/P ratios at a sweep rate of 3 mV s−1. (c,d) Nyquist plots of electrodes prepared by various Ni/P ratios. (e) Bode plots of electrodes prepared with various Ni/P ratios.
Figure 6. The electrochemical test graphs of various Ni/P ratios. (a) Galvanostatic charge-discharge curves of electrodes prepared with various Ni/P ratios at 1 A g−1. (b) Cyclic voltammetry curves of electrode prepared with various Ni/P ratios at a sweep rate of 3 mV s−1. (c,d) Nyquist plots of electrodes prepared by various Ni/P ratios. (e) Bode plots of electrodes prepared with various Ni/P ratios.
Crystals 12 00469 g006
Figure 7. The electrochemical test graph of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P. (a) Galvanostatic charge-discharge curves of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P at 1 A g−1. (b) Cyclic voltammetry curves of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P at a sweep rate of 3 mV s−1. (c) Nyquist plots of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P. (d) Bode plots of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4) and single-phase Ni2P.
Figure 7. The electrochemical test graph of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P. (a) Galvanostatic charge-discharge curves of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P at 1 A g−1. (b) Cyclic voltammetry curves of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P at a sweep rate of 3 mV s−1. (c) Nyquist plots of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase Ni2P. (d) Bode plots of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4) and single-phase Ni2P.
Crystals 12 00469 g007
Figure 8. The electrochemical test graph of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4). (a,b) Galvanostatic charge-discharge curves of Ni2P/Ni3P/Ni (Ni/P = 7:3) under various current densities and cyclic voltammetry curves of Ni2P/Ni3P/Ni (Ni/P = 7:3) at various scan rates. (d,e) Galvanostatic charge-discharge curves Ni2P/Ni5P4 (Ni/P = 5:4) under various current densities and cyclic voltammetry curves of Ni2P/Ni5P4 (Ni/P = 5:4) at various scan rates. (c,f) Cyclic stability test graph of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4).
Figure 8. The electrochemical test graph of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4). (a,b) Galvanostatic charge-discharge curves of Ni2P/Ni3P/Ni (Ni/P = 7:3) under various current densities and cyclic voltammetry curves of Ni2P/Ni3P/Ni (Ni/P = 7:3) at various scan rates. (d,e) Galvanostatic charge-discharge curves Ni2P/Ni5P4 (Ni/P = 5:4) under various current densities and cyclic voltammetry curves of Ni2P/Ni5P4 (Ni/P = 5:4) at various scan rates. (c,f) Cyclic stability test graph of Ni2P/Ni3P/Ni (Ni/P = 7:3) and Ni2P/Ni5P4 (Ni/P = 5:4).
Crystals 12 00469 g008
Figure 9. The electrochemical test graph of the assembled supercapacitors. (a) Galvanostatic charge-discharge curves of Ni2P/Ni3P/Ni//AC and Ni2P/Ni5P4//AC supercapacitors at a current density of 1 A g−1. (b) Cyclic voltammetry curves of Ni2P/Ni3P/Ni//AC and Ni2P/Ni5P4//AC supercapacitors at a scan rate of 100 mV s−1. (c,d) Galvanostatic charge-discharge curves of Ni2P/Ni5P4//AC supercapacitor under various current densities and cyclic voltammetry curves of Ni2P/Ni5P4//AC supercapacitor at various scan rates.
Figure 9. The electrochemical test graph of the assembled supercapacitors. (a) Galvanostatic charge-discharge curves of Ni2P/Ni3P/Ni//AC and Ni2P/Ni5P4//AC supercapacitors at a current density of 1 A g−1. (b) Cyclic voltammetry curves of Ni2P/Ni3P/Ni//AC and Ni2P/Ni5P4//AC supercapacitors at a scan rate of 100 mV s−1. (c,d) Galvanostatic charge-discharge curves of Ni2P/Ni5P4//AC supercapacitor under various current densities and cyclic voltammetry curves of Ni2P/Ni5P4//AC supercapacitor at various scan rates.
Crystals 12 00469 g009
Figure 10. Linear sweep volt–ampere curve, schematic diagram of heterogeneous interface dangling bonds (unpaired electronics) generation, energy band diagrams of metal and semiconductors. (a) The Linear sweep volt–ampere curve of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single phase Ni2P. (b) The schematic diagram of dangling bonds of heterogeneous interface. (c,d) The energy band diagrams of metal and semiconductors.
Figure 10. Linear sweep volt–ampere curve, schematic diagram of heterogeneous interface dangling bonds (unpaired electronics) generation, energy band diagrams of metal and semiconductors. (a) The Linear sweep volt–ampere curve of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single phase Ni2P. (b) The schematic diagram of dangling bonds of heterogeneous interface. (c,d) The energy band diagrams of metal and semiconductors.
Crystals 12 00469 g010
Table 1. The electrochemical performance of composites prepared with various raw material ratios.
Table 1. The electrochemical performance of composites prepared with various raw material ratios.
Ratio of Raw (Ni:P)Prepared CompoundsSpecific Capacity (mAh g−1)Rate Capability (%)
3:1Ni3P/Ni2P/Ni19530
5:2Ni2P/Ni5P414373
12:5Ni5P4/Ni2P/Ni7369
7:3Ni2P/Ni3P/Ni32159
2:1Ni5P4/NiP2/Ni10869
5:4Ni2P/Ni5P421876
1:1Ni5P4/Ni2P11953
Table 2. The electrochemical performance of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase nickel phosphide.
Table 2. The electrochemical performance of Ni2P/Ni3P/Ni (Ni/P = 7:3), Ni2P/Ni5P4 (Ni/P = 5:4), and single-phase nickel phosphide.
Electrode MaterialsSpecific Capacity (mAh g−1)Rate Capability (%)ResRctTime Constant(s)
Ni2P/Ni3P/Ni
(Ni/P = 7:3)
321590.8380.0560.87
Ni2P/Ni5P4
(Ni/P = 5:4)
218761.4730.1114.01
Ni2P58330.8620.17843
Table 3. Comparison of the electrochemical performance of our prepared electrodes with other reported electrode materials in the literature.
Table 3. Comparison of the electrochemical performance of our prepared electrodes with other reported electrode materials in the literature.
ElectrodeSubstrateElectrolyteCurrent Density Specific CapacityRefs.
NixCo1-xOy/Er-GoStainless steel1 M KOH1 A g−1180 mAh g−1[42]
CoP/Ni2PNF6 M KOH1 A g−11557 C g−1[30]
Ni-CoPNF6 M KOH1 A g−1578 C g−1[30]
NiCo2O4/Ni2PNF3 M KOH8 mA cm−22900 F g−1[41]
N-Ni2P/Ni12P5/Ni3S2NF2 M KOH20 mA cm−212.71 F cm−2[44]
Ni2P@N-CNF3 M KOH10 A g−11320.4 F g−1[27]
Co-Ni2PNF3 M KOH1 A g−1864 F g−1[35]
Fe-Ni2PNF3 M KOH1 A g−1856 F g−1[35]
NiCoP/CoPNF2 M KOH1 A g−1152 mAh g−1[37]
Ni2P/Ni3P/NiNF2 M KOH1 A g−1321 mAh g−1This work
Ni2P/Ni5P4NF2 M KOH1 A g−1218 mAh g−1This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, S.-B.; Zhang, W.-B.; Yang, Z.-Q.; Bao, X.; Zhang, L.; Guo, Y.-W.; Han, X.-W.; Long, J. The Preparation and Electrochemical Pseudocapacitive Performance of Mutual Nickel Phosphide Heterostructures. Crystals 2022, 12, 469. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12040469

AMA Style

Guo S-B, Zhang W-B, Yang Z-Q, Bao X, Zhang L, Guo Y-W, Han X-W, Long J. The Preparation and Electrochemical Pseudocapacitive Performance of Mutual Nickel Phosphide Heterostructures. Crystals. 2022; 12(4):469. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12040469

Chicago/Turabian Style

Guo, Shao-Bo, Wei-Bin Zhang, Ze-Qin Yang, Xu Bao, Lun Zhang, Yao-Wen Guo, Xiong-Wei Han, and Jianping Long. 2022. "The Preparation and Electrochemical Pseudocapacitive Performance of Mutual Nickel Phosphide Heterostructures" Crystals 12, no. 4: 469. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12040469

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