Next Article in Journal
Study on the Influencing Factors of Electrical Characteristics of Articulated Split-Zone Insulator Arc
Previous Article in Journal
Thermal Analysis of Parabolic and Fresnel Linear Solar Collectors Using Compressed Gases as Heat Transfer Fluid in CSP Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Design and Synthesis of Ni-Mo-Co Ternary Hydroxides Nanoflakes for Advanced Energy Storage Device Applications

Department of Chemical Engineering and Material Science, Wayne State University, Detroit, MI 48201, USA
*
Author to whom correspondence should be addressed.
Energies 2024, 17(16), 3881; https://doi.org/10.3390/en17163881
Submission received: 1 June 2024 / Revised: 11 July 2024 / Accepted: 28 July 2024 / Published: 6 August 2024
(This article belongs to the Section F3: Power Electronics)

Abstract

:
Three-dimensional interconnected mesoporous nanoflakes of amorphous Ni-Mo-Co trimetallic hydroxides were successfully deposited on a Ni foam (NF) using a facile, environmentally friendly, and scalable electrochemical deposition method. The elemental composition of the nanoflakes, including Ni2+, Mo6+, and Co2+, was characterized by X-ray photoelectron spectroscopy (XPS), while the morphology and particle size of the synthesized nanomaterials were examined using scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The Ni-Mo-Co trimetallic hydroxides on NF were employed as binder-free electrodes for supercapacitors. The Ni-Mo-Co trimetallic hydroxides with a Ni/Mo/Co ratio of 1/1/0.4 exhibited outstanding long-term cyclability over 5000 cycles, with a high reversible specific capacitance of 2700 F g−1 and a high capacitance retention of 96.63% at 10 A g−1. Furthermore, they demonstrated excellent rate performance, maintaining a capacitance of 2429 F g−1 at a current of 50 A g−1, which corresponds to approximately 80% capacitance retention compared to the capacitance at 2 A g−1. The superior performance of these Ni-Mo-Co trimetallic hydroxides can be attributed to their mesoporous hierarchical architecture, which provides large open spaces between the interconnected nanoflakes, numerous electroactive surface sites, facile electron transmission paths, and the synergistic effects of the trimetallic components. These findings demonstrate that Ni-Mo-Co trimetallic hydroxides are promising electrode materials, offering both high capacitance and long-term cyclability for supercapacitors.

1. Introduction

Currently, the development of electric vehicles and consumer electronics is hindered by the inadequate energy and power density, limited cycle life, and safety concerns of current energy storage systems. Batteries have a high energy density but require a long time to recharge. To achieve a high energy density with rapid power delivery and recharging (i.e., high power density), electrochemical capacitors (ECs), also known as supercapacitors (SCs), are promising advanced energy storage systems that can fill the gap between batteries and capacitors [1,2,3].
Supercapacitors, especially pseudocapacitors, have been attracting tremendous attention in recent years as a promising approach to address the relatively low power density of batteries and low energy density of the electric double layer capacitors (EDLCs), the other type of SCs. Different from EDLCs, which store energy via the separation of charge in a Helmholtz double layer, pseudocapacitors store energy via reversible surface Faradaic redox reactions involving charge transfer. As a result, there is a substantial difference in energy density between pseudocapacitors and EDLCs due to their distinct energy storage mechanisms [4]. EDLCs fall short of meeting the stringent requirements of energy-intensive applications due to the limited electric energy stored at the interface between the electrodes, typically composed of porous carbon, and the electrolyte. For instance, carbon-based materials in EDLCs exhibit a capacitance of no more than 150 F g−1 [5]. In pseudocapacitors, transition metal oxides/hydroxides, such as Ni (OH)2, Co (OH)2, NiO, CoO, Co3O4, RuO2, MnO2, and Fe3O4, have garnered increasing interest as pseudocapacitive materials owing to their high theoretical specific capacitance [6,7,8]. For example, by synthesizing a Co (OH)2 nanowire using a dual-template method, Xue et al. demonstrated a specific capacitance of 993 F g−1 at a current density of 1 A g−1 [9]. Hu et al. reported that Ni (OH)2 pseudocapacitive material shows a high specific capacitance of 2217 F g−1 [10]. Although monometallic hydroxides in pseudocapacitors exhibit a higher capacitance compared to EDLCs, several issues remain to be addressed, such as low-rate performance and challenges in achieving a high theoretical capacitance. These issues can be attributed to the intrinsic limitations of the aforementioned monometallic oxides and hydroxides, including poor electrical conductivity, poor electrochemical reversibility, cycling instability, and low-rate performance [11,12,13]. Additionally, the high cost of certain rare metals poses another challenge. For instance, RuO2, despite having a high theoretical specific capacitance of approximately 1000 F g−1, is prohibitively expensive, with costs exceeding USD 1 million for a vehicle-sized supercapacitor [13].
Recently, bimetallic oxides and hydroxides, such as NiCo2O4, NiMoO4, CoMoO4, MnMoO4, MnCo2O4, CoMn(OH)4, NiMn(OH)4, and CoFe(OH)5, have emerged as novel and promising electrode materials for high-performance supercapacitors, aiming to overcome the intrinsic limitations associated with monometallic oxides and hydroxides [14,15,16,17,18,19,20,21,22,23,24,25]. These bimetallic compounds demonstrate synergistic effects that mitigate the defects inherent in their monometallic counterparts. Moreover, the presence of multiple oxidation states in these multi-metallic compounds enables the generation of higher specific capacitances compared to single-metallic oxides [26,27]. For example, Ni (OH)2 suffers from low cyclic stability, poor electric conductivity, and poor rate performance, whereas NiCo(OH)4 possesses a high capacitance and excellent cyclic stability [28,29].
Similarly, as compared to mono-/double-metallic hydroxides/oxides, triple hydroxides/oxides may provide further electrochemical improvement by introducing one more metallic component [27]. For instance, in electrocatalyst applications, NiCoFe layered triple hydroxide nanosheets can be used as an electrocatalyst for water splitting and showed an improved performance compared to their mono-/double-metallic components [30]. NiMnCo or LiNiMnFeO materials have attracted a lot of attention due to their ability to address the intrinsic limitation of LiCoO2 or LiNiO2 for battery applications [31,32,33]. Although tri-metallic compounds exhibit improved electrocatalytic and electrochemical performance compared to mono-/bi-metallic compounds, no tri-metallic hydroxides are currently used in supercapacitors, and the effects of the composition and morphology of tri-metallic hydroxides on supercapacitor capacitance and durability have not been fully elucidated [3].
In the realm of supercapacitors, electrodes commonly rely on amorphous materials due to their potential to achieve higher capacity and superior electrochemical activity compared to crystalline counterparts [34,35]. However, the utilization of amorphous materials may incur challenges, such as pulverization and capacity degradation, owing to their inferior mechanical stabilization properties [36,37]. It is noteworthy that these drawbacks can be mitigated through the optimization of material morphology, for instance, by employing nanoscale three-dimensional structures [15,37,38,39], such as nanotubes [15], nanosheets [40,41], as well as nanobristles [42].
Herein, 3D interconnected mesoporous nanoflakes of amorphous Ni-Mo-Co triple metallic hydroxides (THs), deposited on Ni foam (NF) using a facile, environmentally friendly and scalable electrochemical deposition method, were investigated as supercapacitor electrode materials. The 3D interconnected mesoporous Ni-Mo-Co hydroxides nanoflakes grew vertically onto the surface of the conductive Ni foam, which can be directly used as binder-free electrodes for SCs. A summary of the SC values obtained using Ni, Co, or Mo hydroxide/oxide nanostructured electrodes during the last decade is shown in Table 1. Compared with the reported materials, Ni/Mo/Co of 1/1/0.4 on NF in this study was found to exhibit a better performance, with a reversible specific capacitance of 3074 F g−1 at a current of 2 A g−1 after the aging process. A series of tri-/bi-/mono-metallic hydroxide compounds were prepared for SCs, and correlations between Ni, Mo, and Co composition, as well as morphology, and their electrochemical capacitive performance were investigated in this research.

2. Experimental Section

2.1. Preparation of Ni-Mo-Co Triple Hydroxide (THs) Nanoflakes

All chemicals were of reagent-grade quality and used without further purification. The Ni foam (10 × 10 mm) was ultrasonically cleaned with 2 M HCl, ethanol, and deionized (DI) water and dried in a vacuum oven for 24 h before electrodeposition. The electrodeposition of Ni-Mo-Co triple hydroxides (THs) was performed in a standard three-electrode electrochemical cell at room temperature—graphite foil (20 × 10 mm) as the counter electrode, Ni foam as the working electrode, and Ag/AgCl as the reference electrode—using a CHI 660E model Electrochemical Workstation (Champaign, IL, USA). Electrodeposition was carried out in an aqueous solution containing various Ni/Mo/Co molar ratios of Ni (NO3)2, Na2MoO2, and Co (NO3)2. The total metal cation concentration was 1.2 mM. The THs were electrodeposited onto the dry Ni foam at a constant current of 0.5 mA/cm2 for 300 s at room temperature, right after soaking into the as-prepared solution for 900 s. The as-obtained coated Ni foam was dried in a vacuum oven for another 24 h. The mass of the active materials deposited on the Ni foam was calculated based on the weight of the sample before and after electrodeposition. Typical material loading was around 0.1 mg.

2.2. Material Characterization

X-ray photoelectron spectroscopy (XPS) spectra were taken using a Perkin-Elmer PHI 570 ESCA/SAM spectrometer. The morphology, surface structure, and metal composition of the as-prepared samples were analyzed by field emission scanning electron microscopy (FE-SEM, JEOL JEM-7000F) and energy-dispersive X-ray spectroscopy (EDS, BRUKER QUANTAX EDS for SEM). Transmission electron microscopy (TEM) and selected area electron diffraction (SAED) patterns were obtained using a JEOL JEM-2010 microscope with a LaB6 Filament gun.

2.3. Electrochemical Measurement

Single-electrode supercapacitor measurements were carried out at room temperature using a three-electrode system with an as-prepared Ni foam substrate deposited with active materials, Pt foil, and Hg/HgO as the working electrode, counter electrode, and reference electrode, respectively. Then, 30 c.c. of 1 M aqueous KOH was used as the electrolyte. The cyclic voltammetry (CV), galvanostatic charge–discharge (GCD), and galvanostatic cycling performance (GCP) were tested using an electrochemical workstation (CHI 660E).

3. Results and Discussion

3.1. Electrochemical Performance of the Supercapacitors (SCs)

The electrochemical performances of Ni-Mo-Co hydroxides with varying Ni/Mo/Co ratios were systematically investigated to elucidate the impact of composition on the properties of Ni-Mo-Co triple hydroxides (THs). Figure 1a depicts the cyclic voltammetry (CV) curves for the Ni-Mo-Co hydroxides within a potential window of −0.05 to 0.55 V (vs. Hg/HgO) at a scan rate of 20 mV s−1. The redox peaks observed for the Ni foam are significantly smaller than those of the as-deposited Ni-Mo-Co hydroxide electrodes, indicating that the capacitance contribution from pure Ni foam is negligible (as demonstrated in Figure 1a). The Ni/Mo/Co (1/1/0) hydroxides exhibit a pair of narrow redox peaks within the potential range from 0.33 to 0.47 V. Notably, the anodic peak intensifies and broadens, indicating a higher capacitance, and shifts towards more negative potentials with increasing Co ratios.
The specific capacitance (Cs) as a function of current density of the Ni-Mo-Co TH nanoflakes were plotted in Figure 1b, calculated based on the discharge process in the chronopotentiometry (CP) measurement, using the following equation [34,56]:
C s = I × t / m × V
where Cs represents the specific capacitance (F g−1), I is the constant discharge current (A), ∆t (s) is the duration of the discharge, m (g) is the mass of active material, and ∆V (V) is the voltage windows in the CP measurements. As expected, the specific capacitance of all samples decreases with the increase in current density. This common pseudocapacitive behavior is due to lower rate of redox reactions at a higher current density as a result of a shorter time for ion diffusion [10,57,58]. The Cs of Ni/Mo/Co (1/0/1), Ni/Mo/Co (1/1/1), Ni/Mo/Co (0/0/1), and Ni/Mo/Co (0/1/1) are 2026.4 F g−1, 2557.2 F g−1, 1232.4 F g−1, and 1124.4 F g−1, respectively, at the current density of 2 A g−1. This suggests the incorporation of Ni lead into a high electrochemical energy storage capacity. Similarly, the incorporation of Mo resulting in a higher specific capacitance, as the Cs Mo containing samples of Ni/Mo/Co (0/1/1), Ni/Mo/Co (1/1/0), and Ni/Mo/Co (1/1/1) are 1124.4 F g−1, 2026.4 F g−1, and 2557.2 F g−1 (Figure 1b), respectively, at the current density of 2 A g−1. On the other hand, the Cs of Ni/Mo/Co (0/0/1), Ni/Mo/Co (1/0/0), and Ni/Mo/Co (1/0/1) are 1133.4 F g−1, 1885.64 F g−1, and 2026.4 F g−1, respectively, at the same current density, which are lower than the Mo-containing metal hydroxides. Among all samples, the Ni/Mo/Co (1/1/0.4) sample shows the highest specific capacitances of 3074 F g−1 at 2 A g−1, and 2429 F g−1 at a highest current of 50 A g−1 (ca. 80% capacitance retention), indicating an excellent rate-performance and capacitance retention. On the other hand, the capacitance retention values of Ni/Mo/Co (1/1/0) and Ni/Mo/Co (1/1/1) are just ca. 72% and ca. 74%, respectively. This excellent electrochemical performance may be attributed to the high electrochemical reversibility of the optimized amount of Co-based hydroxides [59,60]. Moreover, the potential difference of anodic and cathodic peaks for Ni/Mo/Co (1/1/0.4) of ca. 100 mV is smaller than those for Ni/Mo/Co (1/1/0) (ca. 140 mV) and Ni/Mo/Co (1/1/1) (ca. 110 mV) (Figure 1a). This suggests a fairly high electrochemical reversibility of Co from charge to discharge [60]. Ni/Mo/Co (1/1/0.4) shows higher Cs (3074 F g−1 at 2 A g−1) than Ni/Mo/Co (1/1/1) and all other samples, which may be the result of the optimal composition of triple metallic compositions. These results indicate that the tri-metallic hydroxide can further improve the specific capacitance as compared to bi-metallic hydroxide, which may be attributed to the high electroactivity, stability, and excellent reversibility with the addition of Co [10,28,59,60]; fast electron transport, rapid ion diffusion, high electrical conductivity with Mo addition [51,61,62]; and remarkable electrochemical energy storage capacity resulting from Ni [63].
Energy density (E) and power density (P) are both critical factors in evaluating the electrochemical performances of energy storage devices including supercapacitors. Figure 2 presents the calculated gravimetric energy density and power density for the as-prepared hydroxide materials (for a detailed calculation, see ESI†). It illustrates that the energy density of Ni/Mo/Co (1/1/0.4) is higher than that of others at the same current density. The Ni/Mo/Co (1/1/0.4) shows the highest energy density of 464.96 W h kg−1 among all materials at a low current density, with a power density of 0.6 kW kg−1 and energy density of 367.39 W h kg−1, which remains the highest among all materials at a high current density with a high power density of 15.13 kW kg−1, indicating a remarkable electrochemical performance.

3.2. Morphology and Structural Analysis

Ni-Mo-Co triple hydroxide (TH) ultrathin mesoporous nanoflakes (NFs) were electrodeposited on Ni foam in the aqueous solution containing appropriate concentrations of Ni (NO3)2, Na2MoO2, and Co (NO3)2. The possible formation mechanism can be ascribed to the co-deposition of Ni2+, Co2+, and Mo6+ cations under alkaline condition. First, hydroxide ions (OH) are produced on the surface of the cathode via a reduction reaction of NO3− and H2O, as shown in Equations (2) and (3) [64]:
NO3 + 7H2O + 8e → NH4+ + 10OH
2H2O + 2e → H2 + 2OH
Subsequently, the Ni-Mo-Co hydroxide was co-electrodeposited according to Equation (4):
xNi2+ + yCo2+ + zMo6+ + 2(x + y + 3z) OH → NixCoyMoz(OH)2(x+y+3z)
The valence states and surface compositions of the as-prepared samples were characterized using X-ray photoelectron spectroscopy (XPS), as illustrated in Figure 3. The XPS survey spectra (Figure 3a) reveal the presence of Ni, Mo, Co, O, and a minor amount of C. High-resolution XPS spectra for Ni 2p, Mo 3d, Co 2p, and O 1s are presented in Figure 3b–e, respectively. Specifically, the Ni 2p spectrum (Figure 3b) exhibits two prominent peaks with binding energies at 873.5 eV for 2p1/2 and 855.9 eV for 2p3/2, corresponding to Ni2+ in nickel hydroxides [65]. The Mo 3d spectrum (Figure 3c) can be deconvoluted into two principal peaks with binding energies at 232.4 eV for 3d5/2 and 235.5 eV for 3d3/2, which are typically attributed to Mo6+ [66]. Similarly, the high-resolution Co 2p spectrum (Figure 3d) displays two main peaks at 781.6 eV and 797.7 eV, along with shake-up satellite peaks at 787.5 eV and 804.0 eV, indicative of Co2+ in cobalt hydroxides [67,68]. The O 1s XPS spectrum (Figure 3e) features a peak at 531.1 eV, which is attributed to metal hydroxide bonds [69]. The XPS analysis confirms that the as-prepared sample contains Ni2+, Mo6+, and Co2+, indicating the successful synthesis of Ni-Mo-Co ternary hydroxides.
The SEM images (Figure 4a,b) show that Ni-Mo-Co THs, with a Ni/Mo/Co solution concentration ratio of 1/1/0.4, are deposited uniformly onto the macroscopic 3D skeleton of the Ni foam. Figure 4b displays the SEM images recorded at a higher magnification of the Ni foam surface, as marked in Figure 4a. The image shows as-deposited ultrathin films consisting of interconnected Ni-Mo-Co hydroxide platelets, forming a mesoporous 3D nanostructure around 50 nm in length and 5 nm in thickness. These nanostructures tend to be vertically oriented on the surface of Ni foam, forming a nanoporous structure, with pore sizes of about 50–100 nm. Different from pure Ni foam substrate, the 3D structure of the highly rippled ultrathin nanoflakes have abundant open space and a large surface-to-volume ratio, which can provide more electroactive surface sites. These 3D structures allow the effective migration of the electrolyte, and thus enhance mass/charge transfer at the electrode/electrolyte interface [70]. The morphology and structure of the as-deposited composites were further investigated by TEM. Figure 4e shows a panoramic view of the as-deposited Ni-Mo-Co TH nanoflakes. The interconnected mesoporous characteristic with a distinct light/dark contrast suggested the existence of nanoflakes and mesoporous structures with open space. Moreover, the nanoflakes show a rippled flake structure with dimensions around 50–100 nm, in accordance with the SEM results. These nanoflakes are quite transparent, indicating their ultrathin features. Figure 4g–k present the EDS mapping images of the Ni-Mo-Co TH nanoflakes supported on the Ni foam, and Figure 4g is the scanning background of the test area. It clearly reveals that Mo, Ni, and Co are uniformly distributed, suggesting the homogeneous deposition of the Ni-Mo-Co TH nanoflakes onto the skeleton of Ni foam.
The high-resolution TEM image (Figure 4f) shows there is no observable lattice fringes for Ni, Mo, and Co of Ni-Mo-Co hydroxide composites, suggesting that the as-deposited Ni-Mo-Co TH nanoflakes are amorphous. This can be further supported by the electron diffraction (SAED) pattern of the as-prepared nanoflakes (the inset of Figure 4e), which show a broad and diffused halo ring. As compared to crystalline metallic hydroxides, amorphous hydroxides exhibited enhanced electrochemical performance, which may be attributed to the following aspects: 1. The long-range disorder and short-range ordered structure can improve the electronic conductivity of the electrode materials. 2. The easier access of the amorphous structure for intercalation and deintercalation of charges; 3. Adequate defects on the amorphous materials favor the charge-transfer rate [34,71]. Moreover, the hierarchically structured electrode with mesoporous Ni-Mo-Co hydroxides formed on a microporous Ni foam could benefit the ion/mass transfer while charging/discharging [72].

3.3. Electrochemical Test of Ni-Mo-Co (1/1/0.4) THs

The electrochemical properties were measured by CV and CP techniques. Figure 5a presents the CV curves of Ni-Mo-Co (1/1/0.4) TH nanoflake electrodes at various scan rates ranging from 1 mV s−1 to 100 mV s−1. A pair of relatively symmetric, board redox peaks can be clearly observed in all CV curves, which is characteristic of a pseudocapacitive behavior. The redox peaks were attributed to the Faradaic reactions of Equations (5)–(7), respectively.
Ni (OH)2 + OH ↔ NiOOH + H2O + e
Co (OH)2 + OH ↔ CoOOH + H2O + e
CoOOH + OH ↔ CoO2 + H2O + e
At the scan rate of 20 mV s−1, the redox peak potentials are found to be 0.37 V and 0.27 V for anodic and cathodic peaks, respectively. Moreover, all of these CV curves exhibit similar shapes, except for the slight shift in the redox peak position with the increasing scan rates resulting from concentration polarization. These CV curves suggest the good electrochemical reversibility, high-rate performance and good electron conductivity of amorphous Ni-Mo-Co TH nanoflakes. The GCD tests were performed within the potential window of −0.05 to 0.55 V (vs. Hg/HgO) at current densities varying from 2 to 50 A g−1 to evaluate the electrochemical capacitive performance (Figure 5b) using the CP technique. The nonlinear, highly symmetric shape, as well as the clear charge/discharge plateaus of the CP curves, are characteristic of the typical pseudocapacitance behavior [10]. Additionally, even at a high current density of 50 A g−1, the charge/discharge curve still shows an almost symmetrical nature without a clear IR decrease (inset of Figure 5b). This suggests that the amorphous Ni-Mo-Co TH nanoflakes exhibit rapid I-V response characteristics and reversible redox reactions, resulting in good electrochemical reversibility.
Moreover, the linear relationship between the current of the redox peaks in CV curves and the corresponding square root of the scan rate (Figure 5c) strongly suggest that the electrochemical process is a diffusion-controlled process of hydroxyl ions [38].
Figure 5d shows the long-term performance of the Ni/Mo/Co (1/1/0.4) hydroxide nanoflake electrode at a current density of 10 A g−1. Interestingly, the specific capacitance of the Ni-Mo-Co hydroxide gradually increased during the first 1300 cycles, which may be attributed to activation of the electrode [73]. After 5000 charge/discharge cycles, a loss of 3.37% of the specific capacitance was observed as compared to initial capacitance (for a detailed calculation, see ESI†), and the morphology of Ni/Mo/Co (1/1/0.4) remains relatively stable (Figure 4c,d). Thus, the remarkable cycling stability of this electrode can be attributed to the stable electrodeposited active materials on the Ni foam, as well as the resulting good structural stability. Moreover, the Ni-Mo-Co ternary hydroxide electrode shows a high specific capacitance, long cyclability, as well as excellent rate performance, which can also be attributed to the following: (1) the mesoporous hierarchical architectures of Ni-Mo-Co hydroxides nanoflakes [74] and the 3D structure affording a continuous network as well as extra active sites for the redox reaction with charge/discharge curves [26]; (2) the amorphous nature of the as-obtained triple metallic hydroxide with more defects, long-range disorder, as well as short-range order, which help to obtain an excellent electrochemical performance [34]; and (3) the binder-free electrode system that can promote electrolyte penetration into the active material, resulting in the electrons being able to transport at a sufficiently fast rate for high rate capability [70,75]. The mesoporous nanostructure of ternary hydroxides (THs) formed on microporous nickel foam substantially increases the specific surface area, thereby offering a greater number of reactive sites for electrochemical reactions in supercapacitors. Furthermore, nanoflake architecture enhances the mechanical stability of the amorphous Ni-Mo-Co THs. Consequently, the as-prepared samples exhibit significantly improved electrochemical reversibility, high-rate performance, and cycling stability in supercapacitor applications.

4. Conclusions

In summary, amorphous Ni-Mo-Co triple-hydroxide 3D hierarchical structures with ultrathin nanoflakes as building units were deposited on a Ni foam via a facile template-free, and scalable electrodeposition method. This Ni-Mo-Co (1/1/0.4) TH on a Ni foam as a binder-free electrode for SCs exhibited a superior performance, with a high specific capacitance and capacity (3074 F g−1 at 2 A g−1, and 2429 F g−1 at highest current of 50 A g−1), outstanding rate performance (80% capacitance retention from 2 A g−1 to 50 A g−1), and remarkable cyclability (96.63% of capacitance retention after 5000 cycles at current density of 10 A g−1), which could be attributed to the large open spaces between the interconnected mesoporous nanoflakes, abundant electroactive surface sites and facile electron transmission paths. This performance is better than most reported mono-/bi-/tri-metallic hydroxides as electrodes for SCs. The Ni-Mo-Co THs were demonstrated to be promising high-capacitance electrode material with long-term cyclability for SCs.

Author Contributions

Conceptualization, K.Y.S.N. and P.L.; methodology, P.L.; validation, P.L. and Z.W.; formal analysis, Z.W. and P.L.; investigation, Z.T.; resources, P.L.; data curation, P.L.; writing—original draft preparation, P.L.; writing—review and editing, Z.W.; visualization, Z.T.; supervision, K.Y.S.N.; project administration, K.Y.S.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Libich, J.; Máca, J.; Vondrák, J.; Čech, O.; Sedlaříková, M. Supercapacitors: Properties and applications. J. Energy Storage 2018, 17, 224–227. [Google Scholar] [CrossRef]
  2. Conway, B.E.; Birss, V.; Wojtowicz, J. The role and utilization of pseudocapacitance for energy storage by supercapacitors. J. Power Sources 1997, 66, 1–14. [Google Scholar] [CrossRef]
  3. Bhojane, P. Recent advances and fundamentals of Pseudocapacitors: Materials, mechanism, and its understanding. J. Energy Storage 2022, 45, 103654. [Google Scholar] [CrossRef]
  4. Ashritha, M.G.; Hareesh, K. Chapter 9—Electrode materials for EDLC and pseudocapacitors. In Smart Supercapacitors; Hussain, C.M., Ahamed, M.B., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 179–198. [Google Scholar]
  5. Schütter, C.; Ramirez-Castro, C.; Oljaca, M.; Passerini, S.; Winter, M.; Balducci, A. Activated Carbon, Carbon Blacks and Graphene Based Nanoplatelets as Active Materials for Electrochemical Double Layer Capacitors: A Comparative Study. J. Electrochem. Soc. 2015, 162, A44. [Google Scholar] [CrossRef]
  6. Sahoo, R.; Pal, A.; Pal, T. Chapter 19—Noble Metal–Transition Metal Oxides/Hydroxides: Desired Materials for Pseudocapacitor. In Noble Metal-Metal Oxide Hybrid Nanoparticles; Mohapatra, S., Nguyen, T.A., Nguyen-Tri, P., Eds.; Woodhead Publishing: Sawston, UK, 2019; pp. 395–430. [Google Scholar]
  7. Liu, R.; Zhou, A.; Zhang, X.; Mu, J.; Che, H.; Wang, Y.; Wang, T.-T.; Zhang, Z.; Kou, Z. Fundamentals, advances and challenges of transition metal compounds-based supercapacitors. Chem. Eng. J. 2021, 412, 128611. [Google Scholar] [CrossRef]
  8. Ma, Y.; Xie, X.; Yang, W.; Yu, Z.; Sun, X.; Zhang, Y.; Yang, X.; Kimura, H.; Hou, C.; Guo, Z.; et al. Recent advances in transition metal oxides with different dimensions as electrodes for high-performance supercapacitors. Adv. Compos. Hybrid Mater. 2021, 4, 906–924. [Google Scholar] [CrossRef]
  9. Xue, T.; Wang, X.; Lee, J.-M. Dual-template synthesis of Co(OH)2 with mesoporous nanowire structure and its application in supercapacitor. J. Power Sources 2012, 201, 382–386. [Google Scholar] [CrossRef]
  10. Hu, C.-C.; Chen, J.-C.; Chang, K.-H. Cathodic deposition of Ni(OH)2 and Co(OH)2 for asymmetric supercapacitors: Importance of the electrochemical reversibility of redox couples. J. Power Sources 2013, 221, 128–133. [Google Scholar] [CrossRef]
  11. Vialat, P.; Mousty, C.; Taviot-Gueho, C.; Renaudin, G.; Martinez, H.; Dupin, J.-C.; Elkaim, E.; Leroux, F. High-Performing Monometallic Cobalt Layered Double Hydroxide Supercapacitor with Defined Local Structure. Adv. Funct. Mater. 2014, 24, 4831–4842. [Google Scholar] [CrossRef]
  12. Huang, B.; Wang, W.; Pu, T.; Li, J.; Zhu, J.; Zhao, C.; Xie, L.; Chen, L. Two-dimensional porous (Co, Ni)-based monometallic hydroxides and bimetallic layered double hydroxides thin sheets with honeycomb-like nanostructure as positive electrode for high-performance hybrid supercapacitors. J. Colloid Interface Sci. 2018, 532, 630–640. [Google Scholar] [CrossRef]
  13. Miller, J.R.; Simon, P. Electrochemical Capacitors for Energy Management. Science 2008, 321, 651–652. [Google Scholar] [CrossRef]
  14. Li, L.; Liu, B.; Hou, S.; Yang, Q.; Zhu, Z. Preparation of bulk doped NiCo2O4 bimetallic oxide supercapacitor materials by in situ growth method. Inorg. Nano-Met. Chem. 2022, 1–10. [Google Scholar] [CrossRef]
  15. Yin, Z.; Zhang, S.; Chen, Y.; Gao, P.; Zhu, C.; Yang, P.; Qi, L. Hierarchical nanosheet-based NiMoO4 nanotubes: Synthesis and high supercapacitor performance. J. Mater. Chem. A 2015, 3, 739–745. [Google Scholar] [CrossRef]
  16. Yu, X.; Lu, B.; Xu, Z. Super long-life supercapacitors based on the construction of nanohoneycomb-like strongly coupled CoMoO4-3D graphene hybrid electrodes. Adv. Mater. 2014, 26, 1044–1051. [Google Scholar] [CrossRef]
  17. Khadka, A.; Samuel, E.; Joshi, B.; Kim, Y.I.; Aldalbahi, A.; El-Newehy, M.; Lee, H.-S.; Yoon, S.S. Bimetallic CoMoO4 Nanosheets on Freestanding Nanofiber as Wearable Supercapacitors with Long-Term Stability. Int. J. Energy Res. 2023, 2023, 2910207. [Google Scholar] [CrossRef]
  18. Prabakaran, P.; Arumugam, G.; Ramu, P.; Selvaraj, M.; Assiri, M.A.; Rokhum, S.L.; Arjunan, S.; Rajendran, R. Construction of hierarchical MnMoO4 nanostructures on Ni foam for high-performance asymmetric supercapacitors. Surf. Interfaces 2023, 40, 103086. [Google Scholar] [CrossRef]
  19. Kumcham, P.; Sreekanth, T.; Yoo, K.; Kim, J. Surfactant-assisted morphology modification of nanostructured MnMoO4 for high-performance asymmetric supercapacitors. New J. Chem. 2024, 48, 7820–7835. [Google Scholar] [CrossRef]
  20. Li, M.; Meng, Z.; Feng, R.; Zhu, K.; Zhao, F.; Wang, C.; Wang, J.; Wang, L.; Chu, P.K. Fabrication of Bimetallic Oxides (MCo2O4, M=Cu, Mn) on Ordered Microchannel Electro-Conductive Plate for High-Performance Hybrid Supercapacitors. Sustainability 2021, 13, 9896. [Google Scholar] [CrossRef]
  21. Gao, Y.; Xia, Y.; Wan, H.; Xu, X.; Jiang, S. Enhanced cycle performance of hierarchical porous sphere MnCo2O4 for asymmetric supercapacitors. Electrochim. Acta 2019, 301, 294–303. [Google Scholar] [CrossRef]
  22. Bai, X.; Cao, D.; Zhang, H. Simultaneously morphology and phase controlled synthesis of cobalt manganese hydroxides/reduced graphene oxide for high performance supercapacitor electrodes. Ceram. Int. 2020, 46 Pt B, 19135–19145. [Google Scholar] [CrossRef]
  23. Zou, X.; Sun, Q.; Zhang, Y.; Li, G.-D.; Liu, Y.; Wu, Y.; Yang, L.; Zou, X. Ultrafast surface modification of Ni3S2 nanosheet arrays with Ni-Mn bimetallic hydroxides for high-performance supercapacitors. Sci. Rep. 2018, 8, 4478. [Google Scholar] [CrossRef]
  24. Yang, Z.; Wang, X.; Zhang, H.; Yan, S.; Zhang, C.; Liu, S. One-Step Synthesis of Ultrathin NiMn Layered Double Hydroxide Nanosheets for Supercapacitors. ChemElectroChem 2019, 6, 4456–4463. [Google Scholar] [CrossRef]
  25. Ma, K.; Cheng, J.P.; Liu, F.; Zhang, X. Co-Fe layered double hydroxides nanosheets vertically grown on carbon fiber cloth for electrochemical capacitors. J. Alloys Compd. 2016, 679, 277–284. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Li, L.; Su, H.; Huang, W.; Dong, X. Binary metal oxide: Advanced energy storage materials in supercapacitors. J. Mater. Chem. A 2015, 3, 43–59. [Google Scholar] [CrossRef]
  27. Gonçalves, J.M.; da Silva, M.I.; Toma, H.E.; Angnes, L.; Martins, P.R.; Araki, K. Trimetallic oxides/hydroxides as hybrid supercapacitor electrode materials: A review. J. Mater. Chem. A 2020, 8, 10534–10570. [Google Scholar] [CrossRef]
  28. Xing, J.; Wu, S.; Ng, K.S. Electrodeposition of ultrathin nickel–cobalt double hydroxide nanosheets on nickel foam as high-performance supercapacitor electrodes. RSC Adv. 2015, 5, 88780–88786. [Google Scholar] [CrossRef]
  29. Li, J.; Wei, M.; Chu, W.; Wang, N. High-stable α-phase NiCo double hydroxide microspheres via microwave synthesis for supercapacitor electrode materials. Chem. Eng. J. 2017, 316, 277–287. [Google Scholar] [CrossRef]
  30. Wang, A.-L.; Xu, H.; Li, G.-R. NiCoFe Layered Triple Hydroxides with Porous Structures as High-Performance Electrocatalysts for Overall Water Splitting. ACS Energy Lett. 2016, 1, 445–453. [Google Scholar] [CrossRef]
  31. Chen-Wiegart, Y.C.K.; Liu, Z.; Faber, K.T.; Barnett, S.A.; Wang, J. 3D analysis of a LiCoO2–Li(Ni1/3Mn1/3Co1/3)O2 Li-ion battery positive electrode using x-ray nano-tomography. Electrochem. Commun. 2013, 28, 127–130. [Google Scholar] [CrossRef]
  32. Ecker, M.; Nieto, N.; Käbitz, S.; Schmalstieg, J.; Blanke, H.; Warnecke, A.; Sauer, D.U. Calendar and cycle life study of Li(NiMnCo)O2-based 18,650 lithium-ion batteries. J. Power Sources 2014, 248, 839–851. [Google Scholar] [CrossRef]
  33. Li, P. Novel Design and Synthesis of Transition Metal Hydroxides and Oxides for Energy Storage Device Applications; Wayne State University: Detroit, MI, USA, 2016. [Google Scholar]
  34. Li, H.; Gao, Y.; Wang, C.; Yang, G. A Simple Electrochemical Route to Access Amorphous Mixed-Metal Hydroxides for Supercapacitor Electrode Materials. Adv. Energy Mater. 2015, 5, 1401767. [Google Scholar] [CrossRef]
  35. Li, Q.; Xu, Y.; Zheng, S.; Guo, X.; Xue, H.; Pang, H. Recent Progress in Some Amorphous Materials for Supercapacitors. Small 2018, 14, 1800426. [Google Scholar] [CrossRef]
  36. Liu, X.; Wang, J.; Yang, G. Amorphous nickel oxide and crystalline manganese oxide nanocomposite electrode for transparent and flexible supercapacitor. Chem. Eng. J. 2018, 347, 101–110. [Google Scholar] [CrossRef]
  37. Cui, L.-F.; Ruffo, R.; Chan, C.K.; Peng, H.; Cui, Y. Crystalline-Amorphous Core–Shell Silicon Nanowires for High Capacity and High Current Battery Electrodes. Nano Lett. 2009, 9, 491–495. [Google Scholar] [CrossRef]
  38. Niu, L.; Li, Z.; Xu, Y.; Sun, J.; Hong, W.; Liu, X.; Wang, J.; Yang, S. Simple Synthesis of Amorphous NiWO4 Nanostructure and Its Application as a Novel Cathode Material for Asymmetric Supercapacitors. ACS Appl. Mater. Interfaces 2013, 5, 8044–8052. [Google Scholar] [CrossRef]
  39. Wang, G.; Zhang, L.; Zhang, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef]
  40. Zhang, G.; Lou, X.W. General Solution Growth of Mesoporous NiCo2O4 Nanosheets on Various Conductive Substrates as High-Performance Electrodes for Supercapacitors. Adv. Mater. 2013, 25, 976–979. [Google Scholar] [CrossRef]
  41. Wang, Y.; Yin, Z.; Wang, Z.; Li, X.; Guo, H.; Wang, J.; Zhang, D. Facile construction of Co(OH)2@Ni(OH)2 core-shell nanosheets on nickel foam as three dimensional free-standing electrode for supercapacitors. Electrochim. Acta 2019, 293, 40–46. [Google Scholar] [CrossRef]
  42. Meng, X.; Deng, D. Bio-inspired synthesis of α-Ni(OH)2 nanobristles on various substrates and their applications. J. Mater. Chem. A 2016, 4, 6919–6925. [Google Scholar] [CrossRef]
  43. Wei, T.-Y.; Chen, C.-H.; Chien, H.-C.; Lu, S.-Y.; Hu, C.-C. A Cost-Effective Supercapacitor Material of Ultrahigh Specific Capacitances: Spinel Nickel Cobaltite Aerogels from an Epoxide-Driven Sol–Gel Process. Adv. Mater. 2010, 22, 347–351. [Google Scholar] [CrossRef]
  44. Lang, X.; Hirata, A.; Fujita, T.; Chen, M. Nanoporous metal/oxide hybrid electrodes for electrochemical supercapacitors. Nat. Nanotechnol. 2011, 6, 232. [Google Scholar] [CrossRef]
  45. Dong, X.-C.; Xu, H.; Wang, X.-W.; Huang, Y.-X.; Chan-Park, M.B.; Zhang, H.; Wang, L.-H.; Huang, W.; Chen, P. 3D Graphene–Cobalt Oxide Electrode for High-Performance Supercapacitor and Enzymeless Glucose Detection. ACS Nano 2012, 6, 3206–3213. [Google Scholar] [CrossRef]
  46. Xia, X.; Tu, J.; Zhang, Y.; Wang, X.; Gu, C.; Zhao, X.B.; Fan, H.J. High-Quality Metal Oxide Core/Shell Nanowire Arrays on Conductive Substrates for Electrochemical Energy Storage. ACS Nano 2012, 6, 5531–5538. [Google Scholar] [CrossRef]
  47. Pu, J.; Tong, Y.; Wang, S.; Sheng, E.; Wang, Z. Nickel–cobalt hydroxide nanosheets arrays on Ni foam for pseudocapacitor applications. J. Power Sources 2014, 250, 250–256. [Google Scholar] [CrossRef]
  48. Chen, H.; Hu, L.; Chen, M.; Yan, Y.; Wu, L. Nickel–Cobalt Layered Double Hydroxide Nanosheets for High-performance Supercapacitor Electrode Materials. Adv. Funct. Mater. 2014, 24, 934–942. [Google Scholar] [CrossRef]
  49. Guo, D.; Luo, Y.; Yu, X.; Li, Q.; Wang, T. High performance NiMoO4 nanowires supported on carbon cloth as advanced electrodes for symmetric supercapacitors. Nano Energy 2014, 8, 174–182. [Google Scholar] [CrossRef]
  50. Cheng, D.; Yang, Y.; Luo, Y.; Fang, C.; Xiong, J. Growth of Ultrathin Mesoporous Ni-Mo Oxide Nanosheet Arrays on Ni Foam for High-performance Supercapacitor Electrodes. Electrochim. Acta 2015, 176, 1343–1351. [Google Scholar] [CrossRef]
  51. Xiao, K.; Xia, L.; Liu, G.; Wang, S.; Ding, L.-X.; Wang, H. Honeycomb-like NiMoO4 ultrathin nanosheet arrays for high-performance electrochemical energy storage. J. Mater. Chem. A 2015, 3, 6128–6135. [Google Scholar] [CrossRef]
  52. Liu, Y.; Fu, N.; Zhang, G.; Xu, M.; Lu, W.; Zhou, L.; Huang, H. Design of Hierarchical Ni-Co@Ni-Co Layered Double Hydroxide Core–Shell Structured Nanotube Array for High-Performance Flexible All-Solid-State Battery-Type Supercapacitors. Adv. Funct. Mater. 2017, 27, 1605307. [Google Scholar] [CrossRef]
  53. Wu, N.; Low, J.; Liu, T.; Yu, J.; Cao, S. Hierarchical hollow cages of Mn-Co layered double hydroxide as supercapacitor electrode materials. Appl. Surf. Sci. 2017, 413, 35–40. [Google Scholar] [CrossRef]
  54. Liu, H.; Yu, T.; Su, D.; Tang, Z.; Zhang, J.; Liu, Y.; Yuan, A.; Kong, Q. Ultrathin Ni-Al layered double hydroxide nanosheets with enhanced supercapacitor performance. Ceram. Int. 2017, 43, 14395–14400. [Google Scholar] [CrossRef]
  55. He, P.; Huang, Q.; Huang, B.; Chen, T. Controllable synthesis of Ni-Co-Mn multi-component metal oxides with various morphologies for high-performance flexible supercapacitors. RSC Adv. 2017, 7, 24353–24358. [Google Scholar] [CrossRef]
  56. Huang, C.; Hu, Y.; Jiang, S.; Chen, H.C. Amorphous nickel-based hydroxides with different cation substitutions for advanced hybrid supercapacitors. Electrochim. Acta 2019, 325, 134936. [Google Scholar] [CrossRef]
  57. Jiang, Y.; Liu, J. Definitions of Pseudocapacitive Materials: A Brief Review. Energy Environ. Mater. 2019, 2, 30–37. [Google Scholar] [CrossRef]
  58. Choi, C.; Ashby, D.S.; Butts, D.M.; DeBlock, R.H.; Wei, Q.; Lau, J.; Dunn, B. Achieving high energy density and high power density with pseudocapacitive materials. Nat. Rev. Mater. 2020, 5, 5–19. [Google Scholar] [CrossRef]
  59. Song, K.; Li, W.; Xin, J.; Zheng, Y.; Chen, X.; Yang, R.; Lv, W.; Li, Q. Hierarchical porous heterostructured Co(OH)2/CoSe2 nanoarray: A controllable design electrode for advanced asymmetrical supercapacitors. Chem. Eng. J. 2021, 419, 129435. [Google Scholar] [CrossRef]
  60. Zhang, X.; Qu, G.; Wang, Z.; Xiang, G.; Hao, S.; Wang, X.; Xu, X.; Ma, W.; Zhao, G. Hollow polyhedron structure of amorphous Ni-Co-S/Co(OH)2 for high performance supercapacitors. Chin. Chem. Lett. 2021, 32, 2453–2458. [Google Scholar] [CrossRef]
  61. Manibalan, G.; Govindaraj, Y.; Yesuraj, J.; Kuppusami, P.; Murugadoss, G.; Murugavel, R.; Rajesh Kumar, M. Facile synthesis of NiO@Ni(OH)2-α-MoO3 nanocomposite for enhanced solid-state symmetric supercapacitor application. J. Colloid Interface Sci. 2021, 585, 505–518. [Google Scholar] [CrossRef] [PubMed]
  62. Shi, M.; Zhao, M.; Jiao, L.; Su, Z.; Li, M.; Song, X. Novel Mo-doped nickel sulfide thin sheets decorated with Ni–Co layered double hydroxide sheets as an advanced electrode for aqueous asymmetric super-capacitor battery. J. Power Sources 2021, 509, 230333. [Google Scholar] [CrossRef]
  63. Mozaffari, S.A.; Mahmoudi Najafi, S.H.; Norouzi, Z. Hierarchical NiO@Ni(OH)2 nanoarrays as high-performance supercapacitor electrode material. Electrochim. Acta 2021, 368, 137633. [Google Scholar] [CrossRef]
  64. Sasaki, Y.; Yamashita, T. Effect of electrolytic conditions on the deposition of nickel hydroxide. Thin Solid Film. 1998, 334, 117–119. [Google Scholar] [CrossRef]
  65. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  66. Choi, J.G.; Thompson, L.T. XPS study of as-prepared and reduced molybdenum oxides. Appl. Surf. Sci. 1996, 93, 143–149. [Google Scholar] [CrossRef]
  67. Wang, Z.; Zeng, W.; Ng, K.Y.S. Facile Synthesis of CoS Nanoparticles Anchored on the Surface of Functionalized Multiwalled Carbon Nanotubes as Cathode Materials for Advanced Li–S Batteries. Ind. Eng. Chem. Res. 2022, 61, 9322–9330. [Google Scholar] [CrossRef]
  68. Gao, P.; Zeng, Y.; Tang, P.; Wang, Z.; Yang, J.; Hu, A.; Liu, J. Understanding the Synergistic Effects and Structural Evolution of Co(OH)2 and Co3O4 toward Boosting Electrochemical Charge Storage. Adv. Funct. Mater. 2022, 32, 2108644. [Google Scholar] [CrossRef]
  69. Hanawa, T.; Hiromoto, S.; Asami, K. Characterization of the surface oxide film of a Co–Cr–Mo alloy after being located in quasi-biological environments using XPS. Appl. Surf. Sci. 2001, 183, 68–75. [Google Scholar] [CrossRef]
  70. Zhang, G.; Li, W.; Xie, K.; Yu, F.; Huang, H. A One-Step and Binder-Free Method to Fabricate Hierarchical Nickel-Based Supercapacitor Electrodes with Excellent Performance. Adv. Funct. Mater. 2013, 23, 3675–3681. [Google Scholar] [CrossRef]
  71. Yan, S.; Abhilash, K.P.; Tang, L.; Yang, M.; Ma, Y.; Xia, Q.; Guo, Q.; Xia, H. Research Advances of Amorphous Metal Oxides in Electrochemical Energy Storage and Conversion. Small 2019, 15, 1804371. [Google Scholar] [CrossRef] [PubMed]
  72. Xiao, Z.; Mei, Y.; Yuan, S.; Mei, H.; Xu, B.; Bao, Y.; Fan, L.; Kang, W.; Dai, F.; Wang, R.; et al. Controlled Hydrolysis of Metal–Organic Frameworks: Hierarchical Ni/Co-Layered Double Hydroxide Microspheres for High-Performance Supercapacitors. ACS Nano 2019, 13, 7024–7030. [Google Scholar] [CrossRef]
  73. Lu, X.; Liu, T.; Zhai, T.; Wang, G.; Yu, M.; Xie, S.; Ling, Y.; Liang, C.; Tong, Y.; Li, Y. Improving the Cycling Stability of Metal–Nitride Supercapacitor Electrodes with a Thin Carbon Shell. Adv. Energy Mater. 2014, 4, 1300994. [Google Scholar] [CrossRef]
  74. Wu, H.B.; Pang, H.; Lou, X.W. Facile synthesis of mesoporous Ni0.3Co2.7O4 hierarchical structures for high-performance supercapacitors. Energy Environ. Sci. 2013, 6, 3619–3626. [Google Scholar] [CrossRef]
  75. Zhang, G.Q.; Wu, H.B.; Hoster, H.E.; Chan-Park, M.B.; Lou, X.W. Single-crystalline NiCo2O4 nanoneedle arrays grown on conductive substrates as binder-free electrodes for high-performance supercapacitors. Energy Environ. Sci. 2012, 5, 9453–9456. [Google Scholar] [CrossRef]
Figure 1. Electrochemical properties of Ni-Mo-Co hydroxide supercapacitors formed in solution with different Ni/Mo/Co ratios: (a) CV curves at scanning rate of 20 mV/s; (b) evolution of specific capacitance with different current densities.
Figure 1. Electrochemical properties of Ni-Mo-Co hydroxide supercapacitors formed in solution with different Ni/Mo/Co ratios: (a) CV curves at scanning rate of 20 mV/s; (b) evolution of specific capacitance with different current densities.
Energies 17 03881 g001
Figure 2. Ragone chart of Ni-Mo-Co hydroxide formed in different solutions, calculated from GCD data.
Figure 2. Ragone chart of Ni-Mo-Co hydroxide formed in different solutions, calculated from GCD data.
Energies 17 03881 g002
Figure 3. XPS spectra of the Ni-Mo-Co hydroxide formed in Ni/Mo/Co (1/1/0.4) solution: (a) Survey scan, (b) Ni 2p spectrum, (c) Mo 3d spectrum, (d) Co 2p spectrum, (e) O 1s spectrum.
Figure 3. XPS spectra of the Ni-Mo-Co hydroxide formed in Ni/Mo/Co (1/1/0.4) solution: (a) Survey scan, (b) Ni 2p spectrum, (c) Mo 3d spectrum, (d) Co 2p spectrum, (e) O 1s spectrum.
Energies 17 03881 g003
Figure 4. Morphology of Ni-Mo-Co hydroxide formed in Ni/Mo/Co (1/1/0.4) solution: (a) low- and (b) high-magnification images of Ni-Mo-Co TH nanoflakes supported on Ni foam; the morphology and structure of Ni-Mo-Co TH nanoflakes (c) before cycling and (d) after 5000 cycles. (e) Low- and (f) high-resolution TEM images of Ni-Mo-Co TH nanoflakes; the low-left inset in (e) represents the corresponding selected area diffraction pattern. (gk) EDS elemental mapping.
Figure 4. Morphology of Ni-Mo-Co hydroxide formed in Ni/Mo/Co (1/1/0.4) solution: (a) low- and (b) high-magnification images of Ni-Mo-Co TH nanoflakes supported on Ni foam; the morphology and structure of Ni-Mo-Co TH nanoflakes (c) before cycling and (d) after 5000 cycles. (e) Low- and (f) high-resolution TEM images of Ni-Mo-Co TH nanoflakes; the low-left inset in (e) represents the corresponding selected area diffraction pattern. (gk) EDS elemental mapping.
Energies 17 03881 g004
Figure 5. (a) CV curves of the Ni-Mo-Co TH supercapacitor (formed in Ni/Mo/Co (1/1/0.4) solution) at different scan rates from 1 to 100 mV s−1; (b) galvanostatic charge/discharge (GCD) curves of the Ni-Mo-Co THs at different current densities from 2 to 50 A g−1. (c) Correlation between Ip and V1/2 of the Ni-Mo-Co TH nanoflake electrode; (d) galvanostatic cycling performance at a current density of 10 A g−1 for 5000 cycles.
Figure 5. (a) CV curves of the Ni-Mo-Co TH supercapacitor (formed in Ni/Mo/Co (1/1/0.4) solution) at different scan rates from 1 to 100 mV s−1; (b) galvanostatic charge/discharge (GCD) curves of the Ni-Mo-Co THs at different current densities from 2 to 50 A g−1. (c) Correlation between Ip and V1/2 of the Ni-Mo-Co TH nanoflake electrode; (d) galvanostatic cycling performance at a current density of 10 A g−1 for 5000 cycles.
Energies 17 03881 g005
Table 1. Summary of supercapacitance values obtained using nickel, cobalt, or molybdenum hydroxide/oxide nanostructured electrodes synthesized by different methods during the last decade. Note: the most cited paper(s), in each year, was (were) only considered for comparison.
Table 1. Summary of supercapacitance values obtained using nickel, cobalt, or molybdenum hydroxide/oxide nanostructured electrodes synthesized by different methods during the last decade. Note: the most cited paper(s), in each year, was (were) only considered for comparison.
YearChemical
Composition
MorphologySynthesis MethodCapacitance (F g−1) and Current DensityDecay% (after No. of Cycles)Ref.
2019NiCo2O4 aerogelsAerogelEpoxide addition procedure1400Excellent stability (2000)[43]
2010Gold/MnO2NanoporousConventional chemical method~1145N/A[44]
20123D graphene/Co3O4NanowireChemical vapor deposition~1100 at 10 A g−1N/A[45]
2012Co3O4/NiONanowiresTwo-step solution-based method853 at 2 A g−1Excellent stability (6000)[46]
2013Ni-Co hydroxideNanosheetsHydrothermal1734 at 6 A g−114% (1000)[47]
2014Ni-Co hydroxideNanosheetsOne-step process2682 at 3 A g−118% (5000)[48]
2014NiMoO4 NanowiresFacile hydrothermal method1517 at 1.2 A g−14.7% (1000)[49]
2015Ni-Mo metal oxideUltrathin mesoporousConventional chemical method954 at 2 A g−122.3% (5000)[50]
2015NiMoO4Ultrathin nanosheetsElectrodeposition1694 at 1 A g−17.2% (9000)[51]
2017Ni-Co@Ni-Co layered double hydroxideCore–shell structured nanotubeConventional chemical method319 at 2 A g−13.1% (3000)[52]
2017Mn-Co layered double hydroxideHierarchical hollow cagesConventional chemical method511 at 2 A g−110% (2000)[53]
2017Ni-Al nanosheetsNanosheetsConventional chemical method1919 at 2 A g−1Recover to 433 after 3000[54]
2017NiCoMn2 metal oxideVarious morphologiesHydrothermal1434.2 at 2 mA cm−25.7% (3000)[55]
NowNi-Mo-Co ternary hydroxideNanosheets One-step electrodeposition 3074 at 2 A g−13.37% (5000)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Z.; Li, P.; Tang, Z.; Ng, K.Y.S. Novel Design and Synthesis of Ni-Mo-Co Ternary Hydroxides Nanoflakes for Advanced Energy Storage Device Applications. Energies 2024, 17, 3881. https://doi.org/10.3390/en17163881

AMA Style

Wang Z, Li P, Tang Z, Ng KYS. Novel Design and Synthesis of Ni-Mo-Co Ternary Hydroxides Nanoflakes for Advanced Energy Storage Device Applications. Energies. 2024; 17(16):3881. https://doi.org/10.3390/en17163881

Chicago/Turabian Style

Wang, Zhao, Peifeng Li, Zhuolun Tang, and Ka Yuen Simon Ng. 2024. "Novel Design and Synthesis of Ni-Mo-Co Ternary Hydroxides Nanoflakes for Advanced Energy Storage Device Applications" Energies 17, no. 16: 3881. https://doi.org/10.3390/en17163881

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