Next Article in Journal
Recent Advances in Flexible Ultrasonic Transducers: From Materials Optimization to Imaging Applications
Previous Article in Journal
Smart Nematic Liquid Crystal Polymers for Micromachining Advances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Achieving Self-Supported Hierarchical Cu(OH)2/Nickel–Cobalt Sulfide Electrode for Electrochemical Energy Storage

Key Laboratory for Comprehensive Energy Saving of Cold Regions Architecture of Ministry of Education, Jilin Jianzhu University, Changchun 130118, China
*
Authors to whom correspondence should be addressed.
Micromachines 2023, 14(1), 125; https://doi.org/10.3390/mi14010125
Submission received: 19 December 2022 / Revised: 29 December 2022 / Accepted: 30 December 2022 / Published: 2 January 2023

Abstract

:
Herein, nickel–cobalt sulfide (NCS) nanoflakes covering the surface of Cu(OH)2 nanorods were achieved by a facile two-step electrodeposition strategy. The effect of CH4N2S concentration on formation mechanism and electrochemical behavior is investigated and optimized. Thanks to the synergistic effect of the selected composite components, the Cu(OH)2/NCS composite electrode can deliver a high areal specific capacitance (Cs) of 7.80 F cm−2 at 2 mA cm−2 and sustain 5.74 F cm−2 at 40 mA cm−2. In addition, coulombic efficiency was up to 84.30% and cyclic stability remained 82.93% within 5000 cycles at 40 mA cm−2. This innovative work provides an effective strategy for the design and construction of hierarchical composite electrodes for the development of energy storage devices.

1. Introduction

In recent years, transition metal sulfides (TMS) as electrode materials have continued to attract attention in the field of energy storage [1]. Among them, nickel sulfide and cobalt sulfide have special advantages in redox activity and theoretical specific capacitance (Cs). At the same time, they exhibit controllable morphology, abundant reserves, and environmental friendliness [2]. Various synthetic pathways have been explored to regulate their morphology and structure to achieve performance optimization, including hydrothermal, ion exchange, and electrodeposition methods [3,4,5]. Among them, electrodeposition technology has attracted continuous attention due to its simple operation, short experimental cycle, and easy realization of effective regulation of deposition amount [6,7]. In particular, the in situ electrodeposition based on the electrode substrate allows the binder-free electrode to be measured directly [8].
At the same time, researchers also optimized electrode performance by constructing nickel (cobalt) sulfide-based composite electrodes, including designing a bimetallic sulfide electrode. For example, He et al. fabricated hierarchical Ni-Co-S (NCS) electrodes using an electrodeposition method, and the Cs of the optimized electrode reached 640 F g−1 at 1 A g−1 and cyclic stability remained 84% within 10,000 cycles. The outstanding energy storage behavior is attributed to the hollow nanosphere structure, which is composed of interconnected NCS nanosheets and thus provides a smooth path for ion transport [9]. In another example, Zheng et al. reported a self-supported hierarchical NCS nanosheet by a two-step hydrothermal that exhibits ultra-high Cs [3]. In fact, the Cs value depends on the electrochemical active site of the constructed NCS electrode. Therefore, optimizing energy storage properties could be achieved by increasing the quality of deposited NCS electrode material [10,11]. However, with the continuous increase of the amount of deposition layer, electrode materials are prone to excessive disorderly accumulation, and even structure collapse and shedding may occur during the electrochemical test, which will inevitably lead to the attenuation of Cs and cyclic stability of the electrode [2,9,12].
To address the above issues, the construction of NSC-based composite electrodes has proved to be an effective strategy [13,14]. For instance, Luo et al. grew NCS on carbon nanotubes by combining chemical vapor deposition with electrodeposition. The introduction of carbon nanotubes, as scaffolds, provides spatial support for the growth of NCS and improves the conductivity of the composite electrode, which further provides an excellent passage for electron transport, and thus dramatically enhances Cs [11].
Inspired by this, we designed and constructed a Cu(OH)2/NCS composition electrode using a facile two-step electrodeposition. First, copper foam (CF) was selected as the only copper source for surface oxidation to generate Cu(OH)2 nanorod arrays, and the electrodeposition of NCS on its surface was then completed. The morphology and properties of Cu(OH)2/NCS electrode were studied by adjusting CH4N2S concentration. The advantage of this in situ electrodeposition strategy is that the electrode to be measured is directly obtained, and this allows each component to form close contact, reduces contact resistance, and thus optimizes the electron transmission path. This is especially true for Cu(OH)2 nanorods introduced by electrodeposition, which provides sufficient space for NCS deposition, and thus endows the hierarchical Cu(OH)2/NCS electrode with more abundant electrochemical active sites and outstanding energy storage characteristics.

2. Materials and Methods

2.1. Materials and Reagents

Cut the commercial copper foam (CF) according to the specified size (1 cm × 1.5 cm), and then use water, ethanol and hydrochloric acid to complete the pretreatment, and vacuum dry for standby. The reagents, including NiCl2·6H2O, Co(NO3)2·6H2O, CH4N2S, and NaOH are exploited without further purification.

2.2. Electrodeposition of Hierarchical Cu(OH)2/NCS on CF

Galvanostatic deposition of Cu(OH)2 is achieved at 0.05 A for 300 s through a three-electrode system consisting of CF (working electrode), saturated calomel electrode (SCE, reference electrode), and Pt plate (counter electrode). A 2 M NaOH solution is used as electrolyte.
A similar three-electrode system, with the working electrode replaced by the previously generated Cu(OH)2/CF, is used to achieve potentiostatic deposition of the NCS layer at −1.1 V for 1200 s. The electrolyte is composed of 0.05 M NiCl2·6H2O, 0.05 M Co(NO3)2·6H2O and 0.75 M CH4N2S. For comparison, potentiostatic deposition of NCS was performed for CH4N2S at different concentrations, including 0.05 M (abbreviated as S-1), 0.25 M (S-2), 0.5 M (S-3), 0.75 M (S-4), and 1 M (S-5).

2.3. Characterization

Samples were evaluated by XRD (Cu Kα radiation with λ = 1.5406 Å), FE-SEM (JSM-7610F), and XPS (ESCALAB 250Xi). Electrochemical characteristics, including cyclic voltammetry (CV), galvanostatic charge-discharge (GCD), and cycling performance, were determined by an electrochemical workstation (Chenhua, CHI 760E). Cu(OH)2/NCS/CF (working electrode), Ag/AgCl (reference electrode), and Pt plate (counter electrode) in 2 M NaOH solution constituted the three-electrode test system.

3. Results

Figure 1 is a schematic diagram of the synthesis path of the Cu(OH)2/NCS composite electrode by a facile two-step electrodeposition. The first step is to achieve surface oxidation to generate blue-green Cu(OH)2 nanorods on the orange-red 3D CF substrate by galvanostatic deposition, and then to achieve the encapsulation of Cu(OH)2 nanorods by NCS through potentiostatic deposition, thus forming the hierarchical Cu(OH)2/NCS composite electrode.
Figure 2a illustrates the XRD spectra of Cu(OH)2 and Cu(OH)2/NCS. Two obvious strong peaks marked by a triangle are ascribed to the CF substrate (JCPDS No.01-1241), while the other seven peaks marked by a square are indexed to (020), (021), (002), (111), (041), (130), and (150) planes of orthorhombic Cu(OH)2 (JCPDS No.13-0420). No significant additional diffraction peaks are obtained in the XRD pattern of NCS due to the weak crystallinity of the NCS layer. This situation has been mentioned and explained in previous reports [4,15,16]. In addition, the XPS survey spectra of Cu(OH)2/NCS confirmed the presence of Cu, Ni, Co, O and S elements in the sample. Figure 2b records the spectrum of Cu 2p; two main peaks at 954.50 and 934.70 eV could be assigned to Cu 2p1/2 and Cu 2p3/2, while two separated sharp diffraction peaks at 952.45 and 932.50 eV originate from the CF substrate. The rests at 963.04, 944.80, and 941.94 eV are their satellite peaks [17,18]. The Ni 2p spectrum is depicted in Figure 2c. Two intense feature peaks at 873.80 and 856.20 eV are ascribed to Ni 2p1/2 and Ni 2p3/2, implying the presence of Ni2+. The other two peaks located at 880.00 and 861.80 eV are their corresponding satellite peaks [5,19]. In the Co 2p spectrum (Figure 2d), spin-orbit double peaks appeared at 797.60 and 796.40 eV indexed to Co 2p1/2, while those at 782.10 and 780.80 eV belong to Co 2p3/2, accompanied by two satellite peaks around 803.55 and 786.85 eV, confirming the coexistence of Co2+ and Co3+ [4,11]. For the O 1s spectrum (Figure 2e), the only characteristic peak at 531.20 eV is indexed to the characteristic signatures of Cu(OH)2 [8]. With respect to the S 2p spectrum in Figure 2f, the main peak decomposed into two peaks at 161.75 and 163.00 eV, corresponding to S 2p1/2 and S 2p3/2, respectively, and derived from S2-. At the same time, a weak broad peak at 167.40 eV could be attributed to the S-O bond. Oxygen is derived from OH- generated by hydrolysis of CH4N2S, which is in good agreement with previous reports [19,20].
Figure 3a–c shows SEM images of Cu(OH)2 grown on CF by galvanostatic deposition at different magnifications. These Cu(OH)2 nanorods are densely and evenly coated on the surface of CF with a diameter of ca. 125 nm and a smooth surface. The corresponding Cu(OH)2/NCS formed by further potentiostatic deposition treatment is shown in Figure 3d–f. These Cu(OH)2 nanorods become fluffy due to the uniform coverage of NCS. In fact, these nanorods are tightly wrapped and wound by large curved nanoflakes with a thickness of ca. 8 nm. Therefore, the top of the nanorods form a flower-like structure.
The electrochemical behaviors of the three electrodes were further tested and compared, including Cu(OH)2/NCS, NCS, and Cu(OH)2 electrode. Figure 4a presents the CV curves of the three electrodes at a scan rate of 5 mV s−1. It is evident that the Cu(OH)2/NCS electrode was observed to have a maximum current response. According to the rule, under the same scan rate, the larger the area enclosed by the CV curve, the higher the Cs [10]. Therefore, the Cu(OH)2/NCS electrode reflects the relatively optimal energy storage property, which is attributed to the following pseudocapacitance reactions [4,6]:
NiS   +   O H NiSOH   +   e
CoS   +   O H CoSOH   +   e
CoSOH   +   O H CoSO   +   H 2 O   +   e
The inference can also be reflected in the GCD curves of the three electrodes. As displayed in Figure 4b, the Cu(OH)2/NCS electrode obviously has the longest discharge time at a discharge current density of 2 mA cm−2. Figure 4c lists a comparison diagram of the Cs of the three electrodes at different discharge current densities. It also confirms that the Cu(OH)2/NCS electrode has the largest Cs, followed by the NCS electrode, while Cu(OH)2 is the lowest. Therefore, the introduction of Cu(OH)2 is a feasible strategy to significantly enhance the energy storage properties of the NCS electrode. In addition, the deposition amount of NCS was regulated by changing the CH4N2S concentration to optimize the performance of Cu(OH)2/NCS composite electrode.
Figure 5 displays the SEM image of the Cu(OH)2/NCS electrode deposits at different CH4N2S concentrations. When the concentration of CH4N2S was low (0.05 M, S-1), the surface of these Cu(OH)2 nanorods began to become rough, indicating a preliminary small deposition of NCS (Figure 5a,b). With the increase in CH4N2S concentration (0.25 M, S-2), it is evident that there are curved nanoflakes produced on the surface of these Cu(OH)2 nanorods, with a thickness of ca. 8 nm (Figure 5c,d). When the CH4N2S concentration increases to 0.5 M (S-3), these nanoflakes are connected to each other and arranged in rows, like ridges growing from nanorods (Figure 5e,f). When the concentration of CH4N2S reaches 0.75 M (S-4), these ridges gradually expand and rise, forming large undulating nanoflakes to wrap these Cu(OH)2 nanorods (Figure 5g,h). The concentration of CH4N2S continued to increase excessively (1 M, S-5), resulting in fragmentation and disordered accumulation of these nanoflakes (Figure 5i,j).
Figure 6 compares the electrochemical properties of the five electrodes. As seen, the GCD curves at current densities of 2 mA cm−2 (Figure 6a); the discharge time length sequence of the five electrodes is S-4 > S-3 > S-5 > S-2 > S-1, which also reflects the order of their Cs. Further, their detailed values of Cs corresponding to different current densities are shown in Figure 6b. It is clear that S-4 has the highest Cs. As a further derivation, when the current density is increased by 20 times, S-4 also reflects a relatively optimal rate capability of 73% (Figure 6c). Figure 6d displays the voltage drop curves of the five electrodes at different current densities. Average RESR data derived from these are shown in Figure 6e. S-4 has a minimum RESR of 1.31 Ω cm−2 in terms of the formula inserted.
According to the above comparative analysis, it can be inferred that in the case of Cu(OH)2 nanorods providing skeleton support, the Cs value increases with the continuous deposition of NCS. At the beginning of the electrodeposition reaction, the S2− produced by the gradual hydrolysis of CH4N2S is captured by Ni2+ and Co2+ ions, the formed NCS precipitated on the surface of Cu(OH)2 nanorods and acts as growth points [9]. The increase of CH4N2S concentration makes it provide more S2−, and thus the area of the deposited nanoflakes gradually increases; that is, the active area of the electrode increases [21]. However, the deposition of NCS nanoflakes is based on the premise that Cu(OH)2 nanorods provide skeleton support. When the amount of S2− is too large, it is bound to collapse Cu(OH)2 nanorods and lead to a disordered accumulation of the NCS deposition layer [2,11]. In this case, excessive accumulation of the NCS layer leads to an increase in the average RESR. Moreover, it is easy to fall off during electrochemical testing, resulting in reduced electrode rate capability [9]. Therefore, S-4 is considered an optimized structure for further consideration and systematic investigation.
Figure 7a reveals the CV curves of the Cu(OH)2/NCS electrode (S-4) scan from 2 to 50 mV s−1. The increase of scan rate is accompanied by a gradual increase of the area enclosed by CV curves, while Cs decrease gradually [17]. The smaller the scanning rate, the more the sufficient pseudocapacitance reaction occurs between OH from the electrolyte and Cu(OH)2/NCS electrode [11]. The GCD curves of the corresponding electrode at a discharge current density of 2–40 mA cm−2 are shown in Figure 7b. It is obvious that the smaller the current density, the longer the discharge time, and thus the greater the Cs. Detailed Cs values at different current densities are shown in Figure 7c. When the current density is 2 mA cm−2, the Cs is up to 7.80 F cm−2, and the coulombic efficiency is up to 84.30% according to the formula [22,23]. Furthermore, the Cs value remains 73.60% of the original when the current density increases by 20 times and reaches 40 mA cm−2. The average RESR derived from the voltage drop curve in Figure 7d is 1.31 Ω cm−2, which has been compared and discussed in Figure 6e. In addition, data from individual Cu(OH)2 and NCS electrodes were also processed (Figure 8), including CV curves, GCD curves at various current densities, and their corresponding Cs.
The Cu(OH)2/NCS electrode was further tested for long-term cycling stability in a three-electrode system at 40 mA cm−2 (Figure 9). The Cs reached 92.68% of the original value at 2000 cycles, and still retained 82.93% within 5000 cycles.
Table 1 lists the relevant Cu(OH)2 and TMS electrodes reported in recent literature.
The relatively superior properties of the Cu(OH)2/NCS electrode are ascribed to the following: (1) CF acts as the electrode substrate, which not only reflects the advantages of good electrical conductivity and 3D porous structure, but also serves as the only source of Cu in Cu(OH)2. The oxidation reaction achieved by the in situ electrodeposition is bound to result in uniform and dense growth of Cu(OH)2. In particular, the binder-free electrode to be measured can be obtained directly [7,10,17]; (2) the Cu(OH)2 nanorod arrays obtained, as skeleton support, provide sufficient growth space for deposition of the NCS layer, and the stretched NCS nanoflakes further expand the contact area between the Cu(OH)2/NCS electrode and the NaOH electrolyte, thereby creating more abundant electrochemical active sites [6,7,11]; (3) this hierarchical structure design provides a smoother path for charge transmission and thus effectively promotes the redox reaction process, while giving full play to the advantages of electrodeposition and the synergistic effect of each component, and obtaining excellent energy storage performance [21,28,30].

4. Conclusions

In summary, an electrodeposition path is presented to design a hierarchical Cu(OH)2/NCS composite electrode on CF. CF substrate acts as copper source and oxidizes to form Cu(OH)2 nanorods. The deposition parameters of NCS were further optimized by adjusting CH4N2S concentration. Benefiting from the synergistic effect of the selected composite components, the Cu(OH)2/NCS electrode delivers a high Cs of 7.80 F cm−2 at 2 mA cm−2 and retains 5.74 F cm−2 at 40 mA cm−2, and excellent cyclic stability of 82.93% after 5000 cycles at 40 mA cm−2. Furthermore, the hierarchical structure design implemented by the in situ electrodeposition strategy provides an effective route for the construction of Cu(OH)2- or NCS-based composite electrodes for energy storage.

Author Contributions

Conceptualization, H.W.; methodology, P.G.; software, B.W. and J.Y.; formal analysis, X.C.; investigation, S.L. and Z.C.; writing—original draft preparation, W.S.; writing—review and editing, S.L. and X.Y.; supervision, C.W.; funding acquisition, Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The Scientific and Technology Development Project of Jilin Province, China (Grant No. 20200401012GX).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rehman, J.; Eid, K.; Ali, R.; Fan, X.F.; Murtaza, G.; Faizan, M.; Laref, A.; Zheng, W.T.; Varma, G.S. Engineering of Transition Metal Sulfide Nanostructures as Efficient Electrodes for High-Performance Supercapacitors. ACS Appl. Energy Mater. 2022, 5, 6481–6498. [Google Scholar] [CrossRef]
  2. Miao, Y.D.; Zhang, X.P.; Zhan, J.; Sui, Y.W.; Qi, J.Q.; Wei, F.X.; Meng, Q.K.; He, Y.Z.; Ren, Y.J.; Zhan, Z.Z.; et al. Hierarchical NiS@CoS with Controllable Core-Shell Structure by Two-Step Strategy for Supercapacitor Electrodes. Adv. Mater. Interfaces 2019, 7, 1901618. [Google Scholar] [CrossRef]
  3. Zheng, L.X.; Song, J.L.; Ye, X.Y.; Wang, Y.Z.; Shi, X.W.; Huajun Zheng, H.J. Construction of self-supported hierarchical NiCo-S nanosheet arrays for supercapacitors with ultrahigh specific capacitance. Nanoscale 2020, 12, 13811–13821. [Google Scholar] [CrossRef] [PubMed]
  4. Ma, F.; Dai, X.Q.; Jin, J.; Tie, N.; Dai, Y.T. Hierarchical core-shell hollow CoMoS4@Ni-Co-S nanotubes hybrid arrays as advanced electrode material for supercapacitors. Electrochim. Acta 2020, 331, 135459. [Google Scholar] [CrossRef]
  5. Tauquir, S.M.; Karnan, M.; Subramani, K.; Sathish, M. One-step superficial electrodeposition of nickel-cobalt-sulfide for high-energy hybrid asymmetric supercapacitor. Mater. Lett. 2022, 323, 132563. [Google Scholar] [CrossRef]
  6. 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]
  7. Wang, Y.; Yin, Z.L.; Wang, Z.X.; Li, X.H.; Guo, H.J.; Wang, J.X.; Zhang, D.C. 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]
  8. He, D.; Wan, G.D.; Liu, G.L.; Bai, J.H.; Suo, H.; Zhao, C. Facile route to achieve mesoporous Cu(OH)2 nanorods on copper foam for high-performance supercapacitor electrode. J. Alloys Compd. 2017, 699, 706–712. [Google Scholar] [CrossRef]
  9. Wen, Y.X.; Liu, Y.P.; Wang, T.; Wang, Z.L.; Zhang, Y.Z.; Wu, X.G.; Chen, X.T.; Peng, S.L.; He, D.Y. High-Mass-Loading Ni-Co-S Electrodes with Unfading Electrochemical Performance for Supercapacitors. ACS Appl. Energy Mater. 2021, 4, 6531–6541. [Google Scholar] [CrossRef]
  10. Wang, J.S.; Hu, L.B.; Zhou, X.Y.; Zhang, S.; Qiao, Q.S.; Xu, L.; Tang, S.C. Three-Dimensional Porous Network Electrodes with Cu(OH)2 Nanosheet/Ni3S2 Nanowire 2D/1D Heterostructures for Remarkably Cycle-Stable Supercapacitors. ACS Omega 2021, 6, 34276–34285. [Google Scholar] [CrossRef]
  11. Wang, X.M.; Deng, Y.Y.; Wang, Z.H.; Li, Z.W.; Wang, L.C.; Ouyang, J.; Zhou, C.; Luo, Y.F. Sophora-like Nickel-Cobalt Sulfide and Carbon Nanotube Composites in Carbonized Wood Slice Electrodes for All-Solid-State Supercapacitors. ACS Appl. Energy Mater. 2022, 5, 7400–7407. [Google Scholar] [CrossRef]
  12. Maile, N.C.; Shinde, S.K.; Koli, R.R.; Fulari, A.V.; Kim, D.Y.; Fulari, V.J. Effect of different electrolytes and deposition time on the supercapacitor properties of nanoflflake-like Co(OH)2 electrodes. Ultrason. Sonochem. 2019, 51, 49–57. [Google Scholar] [CrossRef] [PubMed]
  13. Gao, Y.; Zhao, L.J. Review on recent advances in nanostructured transition-metal-sulfide-based electrode materials for cathode materials of asymmetric supercapacitors. Chem. Eng. J. 2022, 430, 132745–132763. [Google Scholar] [CrossRef]
  14. Miller, E.E.; Hua, Y.; Tezel, F.H. Materials for energy storage: Review of electrode materials and methods of increasing capacitance for supercapacitors. J. Energy Storage 2018, 20, 30–40. [Google Scholar] [CrossRef]
  15. Yu, K.; Tang, W.M.; Dai, J.Y. Double-Layer MnCo2S4@Ni-Co-S Core/Shell Nanostructure on Nickel Foam for High-Performance Supercapacitor. Phys. Status Solidi A 2018, 215, 1800147. [Google Scholar]
  16. Dai, M.Z.; Zhao, D.P.; Liu, H.Q.; Zhu, X.F.; Wu, X.; Wang, B. Nanohybridization of Ni-Co-S Nanosheets with ZnCo2O4 Nanowires as Supercapacitor Electrodes with Long Cycling Stabilities. ACS Appl. Energy Mater. 2021, 4, 2637–2643. [Google Scholar]
  17. Satpathy, B.K.; Patnaik, S.; Pradhan, D. Room-Temperature Growth of Co(OH)2 Nanosheets on Nanobelt-like Cu(OH)2 Arrays for a Binder-Free High-Performance All-Solid-State Supercapacitor. ACS Appl. Energy Mater. 2022, 1, 77–87. [Google Scholar] [CrossRef]
  18. Tang, S.S.; Li, X.G.; Courté, M.; Peng, J.J.; Fichou, D. Hierarchical Cu(OH)2@Co(OH)2 Nanotrees for Water Oxidation Electrolysis. ChemCatChem 2020, 12, 4038–4043. [Google Scholar] [CrossRef]
  19. Pu, X.L.; Ren, X.H.; Yin, H.F.; Tang, Y.; Yuan, H.D. One-step electrodeposition strategy for growing nickel cobalt hydroxysulfide nanosheets for supercapacitor application. J. Alloys Compd. 2021, 865, 158736. [Google Scholar] [CrossRef]
  20. Fu, S.Q.; Yang, X.C.; Zhao, P.D.; Yao, X.; Jiao, Z.; Cheng, L.L. Regulable Electron Transfer on ZnS/CoS2/CC Prepared by an MOF-on-MOF Strategy for Robust LIB Performance. ACS Appl. Energy Mater. 2022, 5, 5159–5169. [Google Scholar] [CrossRef]
  21. Lin, Y.F.; Chen, X.Y.; Chang, P.; Liu, Z.L.; Ren, G.H.; Tao, J.G. Hierarchical design of Ni3S2@Co9S8 nanotubes for supercapacitors with long cycle-life and high energy density. J. Alloys Compd. 2022, 900, 163503. [Google Scholar] [CrossRef]
  22. Li, R.; Wang, S.L.; Huang, Z.C.; Lu, F.X.; He, T.B. NiCo2S4@Co(OH)2 core-shell nanotube arrays in situ grown on Ni foam for high performances asymmetric supercapcitors. J. Power Sources 2016, 312, 156–164. [Google Scholar] [CrossRef]
  23. Liu, G.L.; Zhao, C.; Liu, T.Y.; He, D.; Suo, H. Facile route to achieve book-like tricobalt tetraoxide microstructures on copper foam for high performance supercapacitor. Mater. Lett. 2018, 220, 78–81. [Google Scholar] [CrossRef]
  24. Zhu, D.; Yan, M.L.; Chen, R.R.; Liu, Q.; Liu, J.Y.; Yu, J.; Zhang, H.S.; Zhang, M.L.; Liu, P.L.; Junqing Li, J.Q.; et al. 3D Cu(OH)2 nanowires/carbon cloth for flexible supercapacitors with outstanding cycle stability. Chem. Eng. J. 2019, 371, 348–355. [Google Scholar] [CrossRef]
  25. Meng, Y.; Sun, P.X.; He, W.D.; Teng, B.; Xu, X.J. Construction of Hierarchical Co-Ni-S Nanosheets as Free-Standing Electrode for Superior-Performance Asymmetric Supercapacitors. Appl. Surf. Sci. 2019, 470, 792–799. [Google Scholar] [CrossRef]
  26. Wang, H.N.; Yan, G.W.; Cao, X.Y.; Liu, Y.; Zhong, Y.X.; Cui, L.; Liu, J.Q. Hierarchical Cu(OH)2@MnO2 core-shell nanorods array in situ generated on three-dimensional copper foam for high-performance supercapacitors. J. Colloid Interface Sci. 2020, 563, 394–404. [Google Scholar] [CrossRef]
  27. Yu, Y.W.; Wang, H.M.; Zhang, H.Z.; Tan, Y.R.; Wang, Y.L.; Song, K.F.; Yang, B.Q.; Yuan, L.F.; Shen, X.D.; Hu, X.L. Blanket-like Co(OH)2/CoOOH/Co3O4/Cu(OH)2 composites on Cu foam for hybrid supercapacitor. Electrochim. Acta 2020, 334, 135559. [Google Scholar] [CrossRef]
  28. Fu, H.C.; Zhang, A.T.; Jin, F.H.; Guo, H.W.; Huang, W.J.; Cheng, W.T.; Jingquan Liu, J.Q. Origami and layered-shaped ZnNiFe-LDH synthesized on Cu(OH)2 nanorods array to enhance the energy storage capability. J. Colloid Interface Sci. 2022, 607, 1269–1279. [Google Scholar] [CrossRef]
  29. Zhang, G.L.; Zhu, R.F.; Zhang, R.Q.; Zhang, D.; Sun, C.; Long, Z.; Li, Y.N. 3D hetero-nanostructured electrode constructed on carbon fiber paper with 2D 1T-MoS2/1D Cu(OH)2 for flexible asymmetric solid-state supercapacitors. J. Power Sources 2022, 523, 231031. [Google Scholar] [CrossRef]
  30. Hu, X.M.; Liu, S.C.; Chen, Y.K.; Jiang, J.B.; Cong, H.S.; Tang, J.B.; Sun, Y.X.; Han, S.; Lin, H.L. Rational design of flower-like cobalt-manganese-sulfide nanosheets for high performance supercapacitor electrode materials. New J. Chem. 2020, 44, 11786–11795. [Google Scholar] [CrossRef]
Figure 1. Synthesis path of hierarchical Cu(OH)2/NCS electrode.
Figure 1. Synthesis path of hierarchical Cu(OH)2/NCS electrode.
Micromachines 14 00125 g001
Figure 2. XRD patterns of Cu(OH)2 and Cu(OH)2/NCS (a). XPS spectra of Cu 2p (b), Ni 2p (c), Co 2p (d), O 1s (e), and S 2p (f).
Figure 2. XRD patterns of Cu(OH)2 and Cu(OH)2/NCS (a). XPS spectra of Cu 2p (b), Ni 2p (c), Co 2p (d), O 1s (e), and S 2p (f).
Micromachines 14 00125 g002
Figure 3. FE-SEM images of Cu(OH)2 (ac) and Cu(OH)2/NCS (bf) at different magnifications.
Figure 3. FE-SEM images of Cu(OH)2 (ac) and Cu(OH)2/NCS (bf) at different magnifications.
Micromachines 14 00125 g003
Figure 4. Performances comparison of Cu(OH)2/NCS, NCS, and Cu(OH)2 electrode: (a) CV curves; (b) GCD curves at different current densities and corresponding Cs (c).
Figure 4. Performances comparison of Cu(OH)2/NCS, NCS, and Cu(OH)2 electrode: (a) CV curves; (b) GCD curves at different current densities and corresponding Cs (c).
Micromachines 14 00125 g004
Figure 5. FE-SEM images of NCS electrodeposited on Cu(OH)2 nanorods for different CH4N2S concentrations: (a,b) 0.05 M, S-1; (c,d) 0.25 M, S-2; (e,f) 0.5 M, S-3; (g,h) 0.75 M, S-4; (i,j) 1 M, S-5.
Figure 5. FE-SEM images of NCS electrodeposited on Cu(OH)2 nanorods for different CH4N2S concentrations: (a,b) 0.05 M, S-1; (c,d) 0.25 M, S-2; (e,f) 0.5 M, S-3; (g,h) 0.75 M, S-4; (i,j) 1 M, S-5.
Micromachines 14 00125 g005
Figure 6. Performances comparison of S-1, 2, 3, 4 and 5: (a) GCD curves at different current densities, (b) corresponding Cs and rate capability (c), (d) voltage drops and corresponding average RESR (e).
Figure 6. Performances comparison of S-1, 2, 3, 4 and 5: (a) GCD curves at different current densities, (b) corresponding Cs and rate capability (c), (d) voltage drops and corresponding average RESR (e).
Micromachines 14 00125 g006
Figure 7. Electrochemical properties of Cu(OH)2/NCS electrode (S-4): (a) CV curves; (b) GCD curves at different current densities and corresponding Cs (c) and the voltage drops (d).
Figure 7. Electrochemical properties of Cu(OH)2/NCS electrode (S-4): (a) CV curves; (b) GCD curves at different current densities and corresponding Cs (c) and the voltage drops (d).
Micromachines 14 00125 g007
Figure 8. CV curves, GCD curves at different current densities and corresponding Cs of Cu(OH)2 (a,c,e) and NCS electrode (b,d,f).
Figure 8. CV curves, GCD curves at different current densities and corresponding Cs of Cu(OH)2 (a,c,e) and NCS electrode (b,d,f).
Micromachines 14 00125 g008
Figure 9. Cyclic stability of the Cu(OH)2/NCS electrode.
Figure 9. Cyclic stability of the Cu(OH)2/NCS electrode.
Micromachines 14 00125 g009
Table 1. Comparison of Cs between Cu(OH)2/NCS and literature reports.
Table 1. Comparison of Cs between Cu(OH)2/NCS and literature reports.
Active MaterialSubstrateElectrolyteCurrent Density
(mA cm−2)
Cs
(F cm−2)
Refs.
Cu(OH)2Carbon cloth1 M NaOH10.24[24]
Cu(OH)2Cu foam5 M NaOH21.75[8]
Co-Ni-SNi foam3 M KOH16.30[25]
Cu(OH)2@MnO2Cu foam6 M KOH2 0.71 [26]
Co(OH)2/CoOOH/
Co3O4/Cu(OH)2
Cu foam1 M KOH1 1.94 [27]
ZnNiFe-LDH/Cu(OH)2Cu foam6 M KOH3 6.1 [28]
MoS2/Cu(OH)2Carbon fiber paper6 M KOH1 1.12 [29]
NiMoO4@Ni-Co-SNi foam2 M KOH5 2.27 [6]
Cu(OH)2/Ni-Co-SCu foam2 M NaOH2 7.80this work
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

Lv, S.; Shang, W.; Chi, Y.; Wang, H.; Chu, X.; Wu, B.; Geng, P.; Wang, C.; Yang, J.; Cheng, Z.; et al. Achieving Self-Supported Hierarchical Cu(OH)2/Nickel–Cobalt Sulfide Electrode for Electrochemical Energy Storage. Micromachines 2023, 14, 125. https://doi.org/10.3390/mi14010125

AMA Style

Lv S, Shang W, Chi Y, Wang H, Chu X, Wu B, Geng P, Wang C, Yang J, Cheng Z, et al. Achieving Self-Supported Hierarchical Cu(OH)2/Nickel–Cobalt Sulfide Electrode for Electrochemical Energy Storage. Micromachines. 2023; 14(1):125. https://doi.org/10.3390/mi14010125

Chicago/Turabian Style

Lv, Sa, Wenshi Shang, Yaodan Chi, Huan Wang, Xuefeng Chu, Boqi Wu, Peiyu Geng, Chao Wang, Jia Yang, Zhifei Cheng, and et al. 2023. "Achieving Self-Supported Hierarchical Cu(OH)2/Nickel–Cobalt Sulfide Electrode for Electrochemical Energy Storage" Micromachines 14, no. 1: 125. https://doi.org/10.3390/mi14010125

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