Next Article in Journal
Substantial Improvement of Color-Rendering Properties of Conventional White LEDs Using Remote-Type Red Quantum-Dot Caps
Next Article in Special Issue
Research Progress on Sound Absorption of Electrospun Fibrous Composite Materials
Previous Article in Journal
Formulation and In Vitro Efficacy Assessment of Teucrium marum Extract Loading Hyalurosomes Enriched with Tween 80 and Glycerol
Previous Article in Special Issue
Deformation Mechanism of Depositing Amorphous Cu-Ta Alloy Film via Nanoindentation Test
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synchronous Defect and Interface Engineering of NiMoO4 Nanowire Arrays for High-Performance Supercapacitors

1
Yunnan Key Laboratory of Optoelectronic Information Technology, College of Physics and Electronics Information, Yunnan Normal University, Kunming 650500, China
2
Key Laboratory of Advanced Technique & Preparation for Renewable Energy Materials, Ministry of Education, Yunnan Normal University, Kunming 650500, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(7), 1094; https://doi.org/10.3390/nano12071094
Submission received: 1 March 2022 / Revised: 18 March 2022 / Accepted: 24 March 2022 / Published: 26 March 2022
(This article belongs to the Special Issue Nanotechnologies and Nanomaterials: Selected Papers from CCMR)

Abstract

:
Developing high-performance electrode materials is in high demand for the development of supercapacitors. Herein, defect and interface engineering has been simultaneously realized in NiMoO4 nanowire arrays (NWAs) using a simple sucrose coating followed by an annealing process. The resultant hierarchical oxygen-deficient NiMoO4@C NWAs (denoted as “NiMoO4−x@C”) are grown directly on conductive ferronickel foam substrates. This composite affords direct electrical contact with the substrates and directional electron transport, as well as short ionic diffusion pathways. Furthermore, the coating of the amorphous carbon shell and the introduction of oxygen vacancies effectively enhance the electrical conductivity of NiMoO4. In addition, the coated carbon layer improves the structural stability of the NiMoO4 in the whole charging and discharging process, significantly enhancing the cycling stability of the electrode. Consequently, the NiMoO4−x@C electrode delivers a high areal capacitance of 2.24 F cm−2 (1720 F g−1) at a current density of 1 mA cm−2 and superior cycling stability of 84.5% retention after 6000 cycles at 20 mA cm−2. Furthermore, an asymmetric super-capacitor device (ASC) has been constructed with NiMoO4−x@C as the positive electrode and activated carbon (AC) as the negative electrode. The as-assembled ASC device shows excellent electrochemical performance with a high energy density of 51.6 W h kg−1 at a power density of 203.95 W kg−1. Moreover, the NiMoO4//AC ASC device manifests remarkable cyclability with 84.5% of capacitance retention over 6000 cycles. The results demonstrate that the NiMoO4−x@C composite is a promising material for electrochemical energy storage. This work can give new insights on the design and development of novel functional electrode materials via defect and interface engineering through simple yet effective chemical routes.

1. Introduction

Several alternative energy technologies have been under development globally in a great effort to mitigate the energy and environmental challenges faced and in accordance with the current “carbon neutral” policies. Supercapacitors (SCs), also known as electrochemical capacitors (ECs), have been considered as one of the most promising energy storage devices due to their unique characteristics of high power density (>10 kW/kg), fast charging and discharging capability (within a few seconds), long lifespan (over 100,000 cycles), and good operational safety. SCs have been widely applied in some important fields, including smart electric grids, memory back-ups, (hybrid) electric vehicles, and aerospace crafts. Although SCs have the advantages of high power density and very long calendar lives, their further application is still hindered by their limited energy density. Therefore, it is crucial to develop advanced high-performance SCs with higher energy densities without severely compromising the power density and cyclability.
Recently, asymmetric supercapacitors (ASCs) have been regarded as promising due to their enhanced energy density. ASCs commonly combine pseudocapacitive materials (as positive electrode) and electric double-layer capacitive materials (as negative electrode), making use of the much higher specific capacitance derived from the pseudocapacitive electrode materials and a wider potential window during operation. In this sense, the electrochemical performances of ASCs are largely influenced by the structures and properties of the pseudocapacitive electrode materials. Thus, the development of high energy density ASCs heavily relies on the rational selection and design, as well as delicate fabrication of advanced electrode materials.
The use of transition metal oxides (TMOs) [1], hydroxides [2], sulfides [3], selenides [4], carbides [5], and their composites with conductive carbon and polymers [6], as possible electrode materials for ASCs, has been attempted. Particularly, TMOs, such as NiO [7], MoO3 [8], Co3O4 [9], and MnO2 [10], have received considerable attention due to their low cost, convenience of synthesis, environmental friendliness, and relatively high capacitance. Compared to binary TMOs, ternary TMOs materials containing two different metal cations, such as NiCo2O4 [11], ZnCo2O4 [12,13], CoMoO4 [14], MnMoO4 [15], and NiMoO4 [16], as well as some solid solutions [17], exhibit higher electrochemical activity due to the rich redox reactions stemming from the multiple oxidation states of the transition-metal components. It has been proposed that NiMoO4 has a good reversible capacitance and electrochemical characteristics for SCs/ASCs applications because of the electrochemical active Ni cation and improved electronic conductivity from Mo cation [16,18].
However, the practical application of NiMoO4 electrodes in SCs and ASCs is still hampered by their poor electronic conductivity, insufficient ionic transport and diffusion, and structural instability during long-term cycling [19]. Therefore, the controllable fabrication of NiMoO4 with desired nano- and microstructures and rational structural engineering is highly desired but remains challenging.
Various nanostructured NiMoO4 materials, such as nanosheets and nanorods arrays [20], nanotubes [21], hollow nanorods [22], mesoporous nanospheres [23], nanoparticles, and quantum dots [24], have been designed to boost the electrochemical performances of the NiMoO4 electrodes via increased exposed surface for ion adsorption and insertion, shortened path distances for ion transport and diffusion, and improved electrolyte impregnation and permeation. Specifically, various low-dimensional NiMoO4 nanostructures directly grown on conductive substrates (e.g., Ni/Cu foams [25,26], graphene [27], and carbon substrates [28,29]) are particularly preferred for directional electron transport with reduced charge carrier scattering at grain boundaries and easy integration into flexible devices with some specific applications.
To overcome the poor electronic conductivity of pristine NiMoO4, various NiMoO4/carbon composites have been synthesized by hybridizing NiMoO4 nanostructures with graphene [30,31], carbon nanotubes [32], conducting polymers [33], and porous carbon architectures [34,35] Alternatively, intentional doping of NiMoO4 with several kinds of heteroatoms such as Mn [36,37], P [38], Zn [39], Ce [40], or the creation of oxygen vacancies [14,41,42,43] in the lattice have recently been reported. In addition, NiMoO4 has also been coupled with other metal oxides [44,45,46,47,48,49] or sulfides [50,51,52] to form heterostructure electrodes for supercapacitors with improved electrochemical performances.
However, monotonous strategy sometimes has a limited contribution for the overall electrochemical performance improvement of NiMoO4 materials. In addition, some reported approaches for hybridization or doping of NiMoO4 involve multiple and complex chemical and physical processes that are not economically or environmentally friendly. Thus, the rational design and the design of a NiMoO4-based composite electrode for high-performance supercapacitors remains a challenge.
In this work, we report the simultaneous defect and interface engineering of NiMoO4 nanowires arrays (NWAs) using a simple and effective sucrose coating followed by a thermal treatment approach. In this process, an amorphous carbon shell was uniformly coated on the NiMoO4 surface, effectively improving the electronic transport and structural integrity of the NiMoO4 during electrochemical cycling. Additionally, oxygen-vacancy defects were incorporated into the NiMoO4 during the carbonization process, further enhancing the electronic conductivity of NiMoO4 and redox activity in the NiMoO4 electrode surface. As expected, the resultant NiMoO4−x@C composite exhibited a higher specific capacitance than that of the pristine NiMoO4 NWAs. Furthermore, an asymmetric supercapacitor (ASC) was assembled with the NiMoO4−x@C as positive electrode and activated carbon (AC) as negative electrode, delivering a remarkably high energy density of 51.6 W h kg−1 at a power density of 203 W kg−1 and an excellent cycling stability with a retention of 84.5% after 6000 cycles under a high current density of 10 A g−1.

2. Results and Discussion

The synthesis route of oxygen-deficient NiMoO4@carbon nanowire arrays (NiMoO4−x@C) is schematically shown in Figure 1. The preparation process mainly involved three critical steps. Firstly, a NiMoO4 nanowire arrays (NWAs) precursor (NiMoO4·xH2O NWAs, light green) was directly deposited on a ferronickel foam by a hydrothermal reaction process (Step 1, Figure 1). Secondly, the NiMoO4 NWAs precursor was transferred into NiMoO4 NWAs by annealing in Ar to remove crystal H2O and improve crystallinity (Step 2, Figure 1). Finally, the as-obtained NiMoO4 NWAs were immersed in a sucrose solution, followed by drying and annealing in an Ar atmosphere to fabricate oxygen-deficient NiMoO4@carbon NWAs (NiMoO4−x@C) (Step 3, Figure 1).
The crystal structures of NiMoO4−x@C and neat NiMoO4 NWAs samples were characterized by X-ray diffraction (XRD) analysis as depicted in Figure 2. The two strongest diffraction peaks, located at ca. 45° and 52°, were from the ferronickel foam substrate. The NiMoO4−x@C and NiMoO4 NWAs samples showed similar diffraction peak shapes and locations. The diffraction peaks located at 14.3°, 24.0°, 25.4°, 28.9°, 32.6°, 37.1°, 38.7°, 41.3°, and 47.4° corresponded to the (110), (021), (−112), (220), (022), (−113), (−132), (040), and (−204) crystal planes of orthorhombic NiMoO4 (JCPDS card No. 86-0361) [37]. Compared to the NiMoO4 NWAs samples, the NiMoO4−x@C sample exhibited a slightly lower diffraction peak intensity possibly due to the covering of carbon on the NiMoO4 surface as well as the reduced crystallinity of NiMoO4 with increased structural defects. In addition, no characteristic peaks for carbon phases were noted, suggesting the amorphous nature of the carbon species in the NiMoO4−x@C sample. The formation of amorphous carbon could be partially ascribed to the lower low annealing temperature (400 °C herein).
Raman spectra of pristine NiMoO4 NWAs and NiMoO4−x@C samples are illustrated in Figure 3. The bands at 961 cm−1 and 913 cm−1 corresponded to the symmetric and asymmetric stretching modes of Mo=O bonds, while the band at 706 cm−1 could be ascribed to the stretching mode of Ni/Mo-O bonds of the orthorhombic α-NiMoO4 phase [53]. In addition, two bands ascribed to the presence of carbon species were identified. The band at around 1360 cm−1 could be attributed to the D band from defects and disorders in the amorphous carbon layers, while the other band at around 1590 cm−1 was related to the G band related to the vibration of sp2-bonded carbon atoms [54]. This result implied the successful deposition of amorphous carbon layer on the surface of the NiMoO4 NWAs.
The morphologies of the NiMoO4 samples were firstly observed using scanning electron microscopy (SEM). From Figure 4a, the pristine NiMoO4 NWAs sample was composed of oriented nanowires (NWs) with a smooth surface. In addition, the NiMoO4 NWAs have relatively uniform diameters of ~300 nm, on average, and lengths of several micrometers. After the coating of the carbon, the surface of the NiMoO4−x@C sample became obviously coarse as shown in Figure 4b. The element composition analyses using energy-dispersive X-ray spectra (EDS) analysis indicated the existence of Ni, Mo, O, C, Fe, and Al elements in the NiMoO4−x@C sample (Figure S1, Supporting Information). Note that the Fe and Al signals mainly stemmed from the ferronickel foam substrate and the sample holder, respectively.
The microstructures of the pure NiMoO4 NWAs and NiMoO4−x@C samples were further investigated by transmission electron microscope (TEM) and high-resolution TEM (HRTEM) as shown in Figure 5. The TEM image (Figure 5a) revealed that the pure NiMoO4 nanowire had a smooth surface with a diameter of about 200 nm. A selected-area electron diffraction (SAED) pattern (inset of Figure 5a) taken from this nanowire depicted a clear two-dimensional dot pattern, suggesting its single-crystalline structure in nature. Two diffraction spots, as marked by white circles, could be indexed to the (220) and (−222) crystal facets of orthorhombic NiMoO4. From Figure 5b, the crystal plane with a lattice spacing of 2.73 Å in the HRTEM micrograph corresponded to the (−222) planes of NiMoO4 [49,52]. In contrast, the TEM image in Figure 5c indicated that a layer of amorphous carbon film with a thickness of ca. 20~50 nm had been coated on the NiMoO4 nanowire’s surface, confirming the core-shell structure of the NiMoO4−x@C composite sample with different brightness contrasts of NiMoO4 and carbon. The deposition of amorphous carbon on the surface of NiMoO4 can be further confirmed by HRTEM micrograph as shown in Figure 5d.
Next, the chemical composition and valence states of element on the surface of NiMoO4 NWAs and NiMoO4−x@C samples were identified by X-ray photoelectron spectroscopy (XPS, Figure 6). In the high-resolution Ni 2p spectrum of the pristine NiMoO4 NWAs sample (Figure 6a), two main peaks were observed at binding energies (BEs) of 873.4 eV and 856.3 eV with a spin-orbital splitting energy of 17.1 eV, corresponding to the Ni 2p1/2 and Ni 2p3/2 of Ni2+ in NiMoO4 lattice [55]. In addition, two satellite peaks with BEs of 877.3 eV and 860.6 eV were noted for Ni2+. In Figure 6b, similar peak locations and separations can also be observed in the Ni 2p spectra of NiMoO4−x@C sample, indicating the coating of carbon had little effect on the chemical valence states of the Ni component in NiMoO4. The Mo 6d spectrum of pure NiMoO4 and NiMoO4−x@C samples are shown in Figure 6c,d. Evidently both samples had two strong bands with BEs located at 236.0 eV and 232.9 eV, which could be assigned to Mo 3d3/2 and Mo 3d5/2 of Mo6+ cations in the NiMoO4 lattice [56]. In addition, another pair of doublets was noted for the NiMoO4−x@C sample, verifying the existence of Mo4+ in the NiMoO4−x@C composite [41,42,57] possibly produced during the amorphous carbon coating process. From Figure 6e, the O 1s spectrum of the NiMoO4 NWs sample was deconvoluted into three bands. The band centered at 530.1 eV was assigned to the lattice oxygen with O-Ni/O-Mo bonds, while the bands located at 531.3 eV and 532.9 eV correspond to the O-C and O=C bonds from moisture adsorbed on surface [58]. For the NiMoO4−x@C sample, another band could be noted at 532 eV, suggesting the presence of oxygen vacancies (Vo) [59] at the NiMoO4 surface (Figure 6f). The formation of and Mo4+ and Vo can tune the electronic structures and electrochemical properties of the NiMoO4−x@C composite sample.
Then, the effects of pyrolysis temperatures (from 200~800 °C) during the carbon coating of the morphologies and microstructures of the NiMoO4/carbon composites were investigated. It is noted that some aggregates of residual sucrose were observed after annealing at 200 °C (Figure S2a), suggesting the carbonization of sucrose was incomplete under a lower temperature. This SEM result also coincides well with the thermogravimetric (TGA) and the differential scanning calorimetry (DSC) analyses (Figure S3), where the thermal decomposition process of sucrose mainly occurs between 223 and 389 °C. With the increase of annealing temperature, sucrose was decomposed, and the carbonization process occurred accompanied by the release of some gases (e.g., CO, CO2). At a higher temperature, the generated reductive gases (e.g., CO) reacted with NiMoO4 and generated some oxygen vacancies on the NiMoO4 surface via abstracting some surface oxygen atoms. In contrast, well-defined nanowires were obtained for the samples prepared after annealing at 400 and 600 °C, respectively (Figure S2b,c). However, the nanowire structure was destroyed when the pyrolysis temperature was increased to 800 °C (Figure S2d), which might have been caused by the large inner stain in the NiMoO4 NWAs or at the NiMoO4−x@C interface. Thus, the standard annealing temperature was chosen as 400 °C.
To evaluate the electrochemical performance of NiMoO4 NWAs and NiMoO4−x@C samples, electrochemical measurements were tested by a three-electrode system with 2 M KOH electrolyte (Figure 7). Figure 7a shows the CV curves of NiMoO4 NWAs and NiMoO4−x@C samples at a scan rate of 20 mV s−1 with a potential window of 0 to 0.7 V. Overall, the NiMoO4−x@C sample had a larger integral area than that of the NiMoO4 NWAs sample, indicating a significant increase of capacitance after carbon deposition and introduction of oxygen vacancies. Meanwhile, the CV curves of the two samples exhibited typical oxidation peaks, demonstrating typical pseudocapacitive charge storage characteristics. In addition, the CV curves of four different pyrolysis temperatures of NiMoO4−x@C samples are revealed in Figure S4. The sample collected at 400 °C shows the highest capacitance which is consistent with the result of SEM in Figure S2. The sample collected at 800 °C exhibited an unsatisfactory performance due to its collapsed morphology. Figure 7b shows the GCD curves of the two samples. It revealed that the discharge time of the NiMoO4−x@C sample was almost twice as much as that of the NiMoO4 NWAs sample at a current density of 1 A cm−2. Figure 7c shows the electrochemical impedance spectroscopy (EIS) of NiMoO4 NWAs and NiMoO4−x@C. The direct impedance and charge transfer resistance of the NiMoO4−x@C sample was significantly lower than that of the NiMoO4 NWAs sample. The remarkably reduced size of the semicircle for the NiMoO4−x@C indicated an improved charge transfer kinetics due to enhanced electrical conductivity provided by the carbon shell and oxygen vacancy defects. In addition, the NiMoO4−x@C exhibits the steepest slope in the low-frequency region, clearly indicating the lowest Warburg impedance and, hence, the highest K-ion diffusion capability at the interface between the electrode and electrolyte. Through the AC EIS, we added the corresponding equivalent circuit diagram in Figure 7c.
The true impedance of capacitor can be estimated using the following Equation (1):
Z r e a l = R Ω + j ( w C ) ( R c t + W · j w 1 / 2 ) R c t + W · j w 1 / 2 j ( w C )
Figure 7d shows the capacitance of NiMoO4 NWAs and NiMoO4−x@C samples calculated from different current densities. After coating the carbon layer, the capacitance of the NiMoO4−x@C sample was greatly increased. The areal capacitance can be calculated as high as 2.24 F cm−2 (1720 F g−1) at a current density of 1 mA cm−2. In contrast, the NiMoO4 NWAs electrode only demonstrated a specific capacitance of 1.206 F cm−2 (927 F g−1) at the same current density. The cycling performances of NiMoO4 NWAs and NiMoO4−x@C samples are presented in Figure 7e. The capacitance retention of NiMoO4−x@C is 84.5% at 20 mA cm−2 after 6000 cycles, which is considerably better than that of the NiMoO4 NWAs sample (63.1% after 6000 cycles). To illustrate the difference of the cycling process between NiMoO4 NWAs and NiMoO4−x@C samples, we also obtained the SEM results after cycling as shown in Figure S5. It is evident that the NiMoO4−x@C sample still held some nanorod structures under the protection of amorphous carbon shell. Instead, NiMoO4 NWAs were aggregated after 10,000 cycles, with unsatisfactory cycle abilities. In addition, we made a comparison of the Cs and cycling stability of this work with some previously reported NiMoO4-based electrodes materials as summarized in Table S1.
First-principles density functional theory (DFT) simulations were next adopted to further probe the structure–performance relationship of the NiMoO4−x@C composite electrode in supercapacitors. The optimized geometry configurations of pristine and oxygen-deficient NiMoO4 (110) surface slabs are shown in Figure S6. The pristine NiMoO4 (110) surface was flat and composed of fivefold Ni and Mo atoms and twofold O atoms (Figure S6a). The defective NiMoO4 (110) surface can be produced after eliminating one surface O atom, leaving one threefold Ni and Mo atoms nearby (Figure S6b). The resultant oxygen-deficient NiMoO4 (110) plane retains flat. Next, the adsorption behavior of OH group on the pristine NiMoO4 (110) surface was first investigated. As shown in Figure 8, the OH can be adsorbed on the top of either the surface of the Mo atom (Figure 8a,b) or the Ni atom (Figure 8c,d), yielding an adsorption energy (Eads) of −0.50 and −5.97 eV, respectively. Evidently, the adsorption of OH on the Ni site was much stronger than that on the Mo site, which is consistent with the fact that the Ni in NiMoO4 is electrochemically active for pseudocapacitive charge storage process based on Faradic reactions. The chemisorption of OH on Ni and Mo sites of the NiMoO4 (110) surface was further verified by the interfacial charge transfer from the charge density difference contours (Figure S7).
In the following, the adsorption of OH adsorbed on the oxygen-deficient NiMoO4 (110) surface was evaluated. Specifically, the adsorption on the Ni and Mo sites with lower coordination due to the removal of surface O was considered. Interestingly, it is noted that the OH group was preferred to be adsorbed at the vicinity of the oxygen-vacancy position (Figure 9a), leading to an Eads of −3.63 eV and a charge transfer at the OH/NiMoO4 interface (Figure 9b). This result suggests that the presence of surface oxygen vacancies offers more active sites for OH adsorption, concentration, and subsequent redox reactions for enhanced pseudocapacitive charge storage.
Based on the above experimental data and theoretical simulations, the significantly improved electrochemical performances of the NiMoO4−x@C composite can be mainly attributed to the following points: (i) the deposition of amorphous carbon shell effectively enhances the electron transport of NiMoO4 nanowires and charge transfer at NiMoO4/C heterointerface; (ii) the deposited carbon layer also improves the structural integrity of the NiMoO4 nanowire arrays during long-term electrochemical cycling; (iii) the creation of oxygen vacancies in NiMoO4 accompanied by the coating of the carbon further enhances the electronic conductivity of the NiMoO4 electrode and creates more active sites for pseudocapacitive charge storage. Therefore, the synergy of defect and interface engineering of NiMoO4 NWAs realized by carbon deposition effectively improves the overall electrochemical performance of the resultant NiMoO4−x@C composite electrode in supercapacitors.
To further assess the practical application potential of a NiMoO4−x@C sample, an asymmetric supercapacitor device (NiMoO4−x@C//AC) was assembled with the NiMoO4−x@C as a positive electrode and activated carbon (AC) as a negative electrode. Before testing the ASC device, we performed the CV measurements of NiMoO4−x@C and AC electrodes in a three-electrode system at a scan rate of 5 mV s−1 to estimate the suitable operating voltage range (Figure 10a). The maximum operating voltage of NiMoO4−x@C//AC ASC was determined to be 1.6 V. The CV curves of NiMoO4−x@C//AC ASC at different scan rates are shown in Figure 10b. All the curves display obvious redox peaks, indicating the main contribution from pseudocapacitance. An increased separation of redox peak position can be noted along with the increase of scan rates due to the increased polarization. Figure 10c shows the GCD curves of the NiMoO4−x@C//AC ASC at different current densities while Figure 10d shows the specific capacitance calculated from different current densities. The overall capacitance of the ASC was calculated to be 1.01 F cm−2 (156.25 F g−1) at 1 mA cm−2. Furthermore, the NiMoO4−x@C//AC ASC device has demonstrated a good capacitance retention of 83.6% after 6000 cycles at 20 mA cm−2 (Figure 10e). As a result, the ASC device can power a yellow LED (inset of Figure 10e), showing its potential in practical applications. In addition, the electrochemical performances of our NiMoO4−x@C//AC ASC device are also superior or comparable to some recently reported NiMoO4-based electrode materials for ASCs as summarized in Table S1.

3. Conclusions

In summary, synchronous defect and interface engineering was implemented in NiMoO4 material via the formation of oxygen vacancies and the coating of the carbon on NiMoO4 nanowire arrays through a simple hydrothermal method paired with sucrose pyrolysis. During this process, an amorphous carbon layer was homogeneously deposited on the surface of NiMoO4 nanowires and oxygen vacancies were created on the NiMoO4 surface during the carbonization of sucrose. The deposited carbon layer and formed oxygen vacancies in NiMoO4 boosted the electronic conductivity of NiMoO4 nanowires. In addition, the coated carbon layer also improved the structural integrity of the NiMoO4 electrode during long-term operation in supercapacitors. Consequently, the resultant NiMoO4−x@C heterostructure electrode achieved a high specific capacitance of 2.24 F cm−2 (1720 F g−1) and maintained a good capacitance retention of about 83.6% after 6000 cycles at 20 mA cm−2. In addition, the as-assembled NiMoO4−x@C//activated carbon asymmetric supercapacitor device manifested a high energy density of 51.6 W h kg−1 at a high power density of 203.95 W kg−1, indicating that NiMoO4−x@C composite is a suitable electrode material for supercapacitor applications. The proposed synergistic defect and interface engineering strategy herein can be extended for the design and development of other novel composite electrode materials for applications in electrochemical energy storage and conversion.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano12071094/s1, Figure S1: The EDS analysis of NiMoO4−x@C composite and NiMoO4 NWAs, Figure S2: SEM images of the NiMoO4−x@C composite prepared under different annealing temperature: (a) 200 °C, (b) 400 °C, (c) 600 °C and (d) 800 °C, Figure S3: Thermogravimetric (TG) and differential scanning calorimetry (DSC) analysis of sucrose in Ar atmosphere at a ramping rate of 10 °C min−1, Figure S4: (a) CV and (b) GCD of NiMoO4−x@C samples prepared under different annealing temperatures, Figure S5: SEM images of the (a) NiMoO4−x@C and (b) NiMoO4 NWAs electrodes after 10,000 cycles at 20 mA cm−2, Figure S6: Optimized geometry structures of the (a) pristine NiMoO4 (110) surface and (b) oxygen-deficient NiMoO4 (110) surface obtained by removing one surface O atom (green). The grey, light pink and red balls represent the Ni, Mo and O atoms, respectively [60,61,62,63], Figure S7: Charge density difference contours of after adsorption of one HO molecule on (a) Mo- and (b) Ni-side of NiMoO4 (110) surface, respectively. The yellow and cyan colors denote the electron gain and loss, respectively. The isosurface level is 0.001 e bohr−3, Table S1: Brief comparison of electrochemical performance of current work with recently relevant literature [20,22,30,55,64,65,66,67,68,69,70,71,72].

Author Contributions

Conceptualization, C.Q. and P.W.; methodology, P.W. and X.D.; software, R.Z. and H.-E.W.; validation, P.W., X.D. and R.Z.; formal analysis, P.W., X.D. and C.Q.; investigation, P.W. and X.D.; resources, C.Q., H.-E.W. and Y.L.; data curation, P.W., C.Q. and H.-E.W.; writing—original draft preparation, P.W.; writing—review and editing, C.Q., T.Z. and H.-E.W.; visualization, P.W. and R.Z.; supervision, C.Q. and H.-E.W.; project administration, C.Q. and Y.L.; funding acquisition, C.Q. and H.-E.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (Grant No. 52162037) and Yunnan Fundamental Research Projects (grant No. 2019FD113 & 202101AT070120).

Data Availability Statement

All data included in this study are available upon request by contact with the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive oxide materials for high-rate electrochemical energy storage. Energy Environ. Sci. 2014, 7, 1597–1614. [Google Scholar] [CrossRef] [Green Version]
  2. Ma, H.N.; He, J.; Xiong, D.B.; Wu, J.S.; Li, Q.Q.; Dravid, V.; Zhao, Y.F. Nickel Cobalt Hydroxide Reduced Graphene Oxide Hybrid Nanolayers for High Performance Asymmetric Supercapacitors with Remarkable Cycling Stability. ACS Appl. Mater. Interfaces 2016, 8, 1992–2000. [Google Scholar] [CrossRef] [PubMed]
  3. Yun, Q.B.; Lu, Q.P.; Zhang, X.; Tan, C.L.; Zhang, H. Three-Dimensional Architectures Constructed from Transition-Metal Dichalcogenide Nanomaterials for Electrochemical Energy Storage and Conversion. Angew. Chem. Int. Ed. 2018, 57, 626–646. [Google Scholar] [CrossRef] [PubMed]
  4. Miao, C.X.; Zhou, C.L.; Wang, H.E.; Zhu, K.; Ye, K.; Wang, Q.; Yan, J.; Cao, D.X.; Li, N.; Wang, G.L. Hollow Co-Mo-Se nanosheet arrays derived from metal-organic framework for high-performance supercapacitors. J. Power Sources 2021, 490, 229532. [Google Scholar] [CrossRef]
  5. Lukatskaya, M.R.; Kota, S.; Lin, Z.F.; Zhao, M.Q.; Shpigel, N.; Levi, M.D.; Halim, J.; Taberna, P.L.; Barsoum, M.; Simon, P.; et al. Ultra-high-rate pseudocapacitive energy storage in two-dimensional transition metal carbides. Nat. Energy 2017, 2, 17105. [Google Scholar] [CrossRef]
  6. Ren, J.; Shen, M.; Li, Z.L.; Yang, C.M.; Liang, Y.; Wang, H.E.; Li, J.H.; Li, N.; Qian, D. Towards high-performance all-solid-state asymmetric supercapacitors: A hierarchical doughnut-like Ni3S2 PPy core-shell heterostructure on nickel foam electrode and density functional theory calculations. J. Power Sources 2021, 501, 230003. [Google Scholar] [CrossRef]
  7. Sun, L.; Sun, Y.D.; Fu, Q.; Pan, C.X. Facile preparation of NiO nanoparticles anchored on N/P-codoped 3D carbon nanofibers network for high-performance asymmetric supercapacitors. J. Alloys Compd. 2021, 888, 161488. [Google Scholar] [CrossRef]
  8. Liu, Y.; Wang, Y.F.; Meng, Y.; Plamthottam, R.; Tjiu, W.W.; Zhang, C.; Liu, T.X. Ultrathin Polypyrrole Layers Boosting MoO3 as Both Cathode and Anode Materials for a 2.0 V High-Voltage Aqueous Supercapacitor. ACS Appl. Mater. Interfaces 2022, 14, 4490–4499. [Google Scholar] [CrossRef]
  9. Shu, T.; Wang, H.; Li, Q.; Feng, Z.P.; Wei, F.X.; Yao, K.X.; Sun, Z.; Qi, J.Q.; Sui, Y.W. Highly stable Co3O4 nanoparticles/carbon nanosheets array derived from flake-like ZIF-67 as an advanced electrode for supercapacacitor. Chem. Eng. J. 2021, 419, 129631. [Google Scholar] [CrossRef]
  10. Pandit, B.; Goda, E.S.; Abu Elella, M.H.; Rehman, A.U.; Hong, S.E.; Rondiya, S.R.; Barkataki, P.; Shaikh, S.F.; Al-Enizi, A.M.; El-Bahy, S.M.; et al. One-pot hydrothermal preparation of hierarchical manganese oxide nanorods for high-performance symmetric supercapacitors. J. Energy Chem. 2022, 65, 116–126. [Google Scholar] [CrossRef]
  11. Wei, S.; Wan, C.C.; Zhang, L.Y.; Liu, X.Y.; Tian, W.Y.; Su, J.H.; Cheng, W.J.; Wu, Y.Q. N-doped and oxygen vacancy-rich NiCo2O4 nanograss for supercapacitor electrode. Chem. Eng. J. 2022, 429, 132242. [Google Scholar] [CrossRef]
  12. Liu, C.X.; Zhao, Z.; Liu, Y.; Lu, Q.S. Carbon dots decorated zinc cobaltite nanowires-assembled hierarchical arrays supported on nickel foam as binder-free electrodes for high performance supercapacitors. J. Power Sources 2022, 519, 230780. [Google Scholar] [CrossRef]
  13. Liu, H.Q.; Dai, M.Z.; Zhao, D.P.; Wu, X.; Wang, B. Realizing Superior Electrochemical Performance of Asymmetric Capacitors through Tailoring Electrode Architectures. ACS Appl. Energy Mater. 2020, 3, 7004–7010. [Google Scholar] [CrossRef]
  14. Li, P.X.; Ruan, C.H.; Xu, J.; Xie, Y.B. Supercapacitive performance of CoMoO4 with oxygen vacancy porous nanosheet. Electrochim. Acta 2020, 330, 135334. [Google Scholar] [CrossRef]
  15. Zhu, Z.; Sun, Y.; Li, C.S.; Yang, C.; Li, L.; Zhu, J.H.; Chou, S.L.; Wang, M.M.; Wang, D.D.; Li, Y.L. Mini-review: Progress on micro/nanoscale MnMoO4 as an electrode material for advanced supercapacitor applications. Mater. Chem. Front. 2021, 5, 7403–7418. [Google Scholar] [CrossRef]
  16. Guo, D.; Zhang, P.; Zhang, H.M.; Yu, X.Z.; Zhu, J.; Li, Q.H.; Wang, T.H. NiMoO4 nanowires supported on Ni foam as novel advanced electrodes for supercapacitors. J. Mater. Chem. A 2013, 1, 9024–9027. [Google Scholar] [CrossRef]
  17. Denis, D.K.; Sun, X.; Zhang, J.Y.; Wang, Y.Y.; Hou, L.R.; Li, J.; Yuan, C.Z. Solid Solution Engineering of Co-Ni-Based Ternary Molybdate Nanorods toward Hybrid Supercapacitors and Lithium-Ion Batteries as High-Performance Electrodes. ACS Appl. Energy Mater. 2020, 3, 3955–3965. [Google Scholar] [CrossRef]
  18. Guo, D.; Luo, Y.Z.; Yu, X.Z.; Li, Q.H.; Wang, T.H. High performance NiMoO4 nanowires supported on carbon cloth as advanced electrodes for symmetric supercapacitors. Nano Energy 2014, 8, 174–182. [Google Scholar] [CrossRef]
  19. Hong, J.; Lee, Y.W.; Hou, B.; Ko, W.; Lee, J.; Pak, S.; Hong, J.; Morris, S.M.; Cha, S.; Sohn, J.I.; et al. Solubility-Dependent NiMoO4 Nanoarchitectures: Direct Correlation between Rationally Designed Structure and Electrochemical Pseudokinetics. ACS Appl. Mater. Interfaces 2016, 8, 35227–35234. [Google Scholar] [CrossRef] [Green Version]
  20. Peng, S.J.; Li, L.L.; Wu, H.B.; Madhavi, S.; Lou, X.W. Controlled Growth of NiMoO4 Nanosheet and Nanorod Arrays on Various Conductive Substrates as Advanced Electrodes for Asymmetric Supercapacitors. Adv. Energy Mater. 2015, 5, 1401172. [Google Scholar] [CrossRef]
  21. Yin, Z.X.; Zhang, S.; Chen, Y.J.; Gao, P.; Zhu, C.L.; Yang, P.P.; Qi, L.H. Hierarchical nanosheet-based NiMoO4 nanotubes: Synthesis and high supercapacitor performance. J. Mater. Chem. A 2015, 3, 739–745. [Google Scholar] [CrossRef]
  22. Lin, L.Y.; Liu, T.M.; Liu, J.L.; Sun, R.; Hao, J.H.; Ji, K.M.; Wang, Z.C. Facile synthesis of groove-like NiMoO4 hollow nanorods for high-performance supercapacitors. Appl. Surf. Sci. 2016, 360, 234–239. [Google Scholar] [CrossRef]
  23. Yao, M.M.; Hu, Z.H.; Liu, Y.F.; Liu, P.P. A novel synthesis of size-controllable mesoporous NiMoO4 nanospheres for supercapacitor applications. Ionics 2016, 22, 701–709. [Google Scholar] [CrossRef]
  24. Jinlong, L.; Meng, Y.; Tongxiang, L. Enhanced performance of NiMoO4 nanoparticles and quantum dots and reduced nanohole graphene oxide hybrid for supercapacitor applications. Appl. Surf. Sci. 2017, 419, 624–630. [Google Scholar] [CrossRef]
  25. Huang, L.; Xiang, J.W.; Zhang, W.; Chen, C.J.; Xu, H.H.; Huang, Y.H. 3D interconnected porous NiMoO4 nanoplate arrays on Ni foam as high-performance binder-free electrode for supercapacitors. J. Mater. Chem. A 2015, 3, 22081–22087. [Google Scholar] [CrossRef]
  26. Wang, Z.J.; Wei, G.J.; Du, K.; Zhao, X.X.; Liu, M.; Wang, S.T.; Zhou, Y.; An, C.H.; Zhang, J. Ni Foam-Supported Carbon-Sheathed NiMoO4 Nanowires as Integrated Electrode for High-Performance Hybrid Supercapacitors. ACS Sustain. Chem. Eng. 2017, 5, 5964–5971. [Google Scholar] [CrossRef]
  27. Xu, Y.K.; Xuan, H.C.; Gao, J.H.; Liang, T.; Han, X.K.; Yang, J.; Zhang, Y.Q.; Li, H.; Han, P.D.; Du, Y.W. Hierarchical three-dimensional NiMoO4-anchored rGO/Ni foam as advanced electrode material with improved supercapacitor performance. J. Mater. Sci. 2018, 53, 8483–8498. [Google Scholar] [CrossRef]
  28. Huang, Z.Y.; Zhang, Z.; Qi, X.; Ren, X.H.; Xu, G.H.; Wan, P.B.; Sun, X.M.; Zhang, H. Wall-like hierarchical metal oxide nanosheet arrays grown on carbon cloth for excellent supercapacitor electrodes. Nanoscale 2016, 8, 13273–13279. [Google Scholar] [CrossRef] [Green Version]
  29. Huang, Y.P.; Cui, F.; Zhao, Y.; Lian, J.B.; Bao, J.; Liu, T.X.; Li, H.M. NiMoO4 nanorod deposited carbon sponges with ant-nest-like interior channels for high-performance pseudocapacitors. Inorg. Chem. Front. 2018, 5, 1594–1601. [Google Scholar] [CrossRef]
  30. Murugan, E.; Govindaraju, S.; Santhoshkumar, S. Hydrothermal synthesis, characterization and electrochemical behavior of NiMoO4 nanoflower and NiMoO4/rGO nanocomposite for high-performance supercapacitors. Electrochim. Acta 2021, 392, 138973. [Google Scholar] [CrossRef]
  31. Huang, B.J.; Yao, D.C.; Yuan, J.J.; Tao, Y.R.; Yin, Y.X.; He, G.Y.; Chen, H.Q. Hydrangea-like NiMoO4-Ag/rGO as Battery-type electrode for hybrid supercapacitors with superior stability. J. Colloid Interf. Sci. 2022, 606, 1652–1661. [Google Scholar] [CrossRef] [PubMed]
  32. Yao, P.P.; Li, C.Y.; Yu, J.L.; Zhang, S.; Zhang, M.; Liu, H.C.; Ji, M.W.; Cong, G.T.; Zhang, T.; Zhu, C.Z.; et al. High performance flexible energy storage device based on copper foam supported NiMoO4 nanosheets-CNTs-CuO nanowires composites with core-shell holey nanostructure. J. Mater. Sci. Technol. 2021, 85, 87–94. [Google Scholar] [CrossRef]
  33. Zhu, D.; Sun, X.; Yu, J.; Liu, Q.; Liu, J.Y.; Chen, R.R.; Zhang, H.S.; Song, D.L.; Li, R.M.; Wang, J. Three-dimensional heterostructured polypyrrole/nickel molybdate anchored on carbon cloth for high-performance flexible supercapacitors. J. Colloid Interf. Sci. 2020, 574, 355–363. [Google Scholar] [CrossRef] [PubMed]
  34. Hao, Y.Y.; Huang, H.; Wang, Q.Q.; Wang, Q.; Zhou, G.W. Nitrogen-doped carbon/NiMoO4 nanospheres assembled by nanosheets and ultrasmall nanoparticles for supercapacitors. Chem. Phys. Lett. 2019, 728, 215–223. [Google Scholar] [CrossRef]
  35. Tong, B.L.; Wei, W.T.; Chen, X.L.; Wang, J.; Ye, W.Y.; Cui, S.Z.; Chen, W.H.; Mi, L.W. Designed synthesis of porous NiMoO4/C composite nanorods for asymmetric supercapacitors. Cryst. Eng. Comm. 2019, 21, 5492–5499. [Google Scholar] [CrossRef]
  36. Li, Y.; Zhang, S.M.; Ma, M.Y.; Mu, X.M.; Zhang, Y.X.; Du, J.W.; Hu, Q.; Huang, B.Y.; Hua, X.H.; Liu, G.; et al. Manganese-doped nickel molybdate nanostructures for high-performance asymmetric supercapacitors. Chem. Eng. J. 2019, 372, 452–461. [Google Scholar] [CrossRef]
  37. Yuan, J.J.; Yao, D.C.; Jiang, L.; Tao, Y.R.; Che, J.F.; He, G.Y.; Chen, H.Q. Mn-Doped NiMoO4 Mesoporous Nanorods/Reduced Graphene Oxide Composite for High-Performance All-Solid-State Supercapacitor. ACS Appl. Energy Mater. 2020, 3, 1794–1803. [Google Scholar] [CrossRef]
  38. Wang, F.F.; Ma, K.; Tian, W.; Dong, J.C.; Han, H.; Wang, H.P.; Deng, K.; Yue, H.R.; Zhang, Y.X.; Jiang, W.; et al. P-Doped NiMoO4 parallel arrays anchored on cobalt carbonate hydroxide with oxygen vacancies and mass transfer channels for supercapacitors and oxygen evolution. J. Mater. Chem. A 2019, 7, 19589–19596. [Google Scholar] [CrossRef]
  39. Sharma, P.; Sundaram, M.M.; Watcharatharapong, T.; Laird, D.; Euchner, H.; Ahuja, R. Zn Metal Atom Doping on the Surface Plane of One-Dimesional NiMoO4 Nanorods with Improved Redox Chemistry. ACS Appl. Mater. Interfaces 2020, 12, 44815–44829. [Google Scholar] [CrossRef]
  40. Cui, S.Z.; Wang, F.Q.; Sun, K.J.; Wang, X.; Hu, Q.Z.; Peng, H.; Ma, G.F.; Lei, Z.Q. High-Performance Hybrid Supercapacitors Based on Ce-Doped NiMoO4 Nanosheets and Fe3O4@Bi2O3 Nanoarrays. J. Phys. Chem. C 2021, 125, 18129–18140. [Google Scholar] [CrossRef]
  41. Qing, C.; Yang, C.X.; Chen, M.Y.; Li, W.H.; Wang, S.Y.; Tang, Y.W. Design of oxygen-deficient NiMoO4 nanoflake and nanorod arrays with enhanced supercapacitive performance. Chem. Eng. J. 2018, 354, 182–190. [Google Scholar] [CrossRef]
  42. Sivakumar, P.; Raj, C.J.; Park, J.; Jung, H. Synergistic effects of nanoarchitecture and oxygen vacancy in nickel molybdate hollow sphere towards a high-performance hybrid supercapacitor. Int. J. Energy Res. 2021, 45, 21516–21526. [Google Scholar] [CrossRef]
  43. Zhu, S.; Le, J.Y.; Mao, Y.J.; Chen, S.X.; Han, X.X.; Zeng, Z.L.; Wang, J.; Deng, S.G. Synergistic engineering of fluorine doping and oxygen vacancies towards high-energy and long-lifespan flexible solid-state asymmetric supercapacitor. Ionics 2021, 27, 2649–2658. [Google Scholar] [CrossRef]
  44. Zhang, X.Y.; Wei, L.; Guo, X. Ultrathin mesoporous NiMoO4-modified MoO3 core/shell nanostructures: Enhanced capacitive storage and cycling performance for supercapacitors. Chem. Eng. J. 2018, 353, 615–625. [Google Scholar] [CrossRef]
  45. Shen, J.W.; Wang, Q.G.; Zhang, K.; Wang, S.M.; Li, L.; Dong, S.B.; Zhao, S.T.; Chen, J.; Sun, R.S.; Wang, Y.; et al. Flexible carbon cloth based solid-state supercapacitor from hierarchical holothurian-morphological NiCo2O4@NiMoO4/PANI. Electrochim. Acta 2019, 320, 134578. [Google Scholar] [CrossRef]
  46. Xu, R.; Lin, J.M.; Wu, J.H.; Huang, M.L.; Fan, L.Q.; Xu, Z.D.; Song, Z.Y. A high-performance pseudocapacitive electrode material for supercapacitors based on the unique NiMoO4/NiO nanoflowers. Appl. Surf. Sci. 2019, 463, 721–731. [Google Scholar] [CrossRef]
  47. Yu, D.Y.; Zhang, Z.Q.; Teng, Y.F.; Meng, Y.N.; Wu, Y.P.; Liu, X.L.; Hua, Y.J.; Zhao, X.D.; Liu, X.Y. Fabrication of CuO@NiMoO4 core-shell nanowire arrays on copper foam and their application in high-performance all-solid-state asymmetric supercapacitors. J. Power Sources 2019, 440, 227164. [Google Scholar] [CrossRef]
  48. Liu, Y.; Ma, Z.L.; Xin, N.; Ying, Y.L.; Shi, W.D. High-performance supercapacitor based on highly active P-doped one-dimension/two-dimension hierarchical NiCo2O4/NiMoO4 for efficient energy storage. J. Colloid Interf. Sci. 2021, 601, 793–802. [Google Scholar] [CrossRef]
  49. Zeng, Y.; Liao, J.Z.; Wei, B.B.; Huang, Z.; Zhu, W.J.; Zheng, J.X.; Liang, H.F.; Zhang, Y.Z.; Wang, Z.C. Tuning the electronic structure of NiMoO4 by coupling with SnO2 for high-performance hybrid supercapacitors. Chem. Eng. J. 2021, 409, 128297. [Google Scholar] [CrossRef]
  50. Chen, C.; Yan, D.; Luo, X.; Gao, W.J.; Huang, G.J.; Han, Z.W.; Zeng, Y.; Zhu, Z.H. Construction of Core-Shell NiMoO4@Ni-Co-S Nanorods as Advanced Electrodes for High-Performance Asymmetric Supercapacitors. ACS Appl. Mater. Interfaces 2018, 10, 4662–4671. [Google Scholar] [CrossRef]
  51. Chen, F.S.; Ji, S.; Liu, Q.B.; Wang, H.; Liu, H.; Brett, D.J.L.; Wang, G.X.; Wang, R.F. Rational Design of Hierarchically Core-Shell Structured Ni3S2@NiMoO4 Nanowires for Electrochemical Energy Storage. Small 2018, 14, 1800791. [Google Scholar] [CrossRef] [PubMed]
  52. Acharya, J.; Ojha, G.P.; Kim, B.S.; Pant, B.; Park, M. Modish Designation of Hollow-Tubular rGO-NiMoO4@Ni-Co-S Hybrid Core-shell Electrodes with Multichannel Superconductive Pathways for High-Performance Asymmetric Supercapacitors. ACS Appl. Mater. Interfaces 2021, 13, 17487–17500. [Google Scholar] [CrossRef] [PubMed]
  53. Hu, K.L.; Jeong, S.; Elumalai, G.; Kukunuri, S.; Fujita, J.; Ito, Y. Phase-Dependent Reactivity of Nickel Molybdates for Electrocatalytic Urea Oxidation. ACS Appl. Energy Mater. 2020, 3, 7535–7542. [Google Scholar] [CrossRef]
  54. Wang, Y.; Liu, J.; Chen, X.; Kang, B.; Wang, H.-E.; Xiong, P.; Chen, Q.; Wei, M.; Li, N.; Qian, Q.; et al. Structural engineering of tin sulfides anchored on nitrogen/phosphorus dual-doped carbon nanofibres in sodium/potassium-ion batteries. Carbon 2022, 189, 46–56. [Google Scholar] [CrossRef]
  55. Nti, F.; Anang, D.A.; Han, J.I. Facilely synthesized NiMoO4/CoMoO4 nanorods as electrode material for high performance supercapacitor. J. Alloys Compd. 2018, 742, 342–350. [Google Scholar] [CrossRef]
  56. Zhang, Z.; Liu, Y.; Huang, Z.; Ren, L.; Qi, X.; Wei, X.; Zhong, J. Facile hydrothermal synthesis of NiMoO4@CoMoO4 hierarchical nanospheres for supercapacitor applications. Phys. Chem. Chem. Phys. 2015, 17, 20795–20804. [Google Scholar] [CrossRef]
  57. Zhao, X.; Wang, H.E.; Chen, X.X.; Cao, J.; Zhao, Y.D.; Neale, Z.G.; Cai, W.; Sui, J.H.; Cao, G.Z. Tubular MoO2 organized by 2D assemblies for fast and durable alkali-ion storage. Energy Storage Mater. 2018, 11, 161–169. [Google Scholar] [CrossRef]
  58. Wang, H.E.; Zhao, X.; Yin, K.L.; Li, Y.; Chen, L.H.; Yang, X.Y.; Zhang, W.J.; Su, B.L.; Cao, G.Z. Superior Pseudocapacitive Lithium-Ion Storage in Porous Vanadium Oxides C Heterostructure Composite. ACS Appl. Mater. Interfaces 2017, 9, 43665–43673. [Google Scholar] [CrossRef]
  59. Wang, H.E.; Yin, K.L.; Qin, N.; Zhao, X.; Xia, F.J.; Hu, Z.Y.; Guo, G.L.; Cao, G.Z.; Zhang, W.J. Oxygen-deficient titanium dioxide as a functional host for lithium-sulfur batteries. J. Mater. Chem. A 2019, 7, 10346–10353. [Google Scholar] [CrossRef]
  60. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  61. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  62. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Cryst. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  63. Wang, V.; Xu, N.; Liu, J.C.; Tang, G.; Geng, W.T. VASPKIT: A user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comp. Phys. Comm. 2021, 267, 108033. [Google Scholar] [CrossRef]
  64. Zhao, Y.; Zhang, P.; Fu, W.; Ma, X.; Zhou, J.; Zhang, X.; Li, J.; Xie, E.; Pan, X. Understanding the role of Co3O4 on stability between active hierarchies and scaffolds: An insight into NiMoO4 composites for supercapacitors. Appl. Surf. Sci. 2017, 416, 160–167. [Google Scholar] [CrossRef]
  65. Xuan, H.; Xu, Y.; Zhang, Y.; Li, H.; Han, P.; Du, Y. One-step combustion synthesis of porous CNTs/C/NiMoO4 composites for high-performance asymmetric supercapacitors. J. Alloys Compd. 2018, 745, 135–146. [Google Scholar] [CrossRef]
  66. Huang, Y.; Cui, F.; Zhao, Y.; Lian, J.; Bao, J.; Li, H. Controlled growth of ultrathin NiMoO4 nanosheets on carbon nanofiber membrane as advanced electrodes for asymmetric supercapacitors. J. Alloys Compd. 2018, 753, 176–185. [Google Scholar] [CrossRef]
  67. Hussain, S.; Javed, M.S.; Asim, S.; Shaheen, A.; Khan, A.J.; Abbas, Y.; Ullah, N.; Iqbal, A.; Wang, M.; Qiao, G. Novel gravel-like NiMoO4 nanoparticles on carbon cloth for outstanding supercapacitor applications. Ceram. Int. 2020, 46, 6406–6412. [Google Scholar] [CrossRef]
  68. Feng, X.; Ning, J.; Wang, D.; Zhang, J.; Xia, M.; Wang, Y.; Hao, Y. Heterostructure arrays of NiMoO4 nanoflakes on N-doping of graphene for high-performance asymmetric supercapacitors. J. Alloys Compd. 2020, 816, 152625. [Google Scholar] [CrossRef]
  69. Zhong, Y.; Liu, T.; Zhang, A.; Cui, L.; Liu, X.; Zheng, R.; Liu, J. Controllable synthesis of Ni1-xCoxMoO4 with tunable morphologies for high-performance asymmetric supercapacitors. J. Alloys Compd. 2021, 850, 156734. [Google Scholar] [CrossRef]
  70. Chen, L.; Deng, W.; Chen, Z.; Wang, X. Hetero-architectured core–shell NiMoO4@Ni9S8/MoS2 nanorods enabling high-performance supercapacitors. J. Mater. Res. 2022, 37, 284–293. [Google Scholar] [CrossRef]
  71. Muthu, D.; Vargheese, S.; Haldorai, Y.; Kumar, R.T.R. NiMoO4/reduced graphene oxide composite as an electrode material for hybrid supercapacitor. Mater. Sci. Semicond. Proc. 2021, 135, 106078. [Google Scholar] [CrossRef]
  72. Yi, T.-F.; Qiu, L.-Y.; Mei, J.; Qi, S.-Y.; Cui, P.; Luo, S.; Zhu, Y.-R.; Xie, Y.; He, Y.-B. Porous spherical NiO@NiMoO4@PPy nanoarchitectures as advanced electrochemical pseudocapacitor materials. Sci. Bull. 2020, 65, 546–556. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic illustration of the synthesis process of oxygen-deficient NiMoO4@carbon nanowire arrays (NWAs) (denoted as “NiMoO4−x@C”). Step 1, growth of the NiMoO4 NWAs precursor directly on a ferronickel foam substrate using a hydrothermal process; Step 2, conversion of the NiMoO4 NWAs precursor into the NiMoO4 NWAs via annealing in Ar; Step 3, fabrication of NiMoO4−x@C composite by sucrose coating followed by annealing in Ar.
Figure 1. Schematic illustration of the synthesis process of oxygen-deficient NiMoO4@carbon nanowire arrays (NWAs) (denoted as “NiMoO4−x@C”). Step 1, growth of the NiMoO4 NWAs precursor directly on a ferronickel foam substrate using a hydrothermal process; Step 2, conversion of the NiMoO4 NWAs precursor into the NiMoO4 NWAs via annealing in Ar; Step 3, fabrication of NiMoO4−x@C composite by sucrose coating followed by annealing in Ar.
Nanomaterials 12 01094 g001
Figure 2. XRD patterns of NiMoO4 NWAs (black) and NiMoO4x@C composite (red). The inset (top left) shows the structural model of NiMoO4 crystal, whereas the green, blue, and red balls represent the Ni, Mo, and O atoms, respectively. The two strongest diffraction peaks in the square regions of dotted line are from the ferronickel foam substrate.
Figure 2. XRD patterns of NiMoO4 NWAs (black) and NiMoO4x@C composite (red). The inset (top left) shows the structural model of NiMoO4 crystal, whereas the green, blue, and red balls represent the Ni, Mo, and O atoms, respectively. The two strongest diffraction peaks in the square regions of dotted line are from the ferronickel foam substrate.
Nanomaterials 12 01094 g002
Figure 3. Raman spectra of NiMoO4−x@C (black) and pristine NiMoO4 NWAs (red) samples.
Figure 3. Raman spectra of NiMoO4−x@C (black) and pristine NiMoO4 NWAs (red) samples.
Nanomaterials 12 01094 g003
Figure 4. SEM images of (a) neat NiMoO4 NWAs and (b) NiMoO4−x@C samples.
Figure 4. SEM images of (a) neat NiMoO4 NWAs and (b) NiMoO4−x@C samples.
Nanomaterials 12 01094 g004
Figure 5. TEM images (a,c) and HRTEM micrographs (b,d) of pure NiMoO4 NWAs (a,b) and NiMoO4−x@C (c,d) samples. The insets in (a,c) are corresponding SAED patterns taken from a single nanowire.
Figure 5. TEM images (a,c) and HRTEM micrographs (b,d) of pure NiMoO4 NWAs (a,b) and NiMoO4−x@C (c,d) samples. The insets in (a,c) are corresponding SAED patterns taken from a single nanowire.
Nanomaterials 12 01094 g005
Figure 6. High-resolution XPS spectra of (a,c,e) pristine NiMoO4 and (b,d,f) NiMoO4−x@C samples; (a,b) Ni 2p, (c,d) Mo 3d, (e,f) O 1s.
Figure 6. High-resolution XPS spectra of (a,c,e) pristine NiMoO4 and (b,d,f) NiMoO4−x@C samples; (a,b) Ni 2p, (c,d) Mo 3d, (e,f) O 1s.
Nanomaterials 12 01094 g006
Figure 7. (a) CV curves of the NiMoO4 NWAs and NiMoO4−x@C at a scan rate of 20 mV s−1. (b) GCD curves and (c) EIS spectra of NiMoO4 NWAs and NiMoO4−x@C and its corresponding equivalent fitting circuit. (d) GCD test of NiMoO4−x@C at different current densities. (e) Cycle test of NiMoO4 NWAs and NiMoO4−x@C at 20 mA cm−2 over 6000 cycles.
Figure 7. (a) CV curves of the NiMoO4 NWAs and NiMoO4−x@C at a scan rate of 20 mV s−1. (b) GCD curves and (c) EIS spectra of NiMoO4 NWAs and NiMoO4−x@C and its corresponding equivalent fitting circuit. (d) GCD test of NiMoO4−x@C at different current densities. (e) Cycle test of NiMoO4 NWAs and NiMoO4−x@C at 20 mA cm−2 over 6000 cycles.
Nanomaterials 12 01094 g007
Figure 8. The optimized adsorption configurations of one HO molecule anchored on exposed (a,b) Mo- and (c,d) Ni- atoms of pristine NiMoO4 (110) surfaces, leading to an Eads of −0.50 and −5.97 eV, respectively. The black dotted ellipsoids in (a,c) mark the positions of the adsorbed HO molecule on the NiMoO4 (110) surface.
Figure 8. The optimized adsorption configurations of one HO molecule anchored on exposed (a,b) Mo- and (c,d) Ni- atoms of pristine NiMoO4 (110) surfaces, leading to an Eads of −0.50 and −5.97 eV, respectively. The black dotted ellipsoids in (a,c) mark the positions of the adsorbed HO molecule on the NiMoO4 (110) surface.
Nanomaterials 12 01094 g008
Figure 9. The optimized adsorption configurations of one HO molecule anchored on NiMoO4 (110) surfaces with one surface oxygen vacancy (Vo), leading to an Eads of −3.63 eV. The black dotted ellipsoid in (a) labels the position of the adsorbed HO molecule, which is at the vicinity of the Vo. The yellow and cyan colors in the charge density difference contour (b) represent the electron accumulation and depletion, respectively. The isosurface level is 0.001 e bohr−3.
Figure 9. The optimized adsorption configurations of one HO molecule anchored on NiMoO4 (110) surfaces with one surface oxygen vacancy (Vo), leading to an Eads of −3.63 eV. The black dotted ellipsoid in (a) labels the position of the adsorbed HO molecule, which is at the vicinity of the Vo. The yellow and cyan colors in the charge density difference contour (b) represent the electron accumulation and depletion, respectively. The isosurface level is 0.001 e bohr−3.
Nanomaterials 12 01094 g009
Figure 10. Electrochemical performance of the asymmetric NiMoO4−x@C//AC supercapacitor. (a) CV curves of activated carbon (AC) and NiMoO4−x@C electrodes from −1.0~0 V and 0~0.6 V, respectively at 5 m V s−1; (b) CV curves; (c) GCD curves; (d) areal capacitance and specific capacitance; (e) cyclic stability of NiMoO4−x@C//AC.
Figure 10. Electrochemical performance of the asymmetric NiMoO4−x@C//AC supercapacitor. (a) CV curves of activated carbon (AC) and NiMoO4−x@C electrodes from −1.0~0 V and 0~0.6 V, respectively at 5 m V s−1; (b) CV curves; (c) GCD curves; (d) areal capacitance and specific capacitance; (e) cyclic stability of NiMoO4−x@C//AC.
Nanomaterials 12 01094 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, P.; Ding, X.; Zhe, R.; Zhu, T.; Qing, C.; Liu, Y.; Wang, H.-E. Synchronous Defect and Interface Engineering of NiMoO4 Nanowire Arrays for High-Performance Supercapacitors. Nanomaterials 2022, 12, 1094. https://doi.org/10.3390/nano12071094

AMA Style

Wang P, Ding X, Zhe R, Zhu T, Qing C, Liu Y, Wang H-E. Synchronous Defect and Interface Engineering of NiMoO4 Nanowire Arrays for High-Performance Supercapacitors. Nanomaterials. 2022; 12(7):1094. https://doi.org/10.3390/nano12071094

Chicago/Turabian Style

Wang, Pengcheng, Xinying Ding, Rongjie Zhe, Ting Zhu, Chen Qing, Yingkai Liu, and Hong-En Wang. 2022. "Synchronous Defect and Interface Engineering of NiMoO4 Nanowire Arrays for High-Performance Supercapacitors" Nanomaterials 12, no. 7: 1094. https://doi.org/10.3390/nano12071094

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