Abstract

Improving the electrochemical performance of biomass-derived carbon electrode-active materials for supercapacitor applications has recently attracted considerable attention. Herein, we develop hybrid electrode materials from rice-husk-derived porous carbon (RH-C) materials and β-Ni(OH)2 via a facile solid-state reaction strategy comprising two steps. The prepared RH-C/Ni(OH)2 (C–Ni) was investigated using scanning electron microscopy (SEM) (energy-dispersive X-ray spectrometer (EDS)), X-ray photoelectron spectroscopy (XPS), and X-ray diffraction (XRD) to acquire the physical and chemical information, which was used to demonstrate the successful fabrication of C–Ni. Thermogravimetric analysis (TGA) measurement results confirmed that the thermal stability of C–Ni changed due to the presence of Ni(OH)2. As expected, C–Ni possesses a high capacitance of ∼952 F/g at a current density of 1.0 A/g. This result is higher than that of pure biomass-based carbon materials under the three-electrode system. This facile preparation method, which was used to synthesize the electrode-active materials, can extend to the value-added utility of other waste biomass materials as high-performing supercapacitor electrodes for energy storage applications.
1. Introduction
Recently, the overuse of nonrenewable fossil fuel resources (i.e., coal and oil) has resulted in a considerable environmental pollution and energy crisis, which hinders the sustainable development of modern society.1 Thus, it is necessary to develop new energy technologies from renewable materials, especially from abundant and low-cost biomass-based materials.
In the last few years, supercapacitors, considered an important storage technology, have received considerable interest owing to their high power density,2 high charge–discharge capability, and durable cyclic lifespan.3−5 The performance of supercapacitors typically depends on electrode materials; supercapacitors can be broadly classified into three groups, i.e., (1) carbon-based, (2) metal-oxide-based, and (3) conductive-polymer-based supercapacitors.6
Biomass-derived carbon materials have been favored by many scholars owing to their low cost, wide distribution, and green and renewable characteristics.7,8 Various biomasses (e.g., bio-straw,9 onion,10 soybean root,11 tannic acid,12 and pine nut shell13) have been used as precursors to prepare carbon materials. Agricultural products, like stems and husks of grain or cereal plants,14,15 plant polyphenol,16−18 plant cellulose,19 and so on, are a great source of biomasses. From the long-term point of view, sustainable biomass has a considerable potential for the production of new-generation electrode materials to replace or complement traditional carbon precursors.20,21 According to the energy storage principles, carbon-based supercapacitors are considered electrochemical double-layer capacitors (EDLCs). For EDLCs, the energy storage principle is based on the physical adsorption/desorption of the electrolyte ions at the interface of the electrode, which subsequently leads to the formation of a double layer. Although EDLCs with carbon-based electrodes have the advantages of very high power density and outstanding cycling lifetime, they are restricted in practical application owing to the inferior energy density caused by their low capacity.22 Thus, many scientists are devoted to studying how to efficiently deal with the above-mentioned drawbacks to meet the need of practical applications. For example, He et al.23 determined that the capacitance of ladderlike porous carbon can be improved from 165 to 214.5 F/g at a current density of 0.5 A/g using dual N and S doping. Yang et al.24 have prepared a biomass-derived interconnected hierarchical micro–meso–macroporous carbon with a higher electrochemical performance, which is attributed to the comprehensive effects from the high specific surface area, multilevel porous structure, high level of graphitization, and rich heteroatoms codoping. Up to now, many studies have shown that the energy density of carbon-based materials can be considerably improved by compounding with transition-metal oxides/hydroxides owing to their ultrahigh theoretical specific capacitance values.
Numerous transition-metal oxides/hydroxides, e.g., CuO,25 Ni(OH)2,26,27 MnO2,28,29 Co(OH)2,30 and Co3O4,31 are excellent electrode materials and have been used to construct high energy and power supercapacitor devices because these materials possess high theoretical capacitances. Among the above-mentioned pseudocapacitive materials, Ni(OH)2 is the most promising owing to the advantages of a distinct redox reaction, controllable morphology, and low production cost.32,33 Our group has reported that Ni(OH)2 could significantly increase the electrochemical performance of carbon nanofibers after they were synthesized on the surface of the matrix materials through a facile solid-state reaction to produce supercapacitor electrodes.34
In this study, we successfully prepared C–Ni composite materials through a simple two-step solid-state reaction using biomass-based carbon materials (Figure 1). The samples were characterized by several common techniques, i.e., SEM, XRD, and XPS, to acquire the chemical and physical information about them. TGA measurement was used to determine the thermal stability of composite materials. A three-electrode setup was used to study the electrochemical performance of C–Ni composite. Supercapacitor electrodes, which were obtained by compounding biomass-based carbon materials and Ni(OH)2, delivered a high specific capacitance of ∼952 F/g at a current density of 1.0 A/g. The developed facile electrode material preparation process presented in this work can extend to the value-added utility of other waste biomass materials, which can be used as electrode-active materials in high-performing supercapacitors and other energy storage equipment.
Figure 1.

Schematics of the preparation of the C–Ni composite materials.
2. Results and Discussion
2.1. Characterization of C–Ni Composite Active Materials
Figure 2 presents the microstructure of C–Ni and RH-C. After grinding, the obtained RH-C materials (Figure 2a) appeared as a microsized fragmentary block and possessed a smooth surface; the surface of C–Ni composite materials, as shown in Figure 2b, became relatively tough because of the accumulation of Ni(OH)2 particles on the surface, which suggested the successful preparation of the C–Ni composite. The elemental distribution on the surface of carbonized C–Ni composite materials was investigated using an EDS. According to Figure 2c–f, a uniform distribution of Ni and O could be observed throughout the carbon material surface, which is consistent with the SEM result.
Figure 2.

SEM images of (a) RH-C and (b) C–Ni. (c) SEM image of C–Ni. (d–f) The corresponding SEM–EDS images of (d) C, (e) Ni, and (f) O.
XPS was also used to confirm the presence of Ni on the carbon material surface, and the typical peaks in the full survey spectrum are shown in Figure 3a. For pure carbon materials, the peaks at ∼284 and ∼533 eV are ascribed to C 1s and O 1s, respectively. However, a typical wide scan of C–Ni composite materials showed C 1s (∼284 eV), Ni 2p (∼856 eV), and O 1s (∼531 eV) species, which suggested the presence of Ni in the carbon material substrate. The high-resolution spectra of the Ni 2p peak (Figure 3b) confirm the presence of two major binding energies at ∼874 and ∼856 eV, which are assigned to Ni 2p1/2 and Ni 2p3/2, respectively.35 XRD was used to determine the crystal structure of the samples and used for further qualitative analysis. The XRD patterns of RH-C and C–Ni material are shown in Figure 3c. The XRD profile of RH-C shows two main broad diffraction peaks at 21.6 and 43.8°, which correspond to the (002) and (100) crystalline planes of graphite, respectively.36,37 After the solid-state reaction, several clear diffraction peaks appear at 18.5, 33.2, 39.4, 51.4, 59.3, and 62.8°, which correspond to the (001), (100), (101), (102), (110), and (111) planes of the β-phase nickel hydroxide, respectively.38 This result confirms the formation of C–Ni composite. The thermal stability of composites was evaluated by TGA. The weight loss of the samples was recorded in the TG curves in detail (Figure 3d). For carbon materials, the weight remained nearly unchanged and was 96.74% at 800 °C. The weight of C–Ni composites decreased considerably. Weight loss can be divided mainly into three stages. (1) A small weight loss occurred in the first stage from 30 to 200 °C owing to the evaporation of adsorbed and bound water. (2) The second stage occurring from 200 to 600 °C corresponds to the thermal decomposition of Ni(OH)2 into NiO. (3) During the third stage, the composite had almost no significant weight loss. The results of TGA characterization showed that C–Ni had lower stability than that of pure carbon materials due to the introduction of Ni(OH)2.
Figure 3.
(a) XPS survey spectrum; (b) high-resolution Ni 2p XPS spectra of C–Ni; (c) XRD patterns; and (d) TG curves.
2.2. Electrochemical Performance of C–Ni Composite Materials
The electrochemical performance of C–Ni was examined using cyclic voltammetry (CV), galvanostatic charge–discharge (GCD), and electrochemical impedance spectroscopy (EIS) tests, and the results are presented in Figure 4. Figure 4a reveals that the area under the CV curve of carbon materials without Ni(OH)2 is relatively small, which indicates a lower capacitance, compared with that of C–Ni composite. Two peaks appeared in the CV profile of C–Ni at 10 mV/s scan rates (i.e., a cathodic peak at ∼0.20 V and an anodic peak at ∼0.49 V), which were ascribed to the Faradic reaction of Ni(OH)2. Specifically, the mechanism corresponds to the redox reaction in the reversible conversion between Ni2+ and Ni3+ as follows39,40
As seen in Figure 4b, the current density increased with an increase in the scan rate (between 1 and 20 mV/s), and all CV curves had analogous profiles, which indicated the excellent rate capability of the electrode.32 However, anodic and cathodic peaks were clearly shifted toward positive and negative potentials, respectively, with increasing scan rate due to the controlled internal diffusion and limited charge transfer in the electrode surface.41,42 GCD curves were recorded to further characterize the electrochemical behavior of the as-synthesized C–Ni composite at different current densities (Figure 4c). The nonlinear charge–discharge plateaus of the GCD curves further demonstrated that the typical Faradic redox capacitance of C–Ni materials was mainly due to the reversible conversion between Ni(OH)2 and NiOOH, which was consistent with the CV profiles. Moreover, nearly symmetric nonlinearities in GCD curves confirmed electrochemical reversibility and the rapid I–V response of these materials.43 The specific capacity values under various current densities are presented in Figure 4d. Cs gradually decreased with an increase in the current density owing to ion diffusion resistance. Under faster-charging conditions, electrolytic ions did not have sufficient time to enter the interior of C–Ni composite materials through the pores, which caused the internal active materials to become inaccessible for the redox reaction and provided the corresponding capacity. At a low current density, the electrolytic ions filled the internal structure of these active materials after a sufficient amount of time. C–Ni delivered high gravimetric capacitance of ∼952 F/g at 1 A/g. Even at a high current density of 30 A/g, ∼126 F/g can be achieved for the sample. It should be pointed out that this capacitance value is comparable to those of some of the biomass-derived carbon composite electrodes (see Table 1).
Figure 4.
(a) Typical CV profiles of RH-C and C–Ni at 10 mV/s. (b) Various CV profiles for C–Ni at various scan rates. (c) Various GCD profiles for C–Ni at various current densities. (d) Calculated gravimetric capacitance at different current densities for C–Ni. (e) Nyquist profiles of C–Ni (inset showing the magnified region of 0–2 Ω). (f) Cycle stability performance of C–Ni (inset showing the initial 10 circle GCD curves at the current density of 20 A/g).
Table 1. Specific Capacitance (Cs) of Biomass-Derived Carbon Composite Electrode Materials Reported the Literature.
| biomass-derived carbon composite materials | electrolyte | Cs (F/g) | refs |
|---|---|---|---|
| fungal conidium-derived carbon/MnO2 | 1.0 M Na2SO4 | 263.5 (1 A/g) | (44) |
| pine cone flowers-derived carbon/Ni(OH)2 | 1.0 M KOH | 916.4 (1 A/g) | (45) |
| willow catkin-derived carbon MnO2 | 1.0 M Na2SO4 | 189 (1 A/g) | (46) |
| Faidherbia albida fruit shell-derived carbon MnO2 | 3.0 M KOH | 426.66 (1 A/g) | (47) |
| celery-derived carbon/polyaniline | 1.0 M H2SO4 | 402 (1 A/g) | (48) |
| banana peel-derived carbon MnO2 | 1.0 M Na2SO4 | 139.6 (0.3 A/g) | (49) |
| bamboo leaves-derived carbon/CuOx | 3.0 M KOH | 147 (1 A/g) | (50) |
| husk-derived carbon/Ni(OH)2 | 6.0 M KOH | 952 (1 A/g) | this work |
EIS measurements were used to gain more detailed information about the fundamental behavior of electrode materials, including inherent resistance, charge transfer, and ionic resistance. The corresponding Nyquist plot of the C–Ni composite electrode is shown in Figure 4e. In the low-frequency region, the line with a linear behavior has a larger angle (∼65°) on the X-axis, which indicates good capacitive behavior with a short electron transfer and ion diffusion path in the C–Ni electrode materials. The charge-transfer resistance and solution resistance corresponded to the semicircle radius and the first intercept value of the Z’real-axis, respectively, in the high-frequency region. Both charge-transfer and solution resistances were relatively low (<1.0 Ω), which indicated a low equivalent series resistance (ESR) of the as-prepared electrode materials. This composite electrode material also exhibits good capacitance retention of 81.3% and an acceptable Coulombic efficiency after 1000 cycles even under a high charge/discharged current density of 20 A/g, which is comparable with other similar reports40,51 (Figure 4f).
3. Conclusions
In summary, we successfully fabricated biomass-based carbon and β-Ni(OH)2 hybrid electrode materials via a facile solid-state reaction method involving two steps. The physical and chemical properties of C–Ni composite materials were studied with the use of several characterization techniques, including SEM, XPS, and XRD. The results of TGA showed that the thermal stability of composite materials changed owing to the incorporation of Ni(OH)2. The gravimetric capacitance of C–Ni electrode materials reached ∼952 F/g at 1.0 A/g based on a three-electrode system configuration. The developed material can be applied for the development of high-performing supercapacitor electrodes for energy storage applications. This facile preparation approach of electrode materials can extend to the value-added utility of other waste biomass materials.
4. Materials and Methods
4.1. Materials
Rice husk was purchased from taobao.com (Alibaba, China). A hydrochloric acid solution with >31.6% HCl and potassium hydroxide (KOH) was purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Nickel(II) acetate tetrahydrate [Ni(CH3COO)2·4H2O] and oxalic acid (C2H2O4·2H2O) were provided by Shanghai Shanpu Chemical Co., Ltd. (Shanghai, China) and Tianjing Kaitong Chemical Reagent Co., Ltd. (Tianjing, China), respectively. All reagents were of analytical grade, and they were used as received without further purification; all aqueous solutions used were prepared with deionized water.
4.2. Preparation of RH-C
Risk husk was subjected to carbonization for 2 h at 1000 °C in an N2 atmosphere at the heating and cooling rates of 4 °C/min. The entire process of carbonization was conducted in a closed corundum tube. The as-prepared rice-husk-based carbon was ground into powder and then purified by immersing in a 10% (v/v) HCl solution for 3 h, followed by soaking in a NaOH solution (2 M) at 80 °C for 2 h to remove inorganic salts and silica. Finally, porous carbon materials derived from biomass were obtained for further use in supercapacitor electrodes in the next step.
4.3. Fabrication of C–Ni
C–Ni was synthesized via a facile solid-state reaction, which involved two reaction steps, according to our group’s previous study with minor modification.34 First, a mixture of oxalic acid and nickel acetate (1:1, mole ratio) was ground for 0.5 h and then rinsed numerous times with distilled water and ethanol. Then, the dried precursor and RH-C were loaded in an agate mortar and ground for 0.5 h. Further, 0.123 g of KOH was then added to the compounds and ground for another 1 h. The entire grinding process was performed under room temperature. Finally, the sample was washed with distilled water and eventually dried at 80 °C to achieve the C–Ni composite. The schematic of the preparation process for C–Ni is presented in Figure 1.
4.4. Characterization
The surface morphology of the samples was investigated using SEM (S-3000N; Hitachi, Japan). XPS (Axis Ultra DLD) was performed to determine the surface elemental composition. XRD measurement was used to obtain the crystallographic information and performed on a PANalytical Empyrean instrument (Netherlands) with a Cu Kα radiation in the 2θ range of 10–70°. Thermal stability was analyzed using a METTLER TOLEDO TGA/DSC/110SF instrument in a N2 atmosphere from 30 to 800 °C with a ramping rate of 10 °C/min.
4.5. Electrochemical Testing
A three-electrode setup with 6 M KOH electrolyte was used for all electrochemical measurements (which consisted of a platinum foil as the counter electrode, Hg/HgO electrode as the reference electrode, and a working electrode), and the electrochemical performance was evaluated using an electrochemical workstation (CHI660E; Shanghai Chenhua, China) at room temperature. The preparation of the working electrode was according to a previously reported procedure.52 Briefly, the electrode was prepared by pressing a slurry into a 1 cm × 1 cm nickel foam; the slurry contained active materials, poly(vinylidene fluoride), and carbon black with the mass ratio of 80:5:15 in a small amount of N-methylpyrrolidone solvent; then, the mixture was dried at 80 °C. The prepared C–Ni composite coated onto a Ni-foam substrate served as a working electrode for the electrochemical measurements. Galvanostatic discharge (GCD) curves were used to determine the specific capacitance (Cs) of the electrode (F/g) based on the following equation53
| 1 |
where ΔV is the potential window (V), m is the mass of the as-prepared active material (g), I is the discharge current (A), and Δt is the discharge time (s).
Acknowledgments
Dr. J.C. thanks the Young Elite Scientists Sponsorship Program by CAST (No. 2018QNRC001) and the Chutian Scholar Program of Hubei Provincial Government, China.
Author Contributions
J.C. conceived the overall project and finalized the manuscript. D.Z. performed the experiments and data analyses. J.C. and D.Z. wrote of the manuscript. W.-P.D. and P.F. contributed significantly to the analysis with constructive discussions. Z.-Z.Z., G.-Z.W., J.-R.H., H.-B.W., and T.-L.S. supported the experiment.
This work was funded by the Province Key Laboratory of Cereal Resource Transformation and Utilization, Henan University of Technology (No. PL2018001).
No potential conflict of interest was reported by the authors.
The authors declare no competing financial interest.
References
- Bi Z.; Kong Q.; Cao Y.; Sun G.; Su F.; Wei X.; Li X.; Ahmad A.; Xie L.; Chen C. Biomass-derived porous carbon materials with different dimensions for supercapacitor electrodes: A review. J. Mater. Chem. A 2019, 7, 16028–16045. 10.1039/C9TA04436A. [DOI] [Google Scholar]
- Zhang X.; Pei Z.; Wang C.; Yuan Z.; Wei L.; Pan Y.; Mahmood A.; Shao Q.; Chen Y. Flexible zinc-ion hybrid fiber capacitors with ultrahigh energy density and long cycling life for wearable electronics. Small 2019, 15, 1903817 10.1002/smll.201903817. [DOI] [PubMed] [Google Scholar]
- Kazemi S. H.; Malae K. Electrodeposited Ni(OH)2 nanostructures on electro-etched carbon fiber paper for highly stable supercapacitors. J. Iran. Chem. Soc. 2017, 14, 419–425. 10.1007/s13738-016-0990-z. [DOI] [Google Scholar]
- Li Y.; Zhang D.; Zhang Y.; He J.; Wang Y.; Wang K.; Xu Y.; Li H.; Wang Y. Biomass-derived microporous carbon with large micropore size for high-performance supercapacitors. J. Power Sources 2020, 448, 227396 10.1016/j.jpowsour.2019.227396. [DOI] [Google Scholar]
- Qu Z.; Shi M.; Wu H.; Liu Y.; Jiang J.; Yan C. An efficient binder-free electrode with multiple carbonized channels wrapped by NiCo2O4 nanosheets for high-performance capacitive energy storage. J. Power Sources 2019, 410–411, 179–187. 10.1016/j.jpowsour.2018.11.018. [DOI] [Google Scholar]
- Jiang X.; Cheng W.; Hu H.; Hu Y.; Cao Y.; Yan S.; Di R.; Wang X.; Hou L. Facile preparation of a novel composite Co-Ni(OH)2/carbon sphere for high-performance supercapacitors. Mater. Technol. 2019, 34, 204–212. 10.1080/10667857.2018.1548115. [DOI] [Google Scholar]
- Liu Y.; Shi M.; Yan C.; Zhuo Q.; Wu H.; Wang L.; Liu H.; Guo Z. Inspired cheese-like biomass-derived carbon with plentiful heteroatoms for high performance energy storage. J. Mater. Sci.: Mater. Electron. 2019, 30, 6583–6592. 10.1007/s10854-019-00965-2. [DOI] [Google Scholar]
- Liu Y.; Huang B.; Lin X.; Xie Z. Biomass-derived hierarchical porous carbons: boosting the energy density of supercapacitors via an ionothermal approach. J. Mater. Chem. A 2017, 5, 13009–13018. 10.1039/C7TA03639F. [DOI] [Google Scholar]
- Guan M.; Zhang X.; Wu Y.; Sun Q.; Dong D.; Zhang X.; Wang J. Biomass straw based activated porous carbon materials for high-performance supercapacitors. Res. Appl. Mater. Sci. 2020, 1, 27–30. 10.33142/msra.v1i2.1665. [DOI] [Google Scholar]
- Zhang W.; Xu J.; Hou D.; Yin J.; Liu D.; He Y.; Lin H. Hierarchical porous carbon prepared from biomass through a facile method for supercapacitor applications. J. Colloid Interface Sci. 2018, 530, 338–344. 10.1016/j.jcis.2018.06.076. [DOI] [PubMed] [Google Scholar]
- Guo N.; Luo W.; Guo R.; Qiu D.; Zhao Z.; Wang L.; Jia D.; Guo J. Interconnected and hierarchical porous carbon derived from soybean root for ultrahigh rate supercapacitors. J. Alloys Compd. 2020, 834, 155115 10.1016/j.jallcom.2020.155115. [DOI] [Google Scholar]
- Wang G.; Qin J.; Zhao Y.; Wei J. Nanoporous carbon spheres derived from metal-phenolic coordination polymers for supercapacitor and biosensor. J. Colloid Interface Sci. 2019, 544, 241–248. 10.1016/j.jcis.2019.03.001. [DOI] [PubMed] [Google Scholar]
- Qin L.; Hou Z.; Zhang S.; Zhang W.; Jiang E. Supercapacitive charge storage properties of porous carbons derived from pine nut shells. J. Electroanal. Chem. 2020, 866, 114140 10.1016/j.jelechem.2020.114140. [DOI] [Google Scholar]
- He D.; Zhao W.; Li P.; Liu Z.; Wu H.; Liu L.; Han K.; Liu L.; Wan Q.; Butt F. K.; Qu X. Bifunctional biomass-derived 3D nitrogen-doped porous carbon for oxygen reduction reaction and solid-state supercapacitor. Appl. Surf. Sci. 2019, 465, 303–312. 10.1016/j.apsusc.2018.09.185. [DOI] [Google Scholar]
- Weng W.; Wang S.; Xiao W.; Lou X. W. D. Direct Conversion of Rice Husks to Nanostructured SiC/C for CO2 Photoreduction. Adv. Mater. 2020, 32, 2001560 10.1002/adma.202001560. [DOI] [PubMed] [Google Scholar]
- Cai J.; Zeng F.; Zheng S.; Huang X.; Zhang J.; Zhang P.; Fei P. Preparation of lipid-soluble bilberry anthocyanins through acylation with cinnamic acids and its antioxidation activities. J. Agric. Food Chem. 2020, 68, 7467–7473. 10.1021/acs.jafc.0c01912. [DOI] [PubMed] [Google Scholar]
- Gurung M.; Adhikari B. B.; Alam S.; Kawakita H.; Ohto K.; Inoue K.; Harada H. Adsorptive removal of Cs(I) from aqueous solution using polyphenols enriched biomass-based adsorbents. Chem. Eng. J. 2013, 231, 113–120. 10.1016/j.cej.2013.06.028. [DOI] [Google Scholar]
- Fei P.; Zeng F.; Zheng S.; Chen Q.; Hu Y.; Cai J. Acylation of blueberry anthocyanins with maleic acid: Improvement of the stability and its application potential in intelligent color indicator packing materials. Dyes Pigm. 2021, 184, 108852 10.1016/j.dyepig.2020.108852. [DOI] [Google Scholar]
- Cai J.; Zhang D.; Xu W.; Ding W. P.; Zhu Z. Z.; He J. R.; Cheng S. Y. Polysaccharide-Based Hydrogels Derived from Cellulose: The Architecture Change from Nanofibers to Hydrogels for a Putative Dual Function in Dye Wastewater Treatment. J. Agric. Food Chem. 2020, 68, 9725–9732. 10.1021/acs.jafc.0c03054. [DOI] [PubMed] [Google Scholar]
- Liu X.; Ma C.; Li J.; Zielinska B.; Kalenczuk R. J.; Chen X.; Chu P. K.; Tang T.; Mijowska E. Biomass-derived robust three-dimensional porous carbon for high volumetric performance supercapacitors. J. Power Sources 2019, 412, 1–9. 10.1016/j.jpowsour.2018.11.032. [DOI] [Google Scholar]
- Dutta S.; Bhaumik A.; Wu K. C. W. Hierarchically porous carbon derived from polymers and biomass: effect of interconnected pores on energy applications. Energy Environ. Sci. 2014, 7, 3574–3592. 10.1039/C4EE01075B. [DOI] [Google Scholar]
- Xie M.; Xu Z.; Duan S.; Tian Z.; Zhang Y.; Xiang K.; Lin M.; Guo X.; Ding W. Facile growth of homogeneous Ni(OH)2 coating on carbon nanosheets for high-performance asymmetric supercapacitor applications. Nano Res. 2018, 11, 216–224. 10.1007/s12274-017-1621-4. [DOI] [Google Scholar]
- He D.; Zhao W.; Li P.; Sun S.; Tan Q.; Han K.; Liu L.; Liu L.; Qu X. Bifunctional biomass-derived N, S dual-doped ladder-like porous carbon for supercapacitor and oxygen reduction reaction. J. Alloys Compd. 2019, 773, 11–20. 10.1016/j.jallcom.2018.09.141. [DOI] [Google Scholar]
- Yang S.; Wang S.; Liu X.; Li L. Biomass derived interconnected hierarchical micro-meso-macro- porous carbon with ultrahigh capacitance for supercapacitors. Carbon 2019, 147, 540–549. 10.1016/j.carbon.2019.03.023. [DOI] [Google Scholar]
- Zhang Y.; Huang M.; Li F.; Wen Z. Controlled synthesis of hierarchical CuO nanostructures for electrochemical capacitor electrodes. Int. J. Electrochem. Sci. 2013, 8, 8645–8661. [Google Scholar]
- Qiu H.; An S.; Sun X.; Yang H.; Zhang Y.; He W. Excellent performance MWCNTs-GONRs/Ni(OH)2 electrode for outstanding supercapacitors. Ceram. Int. 2019, 45, 18422–18429. 10.1016/j.ceramint.2019.06.059. [DOI] [Google Scholar]
- Li W.; Shih Y.; Cheng H. Green synthesis of CNTs/Ni(OH)2 nanostructures for electrochemical supercapacitors. Chem. Phys. Lett. 2020, 750, 137499 10.1016/j.cplett.2020.137499. [DOI] [Google Scholar]
- Shah H. U.; Wang F.; Javed M. S.; Ahmad M. A.; Saleem M.; Zhan J.; Khan Z. U. H.; Li Y. In-situ growth of MnO2 nanorods forest on carbon textile as efficient electrode material for supercapacitors. J. Energy Storage 2018, 17, 318–326. 10.1016/j.est.2018.03.015. [DOI] [Google Scholar]
- Pei Z.; Zhu M.; Huang Y.; Huang Y.; Xue Q.; Geng H.; Zhi C. Dramatically improved energy conversion and storage efficiencies by simultaneously enhancing charge transfer and creating active sites in MnOx/TiO2 nanotube composite electrodes. Nano Energy 2016, 20, 254–263. 10.1016/j.nanoen.2015.12.025. [DOI] [Google Scholar]
- Zhao C.; Liang S.; Jiang Y.; Gao F.; Xie L.; Chen L. Mn doped Co(OH)2 nanosheets as electrode materials for high performance supercapacitors. Mater. Lett. 2020, 270, 127751 10.1016/j.matlet.2020.127751. [DOI] [Google Scholar]
- Atchudan R.; Edison T. N. J. I.; Chakradhar D.; Karthik N.; Perumal S.; Lee Y. R. One-pot dual product synthesis of hierarchical Co3O4@N-rGO for supercapacitors, N-GDs for label-free detection of metal ion and bio-imaging applications. Ceram. Int. 2018, 44, 2869–2883. 10.1016/j.ceramint.2017.11.034. [DOI] [Google Scholar]
- Anwer S.; Bin Ari A.; Bharath G.; Cao P.; Patole S. P.; Luo S.; Masood H. T.; Cantwell W. J.; Liao K.; Li Q.; Zheng L. Synthesis of joint-welded carbon nanotube foam@Ni(OH)2 nanosheet-based core–shell 3D architecture for freestanding flexible electrode for supercapacitor applications. Adv. Mater. Interfaces 2019, 6, 1900670 10.1002/admi.201900670. [DOI] [Google Scholar]
- Wen S.; Qin K.; Liu P.; Zhao N.; Shi C.; Ma L.; Liu E. Ultrafine Ni(OH)2 nanoneedles on N-doped 3D rivet graphene film for high-performance asymmetric supercapacitor. J. Alloys Compd. 2019, 783, 625–632. 10.1016/j.jallcom.2018.12.347. [DOI] [Google Scholar]
- Cai J.; Niu H.; Li Z.; Du Y.; Cizek P.; Xie Z.; Xiong H.; Lin T. High-performance supercapacitor electrode materials from cellulose-derived carbon nanofibers. ACS Appl. Mater. Interfaces 2015, 7, 14946–14953. 10.1021/acsami.5b03757. [DOI] [PubMed] [Google Scholar]
- Chen L.; Liu L.; Guo Q.; Wang Z.; Liu G.; Chen S.; Hou H. Preparation of Ni(OH)2 nanoplatelet/electrospun carbon nanofiber hybrids for highly sensitive nonenzymatic glucose sensors. RSC Adv. 2017, 7, 19345–19352. 10.1039/C7RA02064C. [DOI] [Google Scholar]
- Cao W.; Zhang E.; Wang J.; Liu Z.; Ge J.; Yu X.; Yang H.; Lu B. Potato derived biomass porous carbon as anode for potassium ion batteries. Electrochim. Acta 2019, 293, 364–370. 10.1016/j.electacta.2018.10.036. [DOI] [Google Scholar]
- Tang Z.; Pei Z.; Wang Z.; Li H.; Zeng J.; Ruan Z.; Huang Y.; Zhu M.; Xue Q.; Yu J.; Zhi C. Highly anisotropic, multichannel wood carbon with optimized heteroatom doping for supercapacitor and oxygen reduction reaction. Carbon 2018, 130, 532–543. 10.1016/j.carbon.2018.01.055. [DOI] [Google Scholar]
- Tang Y.; Liu Y.; Guo W.; Chen T.; Wang H.; Yu S.; Gao F. Highly oxidized graphene anchored Ni(OH)2 nanoflakes as pseudocapacitor materials for ultrahigh loading electrode with high areal specific capacitance. J. Phys. Chem. C 2014, 118, 24866–24876. 10.1021/jp5075779. [DOI] [Google Scholar]
- He W.; Qiu H.; Meng J.; Liu B.; Cui J.; Zhang Y. Microwave-assisted synthesis of nano petal-like α-Ni (OH)2/RGO and its electrochemical performance. J. Alloys Compd. 2019, 788, 183–190. 10.1016/j.jallcom.2019.02.111. [DOI] [Google Scholar]
- Liu J.; Wang Y.; Hu R.; Munir H. A.; Liu H. High-performance supercapacitor electrode based on 3D rose-like β-Ni(OH)2/rGO nanohybrid. J. Phys. Chem. Solids 2020, 138, 109297 10.1016/j.jpcs.2019.109297. [DOI] [Google Scholar]
- Upadhyay K. K.; Bundaleska N.; Abrashev M.; Bundaleski N.; Teodoro O. M. N. D.; Fonseca I.; de Ferro A. M.; Silva R. P.; Tatarova E.; Montemor M. F. Free-standing N-Graphene as conductive matrix for Ni(OH)2 based supercapacitive electrodes. Electrochim. Acta 2020, 334, 135592 10.1016/j.electacta.2019.135592. [DOI] [Google Scholar]
- Ji J.; Zhang L.; Ji H.; Li Y.; Zhao X.; Bai X.; Fan X.; Zhang F.; Ruoff R. S. Nanoporous Ni(OH)2 thin film on 3D ultrathin-graphite foam for asymmetric supercapacitor. ACS Nano 2013, 7, 6237–6243. 10.1021/nn4021955. [DOI] [PubMed] [Google Scholar]
- Cai J.; Xu W.; Liu Y.; Zhu Z.; Liu G.; Ding W.; Wang G.; Wang H.; Luo Y. Robust construction of flexible bacterial cellulose@Ni(OH)2 paper: toward high capacitance and sensitive H2O2 detection. Eng. Sci. 2018, 5, 21–29. 10.30919/es8d669. [DOI] [Google Scholar]
- Mao C.; Liu S.; Pang L.; Sun Q.; Liu Y.; Xu M.; Lu Z. Ultrathin MnO2 nanosheets grown on fungal conidium-derived hollow carbon spheres as supercapacitor electrodes. RSC Adv. 2016, 6, 5184–5191. 10.1039/C5RA22193E. [DOI] [Google Scholar]
- Nagaraju G.; Cha S. M.; Yu J. S. Ultrathin nickel hydroxide nanosheet arrays grafted biomass-derived honeycomb-like porous carbon with improved electrochemical performance as a supercapacitive material. Sci. Rep. 2017, 7, 45201 10.1038/srep45201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y.; Yu N.; Yan P.; Li Y.; Zhou X.; Chen S.; Wang G.; Wei T.; Fan Z. Fabrication of manganese dioxide nanoplates anchoring on biomass-derived cross-linked carbon nanosheets for high-performance asymmetric supercapacitors. J. Power Sources 2015, 300, 309–317. 10.1016/j.jpowsour.2015.09.077. [DOI] [Google Scholar]
- Mohammed A. A.; Chen C.; Zhu Z. Green and high performance all-solid-state supercapacitors based on MnO2/Faidherbia albida fruit shell derived carbon sphere electrodes. J. Power Sources 2019, 417, 1–13. 10.1016/j.jpowsour.2019.02.003. [DOI] [Google Scholar]
- Du W.; Wang X.; Sun X.; Zhan J.; Zhang H.; Zhao X. Nitrogen-doped hierarchical porous carbon using biomass-derived activated carbon/carbonized polyaniline composites for supercapacitor electrodes. J. Electroanal. Chem. 2018, 827, 213–220. 10.1016/j.jelechem.2018.09.031. [DOI] [Google Scholar]
- Yang G.; Park S.-J. MnO2 and biomass-derived 3D porous carbon composites electrodes for high performance supercapacitor applications. J. Alloys Compd. 2018, 741, 360–367. 10.1016/j.jallcom.2018.01.108. [DOI] [Google Scholar]
- Wang Q.; Zhang Y.; Xiao J.; Jiang H.; Hu T.; Meng C. Copper oxide/cuprous oxide/hierarchical porous biomass-derived carbon hybrid composites for high-performance supercapacitor electrode. J. Alloys Compd. 2019, 782, 1103–1113. 10.1016/j.jallcom.2018.12.235. [DOI] [Google Scholar]
- Lai L.; Li R.; Su S.; Zhang L.; Cui Y.; Guo N.; Shi W.; Zhu X. Controllable synthesis of reduced graphene oxide/nickel hydroxide composites with different morphologies for high performance supercapacitors. J. Alloys Compd. 2020, 820, 153120 10.1016/j.jallcom.2019.153120. [DOI] [Google Scholar]
- Cai J.; Niu H.; Wang H.; Shao H.; Fang J.; He J.; Lin T.; et al. High-performance supercapacitor electrode from cellulose-derived, inter-bonded carbon nanofibers. J. Power Sources 2016, 324, 302–308. 10.1016/j.jpowsour.2016.05.070. [DOI] [Google Scholar]
- Qiu H.; An S.; Sun X.; Yang H.; Zhang Y.; He W. MWCNTs-GONRs/Co3O4@Ni(OH)2 core-shell array structure with a high performance electrode for supercapacitor. Chem. Eng. J. 2020, 380, 122490 10.1016/j.cej.2019.122490. [DOI] [Google Scholar]


