Next Article in Journal
Influence of Cathode Catalyst Layer with SiO2-Coated Pt/Ketjen Black Catalysts on Performance for Polymer Electrolyte Fuel Cells
Previous Article in Journal
Enhanced Catalytic Hydrogen Peroxide Production from Hydroxylamine Oxidation on Modified Activated Carbon Fibers: The Role of Surface Chemistry
Previous Article in Special Issue
Highly Efficient MOF Catalyst Systems for CO2 Conversion to Bis-Cyclic Carbonates as Building Blocks for NIPHUs (Non-Isocyanate Polyhydroxyurethanes) Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Hydrothermal Synthesis and Supercapacitor Performance of Mesoporous Necklace-Type ZnCo2O4 Nanowires

by
John Anthuvan Rajesh
and
Kwang-Soon Ahn
*
School of Chemical Engineering, Yeungnam University, Gyeongsan 712-749, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(12), 1516; https://doi.org/10.3390/catal11121516
Submission received: 15 November 2021 / Revised: 9 December 2021 / Accepted: 10 December 2021 / Published: 13 December 2021
(This article belongs to the Special Issue Applications of Nanoporous Materials in Catalysis)

Abstract

:
In this work, mesoporous ZnCo2O4 electrode material with necklace-type nanowires was synthesized by a simple hydrothermal method using water/ethylene glycol mixed solvent and subsequent calcination treatment. The ZnCo2O4 nanowires were assembled by several tiny building blocks of nanoparticles which led to the growth of necklace-type nanowires. The as-synthesized ZnCo2O4 nanowires had porous structures with a high surface area of 25.33 m2 g−1 and with an average mesopore of 23.13 nm. Due to the higher surface area and mesopores, the as-prepared necklace-type ZnCo2O4 nanowires delivered a high specific capacity of 439.6 C g−1 (1099 F g−1) at a current density of 1 A g−1, decent rate performance (47.31% retention at 20 A g−1), and good cyclic stability (84.82 % capacity retention after 5000 cycles). Moreover, a hybrid supercapacitor was fabricated with ZnCo2O4 nanowires as a positive electrode and activated carbon (AC) as a negative electrode (ZnCo2O4 nanowires//AC), which delivered an energy density of 41.87 Wh kg−1 at a power density of 800 W kg−1. The high electrochemical performance and excellent stability of the necklace-type ZnCo2O4 nanowires relate to their unique architecture, high surface area, mesoporous nature, and the synergistic effect between Zn and Co metals.

1. Introduction

An electrochemical supercapacitor is one of the most effective techniques to store electrical energy, owing to its high power density, high charge/discharge rate, long operation time, higher rate capability, mechanical stability, and environmental friendliness [1,2]. In general, the performance of electrochemical supercapacitors mainly depends on the electrode material types and properties [3,4]. It is well known that supercapacitor electrode materials are classified into three main categories: electrical double layer electrode materials, pseudocapacitor electrode materials, and battery-type electrode materials [5,6]. In the past decade, tremendous efforts have been undertaken to improve the electrochemical performance of supercapacitor electrode materials, including various transition metals-based oxides, hydroxides, sulfides, selenides, nitrides, carbides, phosphides, carbon-based materials with high surface areas, and polymeric materials [7,8,9,10,11,12,13].
Among the various supercapacitor materials, transition metal oxides, particularly AB2O4-type cubic spinel binary transition metal oxides (BTMOs), have been extensively utilized in the growth of new functional materials, due to the facile synthesis methods, controllable morphologies, high surface to volume ratios, varied oxidation states, lower cost, and better electronic conductivity [14,15]. Generally, cubic spinel BTMOs have high surface to volume ratios, which is ultimately applied in the fields of lithium-ion batteries, supercapacitors, electrocatalysts, solar water splitting, and degradation of organic pollutants [14,15,16,17,18,19,20]. Among these, cubic spinel ZnCo2O4 is a promising energy storage material, owing to its various valence states, excellent electrochemical redox properties, better electrical conductivity, low cost, non-toxicity, and different morphologies [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. Currently, a variety of preparation methods are used to synthesize ZnCo2O4 nanomaterials, including hydrothermal [22,24,27,28,30,31,35,36], solvothermal [21,23,25,29], combustion [26], reflexing [32], and electrospinning methods [34]. Similarly, a series of ZnCo2O4 morphologies have been successfully synthesized, such as microspheres [21,30,32,33], nanorods [22,28], nanotubes [34], nanowires [24,31], nanosheets [22,27,29], nanobelt [24], nanoparticles [26], urchin-like [23], quasi cubes [25], and bipyramid nanoframes [35], and utilized widely in energy storage applications, mostly in electrochemical supercapacitors [21,22,23,24,25,26,27,28,30,31,34]. For instance, Bhagwan et al., prepared ZnCo2O4 nanoparticles via a combustion method for supercapacitors, showing a high specific capacitance of 843 F g−1 at a current density of 1 A g−1 [26]. Wang et al. reported the synthesis of porous ZnCo2O4 microspheres by a hydrothermal method, exhibiting a high specific capacitance of 953.2 F g−1 at 4 A g−1 [21]. Cheng et al. synthesized mesoporous ZnCo2O4 nanoflakes hybrids on nickel foam to enhance the specific capacitance (1220 F g−1 at 2 A g−1) [37].
In this work, we illustrate an easy way to synthesize another kind of cubic spinel ZnCo2O4 nanomaterial for effective supercapacitor application. Mesoporous necklace-type ZnCo2O4 nanowires were successfully synthesized by a facile water/ethylene glycol (EG)-mediated hydrothermal method. Various physicochemical characterization techniques revealed that the as-prepared ZnCo2O4 nanowires were built by several tiny building blocks of nanoparticles which led to the growth of necklace-type nanowires. Due to the high surface area, mesoporous structure, unique architecture, and synergistic effect, as-prepared necklace-type ZnCo2O4 nanowires delivered a high specific capacity (439.6 C g−1 at a current density of 1 A g−1), excellent rate performance (208.0 C g−1 at 20 A g−1), and outstanding cyclic stability (84.82 % capacity retention after 5000 cycles). Furthermore, a hybrid supercapacitor (HSC) device was fabricated with ZnCo2O4 nanowires as a positive electrode and activated carbon (AC) as a negative electrode (ZnCo2O4 nanowires//AC), which delivered an energy density of 41.87 Wh kg−1 and a maximum power density of 4000 W kg−1.

2. Results and Discussion

The crystalline phase of the as-prepared ZnCo2O4 sample is examined by XRD measurement. The XRD pattern in Figure 1 reveals that all peaks correspond to the characteristic peaks of spinel phase. All of the identified XRD peaks at around 18.95°, 31.08°, 36.70°, 38.54°, 44.71°, 55.42°, 59.18°, 65.01°, 73.41°, and 76.85° 2θ correspond to the (111), (220), (311), (222), (400), (422), (511), (440), (620), and (533) planes of spinel-type ZnCo2O4 (JCPDS No: 23–1390), respectively [21,22,23,24,25]. Moreover, no other crystalline phases were identified, which indicates the pure phase formation of ZnCo2O4 sample.
The chemical valence states of the ZnCo2O4 sample were verified by XPS (Figure 2a–d). Figure 2a shows the survey scan spectrum, which confirms the presence of Zn, Co, and O elements. The Zn 2p spectrum (Figure 2b) was deconvoluted into two spin–orbit doublets at 1022.1 and 1045.2 eV, which belong to the Zn 2p3/2 and Zn 2p1/2 peaks, respectively [23,25,26,27,29,30,31,38]. Two main peaks at binding energies of 781.3 and 796.3 eV were detected in the Co 2p spectrum (Figure 2c), corresponding to the Co 2p3/2 and Co 2p1/2 labels, respectively [23,25,26,27,29,30,31]. Further, peak deconvolution and fitting were employed for Co 2p3/2 and Co 2p1/2 labels. Two deconvoluted peaks at 781.2 and 796.3 eV were ascribed to Co3+, while another set of deconvoluted peaks situated at 782.9 and 797.9 eV were assigned to Co2+ [23,25,27]. Moreover, two shakeup satellite peaks (sat.) were also observed at 790.1 and 805.4 eV. The Co 2p XPS data indicate that both Co(II) and Co(III) oxidation states were present in the as-prepared ZnCo2O4 nanomaterial [29,30,31]. In the O 1s spectrum (Figure 2d), the peaks located at 530.4 and 531.8 eV were assigned to a metal–oxygen bond and typical chemisorbed oxygen, respectively [23,25,27].
The morphology of the synthesized ZnCo2O4 sample was observed by FE–SEM and FE–TEM measurements. Figure 3 shows the FE–SEM images of the necklace-type ZnCo2O4 nanowires at different magnification. A low-magnification SEM image in Figure 3a shows the flower-like structures in the synthesized material. A closer view of the flower-like structure (Figure 3b,c) reveals that the ZnCo2O4 sample is built by several nanowires. The high-magnification SEM image (Figure 3d) clearly indicates that each ZnCo2O4 nanowire was composed of numerous nanoparticles and that these nanoparticles were interconnected to each other to grow necklace-type nanowires. The EDS spectrum in Figure 3e confirms the presence of Zn, Co, and O elements within the necklace-type ZnCo2O4 nanowires.
The detailed microstructure of the necklace-type ZnCo2O4 nanowires was further characterized by TEM. Figure 4a shows a low-magnification TEM image of a typical flower-like ZnCo2O4 structure. Obviously, this flower-like structure is assembled by numerous nanowires, and the higher magnification TEM image (Figure 4b) clearly indicates that the flower-like structure was composed of many ZnCo2O4 nanowires. Figure 4c,d show the enlarged sections of the regions marked in yellow and red in Figure 4b, illustrating that the necklace-like structures were built by many primary nanoparticles which were interconnected to form a nanowire structure. Such a morphology is highly beneficial for the penetration of electrolyte ions during the charge/discharge process [31,34,39]. The inset image in Figure 4c is the selected area electron diffraction (SAED) pattern taken from the marked region in Figure 4c, which indicates the polycrystalline nature of the prepared necklace-type ZnCo2O4 nanowires [30,31,37]. Lattice fringes of a ZnCo2O4 nanowire were also illustrated in Figure 4d, and the distance between the two adjacent lattice fringes was 0.24 nm, which corresponds to the (311) plane of the cubic spinel AB2O4-type ZnCo2O4 [30,31,33,37].
To determine the average BET surface area and pore size distribution of the necklace-type ZnCo2O4 nanowires, a N2 adsorption/desorption measurement was carried out. As shown in Figure 5a,b, a representative type IV isotherm and a H3 hysteresis loop were observed, which can be assigned to the mesoporous characteristic of these necklace-type ZnCo2O4 nanowires [23,25,30,31,33,34,35]. The calculated average surface area and pore size distribution of the ZnCo2O4 nanowires were 25.33 m2 g−1 and 23.13 nm, respectively. In general, AB2O4-type cubic spinel structures with mesopores showed improved electrochemical supercapacitor properties, due to their high specific surface areas and increased electrolyte ion transport [14,15,23,25]. Therefore, we believe that our mesoporous necklace-type ZnCo2O4 nanowires also provide more electroactive sites and a facile diffusion of electrolyte ions during the intercalation/deintercalation process.
The electrochemical supercapacitor properties of as-prepared necklace-type ZnCo2O4 nanowires were studied by CV and GCD techniques in a traditional three-electrode system using 2M KOH aqueous solution. Figure 6a shows the CV curves of the necklace-type ZnCo2O4 nanowires electrode at various scan rates from 5 to 50 mV s−1 in a voltage window of –0.1~0.5 V. From the CV curves, we can clearly see a pair of oxidation and reduction peaks within 0.1~0.5 V, which are mainly due to the Co–O/Co–O–OH redox reactions during the intercalation and deintercalation process of OH ions [30,31,34,37]. The possible Faradaic redox reactions of the necklace-type ZnCo2O4 nanowires electrode can be given as follows [24,26]:
ZnCo 2 O 4 + OH + H 2 O   ZnOOH   +   2 CoOOH   +   e
CoOOH + OH   CoO 2   +   H 2 O   +   e
As the scan rate increased from 5 to 50 mV s−1, the area of the CV curves and redox peak current were increased, demonstrating the ideal capacity performance of the ZnCo2O4 electrode [23]. Moreover, redox peak shifts can be seen with the increase in scan rate. This effect may be due to the internal resistance of the ZnCo2O4 nanowires electrode during the Faradaic reactions [23,30]. Even at a high scan rate (50 mV s−1), the redox peaks were still detected with a similar shape, showing that the ZnCo2O4 nanowires electrode has excellent rate capability. Figure 6b shows a plot of the anodic and cathodic peak currents of the necklace-type ZnCo2O4 nanowires electrode with respect to the square root of the scan rate. A linear relationship can be observed, indicating that the Faradaic redox reactions of the necklace-type ZnCo2O4 nanowires electrode are diffusion-controlled processes [40].
The charge/discharge measurement was performed to determine the specific capacity, rate capability, and durability of the ZnCo2O4 nanowires electrode. Figure 6c displays the GCD curves of the ZnCo2O4 nanowires electrode at different current densities from 1 to 20 A g−1 within the potential window of 0~0.4 V. The GCD curves were almost similar to the CV profile, which fairly expressed the battery-type behavior of our ZnCo2O4 nanowires electrode. The specific capacity values (Cs) were calculated from the discharge time according to the following equation:
Cs = (I Δt)/(m)
where Cs, I, Δt, and m are the specific capacity (C g−1), the discharging current (A), the discharging time (s), and the mass of the electrode material (g), respectively.
The calculated specific capacity values were high as 439.6, 396.8, 373.2, 354.4, 336.0, 286.0, and 208.0 C g−1 at a current density of 1, 2, 3, 4, 5, 10, and 20 A g−1, respectively. For comparison with the available literature data for ZnCo2O4 nanomaterials, we calculated the specific capacitance (F g−1) values according to the Equation (S1) given in the Supplementary Material. The calculated specific capacitance value of necklace-type ZnCo2O4 nanowires at 1 A g−1 was high as 1099 F g−1. The specific capacitance values of our necklace-type ZnCo2O4 nanowires electrode is comparable to or higher than that of the reported ZnCo2O4 nanomaterials (Table S1). This comparative result suggests that our mesoporous necklace-type ZnCo2O4 nanowires electrode has rapid ion transport during the charge/discharge processes, which ultimately improves the electrochemical supercapacitor performance. The rate performance of the necklace-type ZnCo2O4 nanowires electrode was displayed in Figure 6d. It shows that the specific capacity of the ZnCo2O4 nanowires electrode decreased with increasing current densities from 1 to 20 A g−1. At higher current densities, the diffusion-controlled Faradaic reactions decreased, and surface-controlled Faradaic reactions increased, resulting in a lower specific capacity [40]. The specific capacity retention of the necklace-type ZnCo2O4 nanowires electrode was 47.31%, demonstrating good rate capability at high current densities.
Figure 7 shows the long-term cycling stability of the ZnCo2O4 nanowires electrode at a higher current density of 40 A g−1. After 5000 continuous charge/discharge cycles, the final specific capacity retention value was 84.82 %, suggesting that the ZnCo2O4 nanowires electrode has excellent long-term stability due to the significant number of electroactive sites for the facile transport of electrolyte ions. The Coulombic efficiency (η) of the ZnCo2O4 nanowires electrode was calculated using the following equation [21]:
η = td/tc × 100%
where td and tc are the discharge and charge time, respectively. The calculated Coulombic efficiency of the ZnCo2O4 nanowires electrode over the entirety of the cycles was 100%. This result demonstrates the excellent kinetic reversibility of the Faradaic reactions in the ZnCo2O4 nanowires electrode [23]. The CV, GCD, and stability data showed that the as-prepared necklace-type ZnCo2O4 nanowires electrode has outstanding electrochemical properties and can be considered a potential electrode for battery-type supercapacitors.
To further demonstrate the potential application of the necklace-type ZnCo2O4 nanowires electrode, a HSC device was fabricated with ZnCo2O4 nanowires (positive electrode) and activated carbon (negative electrode), which is represented as ZnCo2O4//AC. Before evaluating the electrochemical properties of the ZnCo2O4//AC HSC device in 2M KOH aqueous solution, the CV and GCD curves of the negative electrode were studied in a three-electrode configuration. The CV curves (Figure S2a) of AC at different scan rates from 10 to 50 mV s–1 clearly show a rectangular shape, indicating an EDLC-type negative electrode [41]. Figure S2b shows the linear and symmetrical GCD curves of the negative electrode at various current densities. The specific capacitances of the negative electrode Cs (F g−1) were calculated using the Equation (S1). The calculated specific capacitance values were 152.3, 131.0, 117.6, 106.8, and 97.5 F g–1 under current densities of 1, 2, 3, 4, and 5 A g–1, respectively.
The comparative CV curves of the positive (ZnCo2O4) and negative (AC) electrodes at a scan rate of 50 mV s−1 in a three-electrode system are presented in Figure 8a. From this CV data, a maximum potential window of 1.6 V was fixed for the ZnCo2O4//AC HSC device. The CV curves of the ZnCo2O4//AC HSC device were obtained at different scanning rates from 10 to 50 mV s−1 and displayed partial redox peaks (Figure 8b), indicating that battery-type and EDLC mechanisms were involved in our HSC device. Furthermore, the shape of the CV curves was maintained, even at a high scan rate, signifying excellent rate performance of the ZnCo2O4//AC HSC device. Figure 8c displays charge/discharge profiles of the ZnCo2O4//AC HSC device evaluated at various current densities from 1 to 5 A g−1. The shape of the GCD curves was consistent with the CV profile and the specific capacitance values estimated according to the Equation (S1). The calculated specific capacitance values were 117.7, 100.8, 91.1, 83.2, and 75.6 F g−1 at current densities of 1, 2, 3, 4, and 5 A g−1, respectively. Remarkably, the ZnCo2O4//AC HSC device delivered a high specific capacitance of 117.7 F g−1 at a current density of 1 A g−1, and retained 64.23 % even at a high current density of 5 A g−1 (75.6 F g−1), displaying outstanding rate performance (Figure 8d). Furthermore, the ZnCo2O4//AC HSC device showed exceptional cycling stability over the 10000 cycles at 20 A g−1 (~73.53% capacitance retention), demonstrating impressive durability (Figure 8e). Moreover, remarkable Coulombic efficiency of more than 97 % during the entire stability test was observed (Figure 8e, blue color), which revealed the excellent reversibility of our HSC device.
Based on the specific capacitance values, the energy density (E, Wh kg−1) and the power density (P, W kg−1) of the ZnCo2O4//AC device were calculated using the following equations [41]:
E = Cs ΔV2/7.2
P = 3600 E/Δt
where Cs (F g−1) is the specific capacitance, Δt (s) is the discharge time, and ΔV (V) is the potential window. Figure 8f shows the Ragone plot of the ZnCo2O4//AC HSC device. The Ragone plot showed a maximum energy density of 41.87 Wh kg−1 for the HSC device at a power density of 800.00 W kg−1. The ZnCo2O4//AC HSC device retained 64.2% of the energy density, even at a power density of 4000.00 W kg−1. These values were comparable, or even superior, to those of the previously reported data for the other HSC devices, such as ZnCo2O4 nanoparticles//AC (26.28 Wh kg−1 at 716 W kg−1) [26], ZnCo2O4 ultra-thin curved sheets//AC (20.31 Wh kg−1 at 855 W kg−1) [42], ZnCo2O4/H:ZnO//AC (3.75677 Wh kg−1 at 653.34 W kg−1) [43], NiCo2O4–MnO2//activated graphene (9.4 Wh kg−1 at 175 W kg−1) [44], NiCo2O4 nanosheets@hollow microrod arrays//AC (15.42 Wh kg−1 at 780 W kg−1 [45], NiCo2O4–rGO//AC (23.32 Wh kg−1 at 324.9 W kg−1) [46], NiCo2O4//RGO (23.9 Wh kg−1 at 650 W kg−1) [47], Ni–Co–oxide//AC (12.0 Wh kg−1 at 100 W kg−1) [48], CuCo2O4@CuCo2O4//AC (18 Wh kg−1 at 125 W kg−1) [49], and FS–CoFe2O4/GS//graphene sheet (28.4 Wh kg−1 at 900 W kg−1) [50]. Overall, the necklace-type ZnCo2O4 nanowires electrode is a promising candidate for electrochemical supercapacitor applications due to their unique morphology, mesoporous structure, high surface area, synergistic effect between Zn and Co metals, and higher electrical conductivity.

3. Experimental

Analytical grade chemicals, such as zinc acetate dihydrate (Zn(CH3COO)2.2H2O), cobalt acetate tetrahydrate (Co(CH3COO)2.4H2O), urea, ammonium fluoride, EG, potassium hydroxide, polyvinylidene difluoride, and N-methyl-2-pyrrolidinone were purchased from Sigma Aldrich Chemicals. Deionized water (DI H2O) was used throughout this work. Absolute ethanol (C2H5OH, 99.9%) was procured from DUKSAN Pure Chemicals, South Korea. Commercial NF (mass density: 346 g/m2, thickness: 1.6 mm) was obtained from MTI Korea.
Necklace-type ZnCo2O4 nanowires were synthesized by a simple hydrothermal method using water/EG mixed solvent and subsequent calcination treatment. First, ~0.219 g of Zn(CH3COO)2.2H2O, ~0.498 g of Co(CH3COO)2.4H2O, ~0.30 g of urea, and ~0.074 g of NH4F were dissolved in a mixed solution of 20 mL water and 20 mL EG with magnetic stirring for 30 min. After that, the clear pink-color solution was transferred to a 50 mL autoclave, and a hydrothermal treatment was carried out at 140 °C for 12 h in an oven. When the hydrothermal process was finished, the as-formed pink-color precipitates were washed with DI H2O and C2H5OH. Then, the cleaned precipitate was dried at 70 °C, and the necklace-type ZnCo2O4 nanowires precursor was successfully collected. Finally, the as-grown precursor was calcined to form necklace-type ZnCo2O4 nanowires at 400 °C for 2 h in a muffle furnace at a heating rate of 5 °C min−1. The detailed experimental photographs, physicochemical characterization, and electrochemical supercapacitor studies were shown in Supplementary Material (SM).

4. Conclusions

Mesoporous necklace-type ZnCo2O4 nanowires were synthesized by a simple one-pot hydrothermal method and subsequent annealing process in air. Crystalline phase, morphology, and surface area analyses revealed that the as-synthesized ZnCo2O4 nanowires possessed AB2O4-type cubic spinel structure and assembled by numerous mesoporous tiny building blocks of nanoparticles. The evaluated supercapacitor properties in a three-electrode system suggested that the necklace-type ZnCo2O4 nanowires had a high specific capacity (439.6 C g−1 at 1 A g−1), decent rate performance (47.31 % capacity retention at 20 A g−1), and long-term durability (84.82 % capacity retention after 5000 cycles). The superior electrochemical properties of the ZnCo2O4 nanowires electrode were attributed to the unique morphology, mesoporous structure, high surface area, and synergistic effect between the two metals. Successively, the HSC device fabricated with the ZnCo2O4 nanowires and AC electrodes delivered a reasonable specific capacitance (117.7 F g−1 at a current density of 1 A g−1), high energy density of 41.87 Wh kg−1, and power density of 800.0 W kg−1. This study on ZnCo2O4 nanowires highlights their great potential as an electrode material for supercapacitor applications due to their high electrochemical properties.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11121516/s1, Figure S1: Photographic images of (a) as-prepared pink-color reaction solution, (b) as-collected precipitates after hydrothermal treatment, (c) dried powder sample before calcination process, and (d) as-prepared necklace-type ZnCo2O4 powder sample after calcination; Figure S2: (a) and (b) CV and GCD curves of activated carbon (negative electrode) measured in 2M KOH solution, respectively; Table S1: Electrochemical performance of the necklace-type ZnCo2O4 nanowires compared to other ZnCo2O4-based nanostructures reported elsewhere.

Author Contributions

Data curation, J.A.R.; Funding acquisition, K.-S.A.; Investigation, J.A.R. and K.-S.A.; Methodology, J.A.R.; Supervision, K.-S.A.; Visualization, K.-S.A.; Writing—original draft, J.A.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation of Korea (NRF), grant number 2018R1D1A3B05042787.This work was also supported by the “Human Resources Program in Energy Technology” of the Korea Institute of Energy Technology Evaluation and Planning (KETEP), granted financial resources from the Ministry of Trade, Industry & Energy, Republic of Korea [No. 20204010600100].

Data Availability Statement

We are excluding this statement because the study did not report any data.

Acknowledgments

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education [grant number 2018R1D1A3B05042787]. This work was also supported by the “Human Resources Program in Energy Technology” of the Korea Institute of Energy Technology Evaluation and Planning (KETEP), granted financial resources from the Ministry of Trade, Industry & Energy, Republic of Korea [No. 20204010600100].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhao, C.; Zheng, W. A review for aqueous electrochemical supercapacitors. Front. Energy Res. 2015, 3, 23. [Google Scholar] [CrossRef] [Green Version]
  2. Gur, T.M. Review of electrical energy storage technologies, materials and systems: Challenges and prospects for large–scale grid storage. Energy Environ. Sci. 2018, 11, 2696–2767. [Google Scholar] [CrossRef]
  3. Wang, G.; Zhang, L.; Zhang, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ho, K.C.; Lin, L.Y. A review of electrode materials based on core–shell nanostructures for electrochemical supercapacitors. J. Mater. Chem. A 2019, 7, 3516–3530. [Google Scholar] [CrossRef]
  5. Guan, L.; Yu, L.; Chen, G.Z. Capacitive and non–capacitive faradaic charge storage. Electrochim. Acta. 2016, 206, 464–478. [Google Scholar] [CrossRef]
  6. Lukatskaya, M.R.; Dunn, B.; Gogotsi, Y. Multidimensional materials and device architectures for future hybrid energy storage. Nat. Commun. 2016, 7, 12647. [Google Scholar] [CrossRef]
  7. Yuan, C.; Wu, H.B.; Xie, Y.; Lou, X.W. Mixed transition–metal oxides: Design, synthesis, and energy–related applications. Angew. Chem. Int. Ed. 2014, 53, 1488–1504. [Google Scholar] [CrossRef]
  8. Cheng, J.P.; Zhang, J.; Liu, F. Recent development of metal hydroxides as electrode material of electrochemical capacitors. RSC Adv. 2014, 4, 38893–38917. [Google Scholar] [CrossRef]
  9. Yu, X.Y.; Lou, X.W. Mixed metal sulfides for electrochemical energy storage and conversion. Adv. Energy Mater. 2018, 8, 1701592. [Google Scholar] [CrossRef]
  10. Lu, T.; Dong, S.; Zhang, C.; Zhang, L.; Cui, G. Fabrication of transition metal selenides and their applications in energy storage. Coord. Chem. Rev. 2017, 332, 75–99. [Google Scholar] [CrossRef]
  11. Theerthagiri, J.; Durai, G.; Karuppasamy, K.; Arunachalam, P.; Elakkiya, V.; Kuppusami, P.; Maiyalagan, T.; Kim, H.S. Recent advances in 2–D nanostructured metal nitrides, carbides, and phosphides electrodes for electrochemical supercapacitors—A brief review. J. Ind. Eng. Chem. 2018, 67, 12–27. [Google Scholar] [CrossRef]
  12. Liu, C.F.; Liu, Y.C.; Yi, T.Y.; Hu, C.C. Carbon materials for high–voltage supercapacitors. Carbon 2019, 145, 529–548. [Google Scholar] [CrossRef]
  13. Suriyakumar, S.; Bhardwaj, P.; Grace, A.N.; Stephan, A.M. Role of polymers in enhancing the performance of electrochemical supercapacitors: A review. Batter. Supercaps 2021, 4, 571–584. [Google Scholar] [CrossRef]
  14. 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]
  15. Chen, D.; Wang, Q.; Wang, R.; Shen, G. Ternary oxide nanostructured materials for supercapacitors: A review. J. Mater. Chem. A 2015, 3, 10158–10173. [Google Scholar] [CrossRef]
  16. Cao, D.; Zheng, L.; Li, Q.; Zhang, J.; Dong, Y.; Yue, J.; Wang, X.; Bai, Y.; Tan, G.; Wu, C. Crystal phase–controlled modulation of binary transition metal oxides for highly reversible Li−O2 batteries. Nano Lett. 2021, 21, 5225–5232. [Google Scholar] [CrossRef]
  17. Zheng, J.H.; Zhang, R.M.; Yu, P.F.; Wang, X.G. Binary transition metal oxides (BTMO) (Co–Zn, Co–Cu) synthesis and high supercapacitor performance. J. Alloys Compd. 2019, 772, 359–365. [Google Scholar] [CrossRef]
  18. Adamson, W.; Bo, X.; Li, Y.; Suryanto, B.H.R.; Chen, X.; Zhao, C. Co–Fe binary metal oxide electrocatalyst with synergistic interface structures for efficient overall water splitting. Catal. Today 2020, 351, 44–49. [Google Scholar] [CrossRef]
  19. Chakrabarty, S.; Mukherjee, A.; Basu, S. RGO–MoS2 supported NiCo2O4 catalyst toward solar water splitting and dye degradation. ACS Sustain. Chem. Eng. 2018, 6, 5238–5247. [Google Scholar] [CrossRef]
  20. Nithya, J.S.M.; Do, J.Y.; Kang, M. Fabrication of flower–like copper cobaltite/graphitic–carbon nitride (CuCo2O4/g–C3N4) composite with superior photocatalytic activity. J. Ind. Eng. Chem. 2018, 57, 405–415. [Google Scholar]
  21. Wang, Q.; Du, J.; Zhu, Y.; Yang, J.; Chen, J.; Wang, C.; Li, L.; Jiao, L. Facile fabrication and supercapacitive properties of mesoporous zinc cobaltite microspheres. J. Power Sources 2015, 284, 138–145. [Google Scholar] [CrossRef]
  22. Mary, A.J.C.; Bose, A.C. Surfactant assisted ZnCo2O4 nanomaterial for supercapacitor application. Appl. Surf. Sci. 2018, 449, 105–112. [Google Scholar] [CrossRef]
  23. Chena, H.; Wang, J.; Han, X.; Liao, F.; Zhang, Y.; Han, X.; Xu, C. Simple growth of mesoporous zinc cobaltite urchin–like microstructures towards high–performance electrochemical capacitors. Ceram. Int. 2019, 45, 4059–4066. [Google Scholar] [CrossRef]
  24. Xu, L.; Zhao, Y.; Lian, J.; Xu, Y.; Bao, J.; Qiu, J.; Xu, L.; Xu, H.; Hua, M.; Li, H. Morphology controlled preparation of ZnCo2O4 nanostructures for asymmetric supercapacitor with ultrahigh energy density. Energy 2017, 123, 296–304. [Google Scholar] [CrossRef]
  25. Chen, H.; Du, X.; Sun, J.; Mao, H.; Wu, R.; Xu, C. Simple preparation of ZnCo2O4 porous quasi–cubes for high performance asymmetric supercapacitors. Appl. Surf. Sci. 2020, 515, 146008. [Google Scholar] [CrossRef]
  26. Bhagwan, J.; Hussain, S.K.; Yu, J.S. Aqueous asymmetric supercapacitors based on ZnCo2O4 nanoparticles via facile combustion method. J. Alloys Compd. 2020, 815, 152456. [Google Scholar] [CrossRef]
  27. Vijayakumar, S.; Nagamuthu, S.; Lee, S.H.; Ryu, K.S. Porous thin layered nanosheets assembled ZnCo2O4 grown on Ni–foam as an efficient electrode material for hybrid supercapacitor applications. Int. J. Hydrog. Energy. 2017, 42, 3122–3129. [Google Scholar] [CrossRef]
  28. Wu, H.; Lou, Z.; Yang, H.; Shen, G. A flexible spiral–type supercapacitor based on ZnCo2O4 nanorod electrodes. Nanoscale 2015, 7, 1921–1926. [Google Scholar] [CrossRef]
  29. Shi, R.; Zhang, Y.; Wang, Z. Facile synthesis of a ZnCo2O4 electrocatalyst with three–dimensional architecture for methanol oxidation. J. Alloys Compd. 2019, 810, 151879. [Google Scholar] [CrossRef]
  30. Rajesh, J.A.; Min, B.K.; Kim, J.H.; Kim, H.; Ahn, K.S. Cubic spinel AB2O4 type porous ZnCo2O4 microspheres: Facile Hydrothermal synthesis and their electrochemical performances in pseudocapacitor. J. Electrochem. Soc. 2016, 163, A2418–A2427. [Google Scholar] [CrossRef]
  31. Rajesh, J.A.; Min, B.K.; Kim, J.H.; Kang, S.H.; Kim, H.; Ahn, K.S. Facile hydrothermal synthesis and electrochemical supercapacitor performance of hierarchical coral–like ZnCo2O4 nanowires. J. Electroanal. Chem. 2017, 785, 48–57. [Google Scholar] [CrossRef]
  32. Wang, F.; Wang, X.; Liu, D.; Zhen, J.; Li, J.; Wang, J.; Zhang, H. High–performance ZnCo2O4@CeO2 core@shell microspheres for catalytic CO oxidation. ACS Appl. Mater. Interfaces 2014, 6, 22216–22223. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, T.; Liu, J.; Liu, Q.; Song, D.; Zhang, H.; Zhang, H.; Wang, J. Synthesis, characterization and enhanced gas sensing performance of porous ZnCo2O4 nano/microspheres. Nanoscale 2015, 7, 19714–19721. [Google Scholar] [CrossRef] [PubMed]
  34. Zhou, G.; Zhu, J.; Chen, Y.; Mei, L.; Duan, X.; Zhang, G.; Chen, L.; Wang, T.; Lu, B. Simple method for the preparation of highly porous ZnCo2O4 nanotubes with enhanced electrochemical property for supercapacitor. Electrochim. Acta 2014, 123, 450–455. [Google Scholar] [CrossRef]
  35. Bai, J.; Wang, K.; Feng, J.; Xiong, S. ZnO/CoO and ZnCo2O4 hierarchical bipyramid nanoframes: Morphology control, formation mechanism, and their lithium storage properties. ACS Appl. Mater. Interfaces 2015, 7, 22848–22857. [Google Scholar] [CrossRef]
  36. Moon, I.K.; Yoon, S.; Oh, J. Three-dimensional hierarchically mesoporous ZnCo2O4 nanowires grown on graphene/sponge foam for high-performance, flexible, all-solid-state supercapacitors. Chem. Eur. J. 2017, 23, 597–604. [Google Scholar] [CrossRef]
  37. Cheng, J.; Lu, Y.; Qiu, K.; Yan, H.; Hou, X.; Xu, J.; Han, L.; Liu, X.; Kim, J.K.; Luo, Y. Mesoporous ZnCo2O4 nanoflakes grown on nickel foam as electrodes for high performance supercapacitors. Phys. Chem. Chem. Phys. 2015, 17, 17016–17022. [Google Scholar] [CrossRef]
  38. Zhang, X.; Tian, F.; Gao, M.; Yang, W.; Yu, Y. L-Cysteine capped Mo2C/Zn0.67Cd0.33S heterojunction with intimate covalent bonds enables efficient and stable H2-Releasing photocatalysis. Chem. Eng. J. 2022, 428, 132628. [Google Scholar] [CrossRef]
  39. Zhang, X.; Tian, F.; Lan, X.; Liu, Y.; Yang, W.; Zhang, J.; Yu, Y. Building P-doped MoS2/g-C3N4 layered heterojunction with a dual-internal electric field for efficient photocatalytic sterilization. Chem. Eng. J. 2022, 429, 132588. [Google Scholar] [CrossRef]
  40. Rajesh, J.A.; Park, J.-H.; Quy, V.H.V.; Kwon, J.M.; Chae, J.; Kang, S.-H.; Kim, H.; Ahn, K.-S. Rambutan–like cobalt nickel sulfide (CoNi2S4) hierarchitecture for high–performance symmetric aqueous supercapacitors. J. Ind. Eng. Chem. 2018, 63, 73–83. [Google Scholar] [CrossRef]
  41. Lee, Y.-H.; Kang, J.S.; Jo, I.-K.; Sung, Y.-E.; Ahn, K.-S. Double–layer cobalt selenide/nickel selenide with web–like nanostructures as a high–performance electrode material for supercapacitors. J. Electroanal. Chem. 2021, 895, 115479. [Google Scholar] [CrossRef]
  42. Zhou, Y.; Chen, L.; Jiao, Y.T.; Li, Z.; Gao, Y. Controllable fabrication of ZnCo2O4 ultra–thin curved sheets on Ni foam for high–performance asymmetric supercapacitors. Electrochim. Acta 2019, 299, 388–394. [Google Scholar] [CrossRef]
  43. Boruah, B.D.; Maji, A.; Misra, A. Synergistic effect in the heterostructure of ZnCo2O4 and hydrogenated zinc oxide nanorods for high capacitive response. Nanoscale 2017, 9, 9411–9420. [Google Scholar] [CrossRef]
  44. Kuang, M.; Wen, Z.Q.; Guo, X.L.; Zhang, S.M.; Zhang, Y.X. Engineering firecracker–like beta–manganese dioxides@spinel nickel cobaltates nanostructures for high–performance supercapacitors. J. Power Sources. 2014, 270, 426–433. [Google Scholar] [CrossRef]
  45. Lu, X.-F.; Wu, D.-J.; Li, R.-Z.; Li, Q.; Ye, S.-H.; Tong, Y.-X.; Li, G.-R. Hierarchical NiCo2O4 nanosheets@hollow microrod arrays for high–performance asymmetric supercapacitors. J. Mater. Chem. A 2014, 2, 4706–4713. [Google Scholar] [CrossRef]
  46. Wang, X.; Liu, W.S.; Lu, X.; Lee, P.S. Dodecyl sulfate–induced fast faradic process in nickel cobalt oxide–reduced graphite oxide composite material and its application for asymmetric supercapacitor device. J. Mater. Chem. 2012, 22, 23114–23119. [Google Scholar] [CrossRef]
  47. Chen, H.; Jiang, J.; Zhang, L.; Qi, T.; Xia, D.; Wan, H. Facilely synthesized porous NiCo2O4 flowerlike nanostructure for high–rate supercapacitors. J. Power Sources 2014, 248, 28–36. [Google Scholar] [CrossRef]
  48. Tang, C.; Tang, Z.; Gong, H. Hierarchically porous Ni–Co oxide for high reversibility asymmetric full–cell supercapacitors. J. Electroanal. Chem. 2012, 159, A651–A656. [Google Scholar] [CrossRef]
  49. Zhang, Y.; Xu, J.; Zheng, Y.; Zhang, Y.; Hu, X.; Xu, T. Construction of CuCo2O4@CuCo2O4 hierarchical nanowire arrays grown on Ni foam for high–performance supercapacitors. RSC Adv. 2017, 7, 3983–3991. [Google Scholar] [CrossRef] [Green Version]
  50. Choi, J.; Seong, K.-D.; Kang, J.; Hwang, M.; Kim, J.M.; Jin, X.; Piao, P. Fluoride ion–mediated morphology control of fluorine–doped CoFe2O4/graphene sheet composites for hybrid supercapacitors with enhanced performance. Electrochim. Acta 2018, 279, 241–249. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of as-prepared necklace-type ZnCo2O4 nanowires.
Figure 1. XRD pattern of as-prepared necklace-type ZnCo2O4 nanowires.
Catalysts 11 01516 g001
Figure 2. XPS data of the necklace-type ZnCo2O4 nanowires. (a) Survey scan, (b) Zn 2p, (c) Co 2p, and (d) O 1s.
Figure 2. XPS data of the necklace-type ZnCo2O4 nanowires. (a) Survey scan, (b) Zn 2p, (c) Co 2p, and (d) O 1s.
Catalysts 11 01516 g002aCatalysts 11 01516 g002b
Figure 3. (a,b) and (c,d) Low-magnification and high-magnification FE–SEM images of the necklace-type ZnCo2O4 nanowires, respectively; (e) corresponding EDS spectrum.
Figure 3. (a,b) and (c,d) Low-magnification and high-magnification FE–SEM images of the necklace-type ZnCo2O4 nanowires, respectively; (e) corresponding EDS spectrum.
Catalysts 11 01516 g003
Figure 4. (a,b) Low-magnification FE–TEM images, (c) high-magnification FE–TEM image, and (d) HRTEM image of the necklace-type ZnCo2O4 nanowires. Inset image in (c) is SAED pattern.
Figure 4. (a,b) Low-magnification FE–TEM images, (c) high-magnification FE–TEM image, and (d) HRTEM image of the necklace-type ZnCo2O4 nanowires. Inset image in (c) is SAED pattern.
Catalysts 11 01516 g004
Figure 5. (a) BET N2 adsorption–desorption isotherm of the necklace-type ZnCo2O4 nanowires and (b) corresponding BJH pore size distribution curve.
Figure 5. (a) BET N2 adsorption–desorption isotherm of the necklace-type ZnCo2O4 nanowires and (b) corresponding BJH pore size distribution curve.
Catalysts 11 01516 g005
Figure 6. Electrochemical performance of the necklace-type ZnCo2O4 nanowires for the supercapacitors: (a) cyclic voltammogram curves measured at various scan rates from 5 to 50 mVs−1; (b) the relationship between cathodic/anodic peak current and the square root of the scan rate; (c) GCD curves at various current densities from 1 to 20 A g−1; and (d) specific capacity as a function of current density according to the data in (c).
Figure 6. Electrochemical performance of the necklace-type ZnCo2O4 nanowires for the supercapacitors: (a) cyclic voltammogram curves measured at various scan rates from 5 to 50 mVs−1; (b) the relationship between cathodic/anodic peak current and the square root of the scan rate; (c) GCD curves at various current densities from 1 to 20 A g−1; and (d) specific capacity as a function of current density according to the data in (c).
Catalysts 11 01516 g006
Figure 7. Electrochemical cycling stability of the necklace-type ZnCo2O4 nanowires measured at 40 A g−1 for 5000 cycles.
Figure 7. Electrochemical cycling stability of the necklace-type ZnCo2O4 nanowires measured at 40 A g−1 for 5000 cycles.
Catalysts 11 01516 g007
Figure 8. (a) CV curves of AC (negative potential) and ZnCo2O4 nanowires (positive potential) at 50 mV s−1; (b) CV curves of HSC device at different scan rates; (c) GCD curves at different current densities; (d) plot of specific capacitances vs. different current densities; (e) cycling performances of the HSC device for 10000 cycles at 20 A g−1; and (f) Ragone plot of HSC device at different current densities.
Figure 8. (a) CV curves of AC (negative potential) and ZnCo2O4 nanowires (positive potential) at 50 mV s−1; (b) CV curves of HSC device at different scan rates; (c) GCD curves at different current densities; (d) plot of specific capacitances vs. different current densities; (e) cycling performances of the HSC device for 10000 cycles at 20 A g−1; and (f) Ragone plot of HSC device at different current densities.
Catalysts 11 01516 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rajesh, J.A.; Ahn, K.-S. Facile Hydrothermal Synthesis and Supercapacitor Performance of Mesoporous Necklace-Type ZnCo2O4 Nanowires. Catalysts 2021, 11, 1516. https://doi.org/10.3390/catal11121516

AMA Style

Rajesh JA, Ahn K-S. Facile Hydrothermal Synthesis and Supercapacitor Performance of Mesoporous Necklace-Type ZnCo2O4 Nanowires. Catalysts. 2021; 11(12):1516. https://doi.org/10.3390/catal11121516

Chicago/Turabian Style

Rajesh, John Anthuvan, and Kwang-Soon Ahn. 2021. "Facile Hydrothermal Synthesis and Supercapacitor Performance of Mesoporous Necklace-Type ZnCo2O4 Nanowires" Catalysts 11, no. 12: 1516. https://doi.org/10.3390/catal11121516

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