Next Article in Journal
Multifunctional Hybrid Nanozymes for Magnetic Enrichment and Bioelectrocatalytic Sensing of Circulating Tumor RNA during Minimal Residual Disease Monitoring
Next Article in Special Issue
Hydrothermally Derived Mg-Doped TiO2 Nanostructures for Enhanced H2 Evolution Using Photo- and Electro-Catalytic Water Splitting
Previous Article in Journal
Catalytic Activity of Ni Based Materials Prepared by Different Methods for Hydrogen Production via the Water Gas Shift Reaction
Previous Article in Special Issue
Augmenting the Photocatalytic Performance of Direct Z-Scheme Bi2O3/g-C3N4 Nanocomposite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogen Evolution Reaction Activities of Room-Temperature Self-Grown Glycerol-Assisted Nickel Chloride Nanostructures

1
School of Electrical Engineering, Korea University, Seoul 02841, Republic of Korea
2
Department of Chemical Engineering (BK21 FOUR), Dong-A University, 37 Nakdong-daero, Saha-gu, Busan 49315, Republic of Korea
3
Department of Future Energy Convergence, Seoul National University of Science & Technology, 232 Gongneung-ro, Nowon-gu, Seoul 01811, Republic of Korea
4
School of Physical Sciences, Swami Ramanand Teerth Marathwada University, Nanded 431606, India
5
Vidya Pratishtan’s Arts, Science and Commerce College, Baramati 413133, India
6
Henan Key Laboratory of Materials on Deep-Earth Engineering, School of Materials Science and Engineering, Henan Polytechnic University, Jiaozuo 454000, China
*
Authors to whom correspondence should be addressed.
Submission received: 13 December 2022 / Revised: 28 December 2022 / Accepted: 7 January 2023 / Published: 12 January 2023
(This article belongs to the Special Issue Advanced Nanostructured Materials for Modern Catalysis Applications)

Abstract

:
Three-dimensional nanomaterials of desired structural/morphological properties and highly porous with a high specific surface area are important in a variety of applications. In this work, glycerol-mediated self-growth of 3-D dandelion flower-like nickel chloride (NiCl2) from nickel-foam (NiF) is obtained for the first time using a room-temperature (27 °C) processed wet chemical method for electrocatalysis application. Glycerol-mediated self-grown NiCl2 flowers demonstrate an excellent electrocatalytic performance towards the hydrogen evolution reaction (HER), which is much superior to the NiF (303 mV) and NiCl2 electrode prepared without glycerol (208 mV) in the same electrolyte solution. With a Tafel slope of 41 mV dec−1, the NiCl2 flower electrode confirms improved reaction kinetics as compared to the other two electrodes, i.e., NiF (106 mVdec−1) and NiCl2 obtained without glycerol (56 mV dec−1). The stability of the glycerol-based NiCl2 electrode has further been carried out for 2000 cycles with the overpotential diminution of just 8 mV, approving an electrocatalyst potential of glycerol-based NiCl2 electrode towards HER kinetics. This simple and easy growth process involves nucleation, aggregation, and crystal growth steps for producing NiCl2 nanostructures for electrocatalytic water splitting application through the HER process.

1. Introduction

The large-scale reliance on fossil fuels, and their predictable reduction, is of serious concern not only for worldwide environmental pollution but also the energy crisis, which has encouraged researchers to look for new renewable and clean energy resources [1,2,3,4]. Hydrogen is a viable alternative to fossil fuels [5,6]. After water electrolysis, high-purity hydrogen evolution reactions (HER) take place at the cathode, while the anode produces oxygen evolution reactions (OER). Both HER and OER kinetics are essential for the overall water-splitting process. Because the hydrogen adsorption free energy on these metal surfaces is thermally neutral, platinum group metals are known to be the best HER catalysts [6,7]. Due to the high cost of platinum, its use as an HER catalyst has been substantially limited [8,9,10,11,12,13,14,15]. The development and synthesis of HER catalysts with high catalytic activity are some of the top issues for researchers across the globe. Versatile nickel-based electrodes, including NiS2, NiO, NiSe, etc., have widely been envisaged for HER as well as OER electrocatalysts [16,17,18,19,20,21,22]. An electrocatalyzer assembly consisting of a cathode (platinum–carbon) for the HER with zero overpotential and an anode (ruthenium oxide/iridium oxide) for the OER with a 1.23 V overpotential for water splitting (at 10 mA cm−2 current density) is considered to be a standard measurement [7]. To-date, very few reports describing the use of self-supported Ni-based electrocatalysts in HER studies are known. For example, Zhu et al. revealed an in-situ chemical etching (ICE) approach for obtaining 3D Ni3S2 superstructures (needle array and leaf pattern) on a nickel-foam (NF) substrate for probable HER/OER catalyst with the lowest overpotential of 182/340 mV [23]. Shinde et al. have grown promising NiF2 nanorod arrays directly on a 3D NF via a reduction procedurethat obtained172 mV HER overpotential [24]. According to Wang et al., a NiS/NiF electrode delivered 158 mV overpotential for the HER and 50 mV overpotential for the OER [25]. These processes produce only one form of nanostructure, and their sophisticated customization necessitates the precursor composition ratio variation [26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41]. If the product is in powder form, several annealing courses are required to remove the binders, which eventually can reduce electrochemical performance by obstructing the porosity of the employed matrix or host materials.
This issue has been resolved in this manuscript to some limit by self-growing dandelion flower-like NiCl2@NiF nanostructured electrode material at room-temperature (25°C) in the presence of glycerol. The use of glycerol demonstrates a significant impact on the shape, surface area, and electrochemical performance of 3D NiCl2@NiF nanostructure over pristine NiF and NiCl2@NiF (obtained without glycerol) electrodes. Glycerol in a precursor solution can aid in the formation of a well-controlled dandelion flower-like nanostructure. We present a simple and scalable self-grown wet chemical, i.e., hydrothermal synthesis method, for obtaining the NiCl2@NiF electrode, which is intended for water splitting with a low overpotential of 176 mV for HER electrocatalysis with a Tafel slope of 41 mV dec−1 and long-term stability up to 2000 cycles.

2. Results

2.1. Morphology Evolution Studies

The FE-SEM depictions of the NiCl2@NiF electrodes obtained with and without glycerol are shown in Figure 1. From Figure 1a, several nanorods are interlocked with each other and well-cover the 3D NiF skeleton. The diameters of these nanorods are found to be in the range of 1–3 µm, and the average length is 10 (±4) µm. The compact nanorod structure of this electrode may offer a smaller surface area and hence may seriously impact on its electrochemical performance. Moreover, the ion percolation is complicated with the present form of nanorods. So, it is hopefully possible to control the NiCl2 nanostructure using glycerol as a surfactant.
As shown in Figure 1b, well-controlled parallel and conical arranged nanorods are grown together to form dandelion flowers on the NiF (see Figure 1(b-1) for a closer view). Some NiF spaces are exposed, suggesting the selective growth of NiCl2 for obtaining dandelion flower-like architecture from NiF without adding any nickel salt. Compared with Figure 1a, the dandelion flowers are uniformly distributed on the NiF substrate surface, suggestinga decrease in the self-aggregation of NiCl2, which is totally disappeared (Figure 1(b,b-1)). Vertically inclined nanorods are confirmed in the dandelion flowers. These upright standing nanorods have an average height of 10 (±3) µm with a separation distance of 5–10 µm (see Figure 1(b-2)), which will effectively help the deep penetration of the electrolyte ion for majority access and hassle-free charge/mass transportation. Figure S1 shows the related FE-SEM-assisted EDX elemental mapping measurements of the NiCl2 electrodes obtained with and without glycerol (a, a-2 and b, b-2). Both Ni and Cl elements are evenly distributed over the NiCl2 electrode surface (with and without glycerol) in a 1:2 stoichiometric ratio, indicating that the 3D NiF surface has converted to NiCl2. The TEM image of the NiCl2 sample obtained with glycerol, as shown in Figure 1c, clearly displays a dandelion flower-like nanostructure, which is constructed by several nanorods. As screened at higher magnification images, these interconnected nanorods are well-distinguished from one another. The HR-TEM image shown in Figure 1c is a TEM image of the NiCl2 obtained with glycerol where a dandelion flower-like morphology, consistent with the FE-SEM image, is constructed by the nanorods, which is consistent with the above FE-SEM observations as shown in the Figure 1b.
The formation of the NiCl2 is confirmed by HR-TEM imaging (Figure 1(c,c-1)) as the visible lattice fringes with an interlayer spacing of 0.271 nm, in close agreement with the XRD result (discussed later), are observed. The same part of the nanorods, as seen in the HR-TEM image, does not have clearly reflected lattice fringes, suggesting the presence of an amorphous or non-crystallineNiCl2. The well-ordered circular arrays of lattice points in the selected-area electron diffraction pattern (inset of Figure 1(c-1)) are noted, indicating the significant single crystallinity of the NiCl2 nanorods [32]. The above results suggest that the formation process of the NiCl2 dandelion flowers strongly depends on the use of glycerol in the precursor solution, which has been described as a growth process (see Scheme 1), in which the use of an appropriate amount of glycerol with acetone to form a uniform and transparent mixed solution can act as a micro-emulsion. This can be confirmed to be a Tyndall-effect-free and miscible solvent under thermodynamic equilibrium [40], so it is reasonable to control the self-aggregation of a NiCl2 dandelion flower in precursor solution through the following reaction steps. The NiF is initially oxidized to create nickel ions on the surface when dipped in a homogeneous solution containing acetone, glycerol, and hydrochloric acid (Equation (1)). Glycerol surfactant and hydrochloric acid are disintegrated in acetone at the same time, yielding glycerol, chlorine, and hydrogen free radical ions (see Equations (2) and (3)). Furthermore, Ni+ ions serve as nucleation sites, attracting glycerol surfactant molecules to create an intermediate complex known as Ni[O-CH2-CH(OH)-CH2-OH]3. Furthermore, this intermediate combination transfers electrons to Cl- ions, converting NiCl2 primary nanorods on the NiF surface, which serves as a basic mother root for the growth of a dandelion flower NiCl2 nanostructure (see bottom reaction pathway of schematic 1). As time proceeds, seed nuclei will grow in the vertical direction without aggregation under the control of the glycerol surfactant (see Scheme 1), forming dandelion flower-type NiCl2, which is summarized in flowing reaction kinetics [32]. On the other hand, in the case of the NiCl2 sample prepared without glycerol surfactant, the growth of the NiCl2 nanorods is out of control; hence the formation of scattered and irregular nanorods over the NiF surface (see upper reaction pathway in Scheme 1) is seen. A very similar type of growth mechanism for the Fe2O3 nanoparticles obtained in the presence of glycerol has been reported elsewhere [32,40].
Ni foam     Ni 2 +
[ HO CH 2 CH ( OH ) CH 2 OH ] [ HO CH 2 CH ( OH ) CH 2 O ] + H +
HCl H + + Cl
Ni 2 + + 3 [ HO CH 2 CH ( OH ) CH 2 O ]   Ni [ O CH 2 CH ( OH ) CH 2 OH ] 3
Ni [ O CH 2 CH ( OH ) CH 2 OH ] 3 + 2 Cl   NiCl 2

2.2. Structural Elucidation

Figure 2 confirms the X-ray diffraction patterns of the NiCl2 electrode materials on the NiF obtained with and without glycerol (A and B). Two intensive peaks, marked as symbol ‘∆’, of metallic nickel (see Figure 2a(C)) in all patterns are recognized ((JCPDS card no. 04-0850)). The newly reflected diffraction peaks denoted with a ‘*’ symbol and positioned at a different 2-theta location according to JCPDS card no. 01-072-0044 are assigned to NiCl2. According to the XRD results, the intensity of the diffraction peak in the glycerol-meditated NiCl2 sample is higher compared to the non-glycerol-involved sample, showing higher crystallinity in the prior over the latter. The detailed surface optimization of the glycerol-mediated NiCl2 electrode was obtained using XPS, and the corresponding results are presented in Figure 2b–d. The presence of Ni and Cl elements is confirmed by the survey spectrum (Figure 2b). The deconvoluted Ni 2p high-resolution spectrum (Figure 2c) adduces two major peaks at 856.3 and 874.1 eV for Ni 2p3/2 and Ni 2P1/2, respectively, with a spin energy separation distance of 17.8 eV. The satellite peaks at 862.2 and 882.5 eV, which are in good agreement with previous findings, are also evidenced [32].
The prominent peak obtained at 198.56 eV in the Cl high-resolution spectrum indicates that NiCl2 develops on the NiF skeleton during the self-growth process at ambient temperature. The specific surface area (IV-type isotherm and H3 hysteresis type), pore-size distribution, and pore surface area estimations of the optimized with glycerol-mediated NiCl2 electrode determined by N2 adsorption–desorption measurements are, respectively, 15.98 m2g−1, 7.4 nm, and 47.66 m2g−1 (Figure 2e) [32]. The nature of the adsorption and the desorption branches is assigned to the mesoporous behavior [32].

2.3. HER Activity

Furthermore, NiCl2 electrocatalysts i.e., prepared with and without glycerol were utilized to study the HER water catalysis activity in 1.0 M KOH electrolyte solution (Figure 3) and compared with pristine NiF.
The polarization curves of the as-prepared with and without glycerol-mediated NiCl2 electrode materials are shown in Figure 3a. Both electrode materials demonstrate considerable HER activity. As compared to the nanorod-type NiCl2 (208 mV and pristine NiF (303 mV) electrodes, the ‘with dandelion flower-like NiCl2 electrode reveals an enhanced electrocatalytic activity with as small as 176 mV overpotential, proving an enhancement in electrocatalytic activity due to the change in morphology caused by the glycerol incorporation of the NiCl2 electrode for HER. The Tafel slopes for the corresponding electrode materials, as shown in Figure 3b, are lowered from 41 mV dec−1 to 57 mV dec−1 when the HER rate is increased (see supporting information (SI) for more details on the equations used for computation). The Volmer–Heyrovsky mechanism is obeyed on the electrode surface, as evidenced by the tiny Tafel slope of 41 mVdec−1 [39,40,41,42]. As shown in Figure 3c, the electrochemical impedance spectroscopy (EIS) analysis performed for the three aforementioned electrodes from the 100 kHz to 10 kHz frequency range with a potential amplitude of 5 mV adduces clear and visible difference. Compared to the NiCl2 obtained without glycerol and pristine NiF electrodes, the glycerol-mediated NiCl2 electrode has low equivalent series resistance (2.62 Ω) and 1.78 Ω charge transfer resistance, the NiCl2 electrode obtained with glycerol has improved electron mobility and electrocatalytic activity [24,41]. The cyclability test of the dandelion flower-type NiCl2 electrocatalyst for HER was further examined by successive CV plots in a three-electrode system at a scan rate of 10 mV s−1 for 2000 cycles (Figure 3d). A negligible decrease of 8 mV has been noticed in the 1st and 2000th polarization curves after long-term HER activity scanning, demonstrating its excellent HER catalytic ability with moderate chemical stability. The inset of Figure 3d presents a genuine photograph of the three-electrode system used for the electrocatalytic study, where a clear progression of oxygen bubble evaluation, indicating strong HER activity, is confirmed. This obtained over-potential is comparable with those reported in the literature (Table 1). According to the electrocatalysis results, the glycerol-containing NiCl2 electrode exhibits excellent catalytic activity and stability towards the HER for several reasons (Scheme 2), which are as follows; (i) the porous nano-array structure of dandelion flower with a high surface area results in a greater number of active sites for electrocatalytic activity, (ii) the direct binder-free growth of mesoporous dandelion flower-like NiCl2 on 3D conductive NiF substrate provides good connectivity with zero dead volume, (iii) the direct deposition of the NiCl2 dandelion flower on the NiF surface and the close interconnection of these nanorods, forming a network that maintains good structural stability with efficient electron transport, and (iv) the room-temperature NiCl2 synthesis approach as shown in Scheme 2 clearly depicts the important features of the dandelion flower-type structure for better catalytic performance.

3. Materials and Methods

3.1. Chemicals

NiF, with a pore density of 110 pore size and a mass density of 320 g m−2, was obtained (Artenano Company Limited, Central, Hong Kong). Concentrated hydrochloric acid (HCl, 37%) was obtained from Daejung Chemicals (Seongnam-si, Republic of Korea). Acetone and anhydrous ethanol standard solutions were purchased from SK Chemicals (Seongnam-si, Republic of Korea). Deionized (DI) water, obtained from (Millipore Darmstadt, Darmstadt, Germany), was used throughout the experiments as a solvent. The NiCl2 dandelion flowers were fabricated under optimized conditions on NiF at room-temperature. A NiF (2.00 × 6.00 cm2) was ultrasonicated for 20 minin ethanol, acetone, and finally, water to ensure that the NiF surface was thoroughly cleaned and then dried in a vacuum oven at 60 °C for 12 h using a process described elsewhere [32]. The NiF was then placed in a glass falcon tube (25 mL) that was filled with 20 mL acetone and 0.1 M HCl, in addition to a few drops of glycerol, which was stirred continuously for 10 min. The glass falcon tube was then sealed with a lid and left at room-temperature for 4 h. Interestingly, after a few hours, the NiF was changed from a polished silvery texture to a rough texture with a bright green tint (see on-ground experimental digital images displayed in Scheme 1). Finally, the NiF was removed and properly washed with deionized water multiple times before being thoroughly dried in a vacuum at 60 °C for 12 h and labeled as electrode ‘with glycerol’. For comparison, a NiCl2 sample was obtained under identical conditions but just without the addition of glycerol into the for mentioned solution, which was referred to as ‘without glycerol’.

3.2. Characterizations

The structural interpretation and surface morphology evolution studies for the NiCl2 electrodes, i.e., with and without glycerol, were carried out using X-ray diffraction (XRD, D8-Discovery Bruker, 40 kV, 40 mA, Cu Kα, λ = 1.5406 Å) patterns and field-emission scanning electron microscopy (FE-SEM, Hitachi, S-4800, 15 kV) digital images, equipped with an energy-dispersive X-ray spectroscopy (EDX) measurement unit. Both transmission and high-resolution transmission electron microscopy (HRTEM, JEOL 2100F) images were recorded over the optimized NiCl2 nanostructured electrode. The X-ray photoelectron spectroscopy (XPS, VG Scientifics ESCALAB250) measurement was carried out to analyze the chemical bonding status of the NiCl2 nanostructures. The surface area and pore-size distribution for the NiCl2 nanostructures were obtained from Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) plots.

3.3. HER Confirmation

Both NiCl2 nanostructure electrocatalysts were envisaged in a three-electrode system (in addition to NiF as a reference) in the presence of a Hg/HgO reference electrode and a platinum counter electrode. A nitrogen purge was used for 10 min before the trials to normalize the system in terms of external and internal oxygen.

3.4. Formulae Used

The formulae used for estimating the overpotential and the Tafel slope are provided in the supporting information.

4. Conclusions

In summary, the glycerol-mediated self-growth of dandelion flower-type NiCl2 nanostructure is proposed at room-temperature viaa soft chemical approach. After sorting out the phase formation confirmation through the XRD pattern analysis. Change in the dandelion flower-type surface appearance with the 15.98 m2g−1 surface area of the as-obtained NiCl2 on glycerol loading has been evidenced by the FE-SEM images and BJH curves, respectively. Furthermore, the HER activities of the aforementioned electrodes studied for water-splitting applications demonstrate the feasibility of NiCl2 as an HER electrode material in the water-splitting process. The NiCl2 electrode prepared in the presence of glycerol demonstrates an overpotential of 176 mV, while the electrode prepared without glycerol confirms the overpotential of 208 mV, suggesting the impact of surface modification on HER performance. The NiCl2 dandelion flower-like electrode reveals the superior reaction kinetics with the small Tafel slope of 41 mV dec−1 as compared to that of the nanorod-type NiCl2 (57 mV dec−1) electrode obtained without glycerol. Moreover, the NiCl2 dandelion flower electrode is found to be stable even after 2000 operation cycles with a minute decrease of 8 mV in overpotential. The considerable HER activity of the NiCl2 dandelion flower-type electrode is attributed to its high surface area, flower-like morphology, and low internal resistance due to uncomplicated charge transport.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal13010177/s1, Figure S1: Figure (a-a-2, b-b-2) surface Ni and Cl element mapping and EDX spectra of with and without glycerol NiCl2 at different magnifications.

Author Contributions

Conceptualization, N.M.S., B.G.G., N.M.S. and S.D.R.; methodology, N.M.S. and S.H.N.; software, R.J.D.; validation, N.M.S., B.G.G. and Q.X.; formal analysis, N.M.S. and Q.X.; investigation, N.M.S. and S.D.R.; data curation, N.M.S.; writing—original draft preparation, R.S.M., J.-S.K., J.J.P.; review and editing, visualization, supervision, R.S.M., J.-S.K., J.J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korean Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea government (MOTIE) (20224000000400, Cultivating Specialized Professionals for Core Materials in the ESS Secondary Battery Industry (Middle Market Enterprise). This work was supported by Dong-A University Foundation Grant in 2021.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, G.; Wang, P.; Li, H.; Hu, B.; Sun, Y.; Huang, R.; Liu, L. Spin-state reconfiguration induced by alternating magnetic field for efficient oxygen evolution reaction. Nat. Commun. 2021, 12, 4827. [Google Scholar] [CrossRef]
  2. Anantharaj, S.; Rao Ede, S.; Sakthikumar, K.; Karthick, K.; Mishra, S.; Kundu, S. A Recent Trends and Perspectives in Electrochemical Water Splitting with an Emphasis on Sulfide, Selenide, and Phosphide Catalysts of Fe, Co, and Ni: A Review. ACS Catal. 2016, 6, 8069–8097. [Google Scholar] [CrossRef]
  3. Zhou, G.; Wang, P.; Hu, B.; Shen, X.; Liu, C.; Tao, W.; Huang, P.; Liu, L. Spin-related symmetry breaking induced by half-disordered hybridization in BixEr2-xRu2O7 pyrochlores for acidic oxygen evolution. Nat. Commun. 2022, 13, 4106. [Google Scholar] [CrossRef]
  4. Suen, N.; Hung, S.; Quan, Q.; Zhang, N.; Xu, Y.; Chen, M. Electrocatalysis for the Oxygen Evolution Reaction: Recent Development and Future Perspectives. Chem. Soc. Rev. 2017, 46, 337–365. [Google Scholar] [CrossRef]
  5. Zhou, G.; Shan, Y.; Wang, L.; Hu, Y.; Guo, J.; Hu, F.; Shen, J.; Gu, Y.; Cui, J.; Liu, L.; et al. Photoinduced semiconductor-metal transition in ultrathin troilite FeS nanosheets to trigger efficient hydrogen evolution. Nat. Commun. 2019, 10, 399. [Google Scholar] [CrossRef] [Green Version]
  6. Fu, Y.; Shan, Y.; Zhou, G.; Long, L.; Wang, L.; Yin, K.; Guo, J.; Shen, J.; Liu, L.; Wu, X. Electric strain in dual metal Janus nanosheets induces structural phase transition for efficient hydrogen evolution. Joule 2019, 3, 2955–2967. [Google Scholar] [CrossRef]
  7. Chaudhari, N.; Jin, H.; Kim, B.; Lee, K. Nanostructured Materials on 3D Nickel Foam as Electrocatalysts for Water Splitting. Nanoscale 2017, 9, 12231–12247. [Google Scholar] [CrossRef]
  8. Giri, A.; Park, G.; Yang, H.; Pal, M.; Kwak, J.; Jeong, U. Synthesis of 2D Metal Chalcogenide Thin Films through the Process Involving Solution-Phase Deposition. Adv. Mater. 2013, 30, 1707577. [Google Scholar] [CrossRef]
  9. Chou, S.; Lin, J. Cathodic Deposition of Flaky Nickel Sulfide Nanostructure as an Electroactive Material for High-Performance Supercapacitors. J. Electrochem. Soc. 2013, 160, D178–D182. [Google Scholar] [CrossRef]
  10. Yu, S.; Yoshimura, M. Ferrite/Metal Composites Fabricated by Soft SolutionProcessing. Adv. Funct. Mater. 2002, 12, 277–285. [Google Scholar] [CrossRef]
  11. Yang, S.; Yao, H.; Gao, M.; Yu, S. Monodisperse Cubic Pyrite NiS2 Dodecahedrons and Microspheres Synthesized by a Solvothermal Process in a Mixed Solvent: Thermal Stability and Magnetic Properties. CrystEngComm 2009, 11, 1383–1390. [Google Scholar] [CrossRef]
  12. Li, X.; Li, Q.; Wu, Y.; Rui, M.; Zeng, H. Two-Dimensional, Porous Nickel–Cobalt Sulfide for High-Performance Asymmetric Supercapacitors. ACS Appl. Mater. Interf. 2015, 7, 19316–19323. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, Q.; Lu, Z.; Liu, J.; Lei, X.; Chang, Z.; Luo, L.; Sun, X. Metal oxide and Hydroxide Nanoarrays: Hydrothermal Synthesis and Applications as Supercapacitors and Nanocatalysts. Prog. Nat. Sci. Mater. Int. 2013, 23, 351–366. [Google Scholar] [CrossRef] [Green Version]
  14. Liu, Y.; Li, Y.; Kang, H.; Jin, T.; Jiao, L. Design, Synthesis, and Energy-related Applications of Metal Sulfides. Mater. Horiz. 2016, 3, 402–421. [Google Scholar] [CrossRef]
  15. Guo, K.; Yang, F.; Cui, S.; Chen, W.; Mi, L. Controlled synthesis of 3D Hierarchical NiSe Microspheres for High-performance Supercapacitor Design. RSC Adv. 2016, 6, 46523–46530. [Google Scholar] [CrossRef]
  16. Lin, H.; Liu, F.; Wang, X.; Ai, Y.; Yao, Z.; Chu, L.; Han, S.; Zhuang, X. Graphene-Coupled Flower-Like Ni3S2 for a Free-Standing 3D Aerogel with an Ultra-High Electrochemical Capacity. Electrochim. Acta 2016, 191, 705–715. [Google Scholar] [CrossRef]
  17. Zhang, Y.; Qian, L.; Zhao, W.; Li, X.; Huang, X.; Mai, X.; Wang, Z.; Shao, Q.; Yan, X.; Guo, Z. Highly Efficient Fe-N-C Nanoparticles Modified Porous Graphene Composites for Oxygen Reduction Reaction. J. Electrochem. Soc. 2018, 165, H510–H516. [Google Scholar] [CrossRef]
  18. Krishnamoorthy, K.; Kumar Veerasubramani, G.; Radhakrishnan, S.; Kim, S. One pot Hydrothermal Growth of Hierarchical Nanostructured Ni3S2 on Ni foamfor Supercapacitor Application. Chem. Eng. J. Chem. 2014, 251, 116–122. [Google Scholar]
  19. Zhuo, M.; Zhang, P.; Chen, Y.; Li, Q. Facile Construction of Graphene-likeNi3S2 Nanosheets Through the Hydrothermally Assisted Sulfurization of NickelFoam and Their Application as Self-supported Electrodes for Supercapacitors. RSC Adv. 2015, 5, 25446–25449. [Google Scholar] [CrossRef]
  20. Li, W.; Wang, S.; Xin, L.; Wu, M.; Lou, X. Single-crystal β-NiS Nanorod Arrays with a Hollow-structured Ni3S2 Framework for Supercapacitor Applications. J. Mater. Chem. A 2016, 4, 7700–7709. [Google Scholar] [CrossRef]
  21. Li, J.; Wang, S.; Xiao, T.; Tana, X.; Xiang, P.; Jiang, L.; Deng, C.; Li, W.; Li, M. Controllable preparation of Nanoporous Ni3S2 films by Sulfuration of nickel foam as promising asymmetric supercapacitor electrodes. Appl. Surf. Sci. 2017, 420, 919–926. [Google Scholar] [CrossRef]
  22. Zhang, D.; Li, J.; Luo, J.; Xu, P.; Wei, L.; Zhou, D.; Xu, W.; Yuan, D. Ni3S2 nanowires grown on nickel foam as an efficient bifunctional electrocatalyst for water splitting with greatly practical prospects. Nanotechnology 2018, 29, 245402. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, T.; Zhu, L.; Wang, J.; Ghim, H. In Situ Chemical Etching of Tunable 3D Ni3S2 Superstructures For Bifunctional Electrocatalysts For Overall Water Splitting. J. Mater. Chem. A 2016, 4, 13916–13922. [Google Scholar] [CrossRef]
  24. Shinde, N.; Raut, S.; Ghule, B.; Gunturu, K.; Pak, J.; Mane, R. Recasting Ni-foam into NiF2 nanorod arrays via a hydrothermal process for hydrogen evolution reaction application. Dalton Trans. 2021, 50, 6500–6505. [Google Scholar] [CrossRef] [PubMed]
  25. Zhu, W.; Yue, X.; Zhang, W.; Yu, S.; Zhang, Y.; Wang, J.; Wang, J. Nickel sulfide Microsphere Film on Ni foam as an Efficient Bifunctional Electrocatalyst for Overall Water Splitting. J. Chem. Commun. 2016, 52, 1486–1489. [Google Scholar] [CrossRef] [PubMed]
  26. Maiaugree, W.; Tangtrakarn, A.; Lowpa, S.; Ratchapolthavisin, N.; Amornkitbamrung, V. Facile Synthesis of Bilayer Carbon/Ni3S2Nanowalls for a Counter Electrode of Dye-sensitized Solar Cell. Electrochim. Acta 2015, 174, 955–962. [Google Scholar] [CrossRef]
  27. Li, X.; Chen, Y.; Zou, J.; Zeng, X.; Zhou, L.; Huang, H. Stable Freestanding Li-ion Battery Cathodes by in Situ Conformal Coating of Conducting PolypyrroleonNiS-carbon Nanofiber Films. J. Power Sources 2016, 331, 360–365. [Google Scholar] [CrossRef]
  28. Salavati-niasari, M.; Davar, F.; Emadi, H. Hierarchical Nanostructured Nickel sulfide architectures through Simple Hydrothermal Method in the Presence of Thioglycoloic Acid. Chalco. Lett. 2010, 7, 647–655. [Google Scholar]
  29. Al-Naggar, H.; Shinde, N.M.; Kim, J.; Mane, R.S. Water splitting performance of metal and non-metal-doped transition metal oxide electrocatalysts. Coord. Chem. Rev. 2022, 474, 21486–21644. [Google Scholar]
  30. Feng, N.; Hu, D.; Wang, P.; Sun, X.; Li, X.; He, D. Growth of Nanostructured Nickel Sulfide Films on Ni foam as High-performance Cathodes for Lithium ion Batteries. Phys. Chem. Chem. Phys. 2013, 15, 9924–9930. [Google Scholar] [CrossRef]
  31. Liang, K.; Marcus, K.; Guo, L.; Li, Z.; Zhou, L.; Li, Y.; De Oliveira, S.; Orlovskaya, N.; Sohn, Y.; Yang, Y. A freestanding NiSx Porous Film as a Binder-free Electrode for Mg-ion Batteries. Chem. Commun. 2017, 53, 7608–7611. [Google Scholar] [CrossRef] [PubMed]
  32. Shinde, N.M.; Shinde, P.V.; Yun, J.M.; Mane, R.S.; Kim, K.H. Room-temperature chemical synthesis of 3-D dandelion-type nickel chloride (NiCl2@ NiF) supercapattery nanostructured materials. J. Colloid Interface Sci. 2020, 578, 547–554. [Google Scholar] [CrossRef] [PubMed]
  33. Yan, S.; Shi, Y.; Sun, L.; Xiao, Z.; Sun, B.; Xu, X. Controlled Synthesis of NiSNanoparticle/CdS Nanowire Heterostructures via Solution Route and Their Optical Properties. Mater. Sci. Eng. B 2013, 178, 109–116. [Google Scholar] [CrossRef]
  34. Cheng, Z.; Abernathy, H.; Liu, M. Raman Spectroscopy of Nickel SulfideNi3S2. J. Phys. Chem. C 2007, 111, 17997–18000. [Google Scholar] [CrossRef]
  35. Akbarzadeh, R.; Dehghani, H.; Behnoudnia, F. Sodium Thiosulfate-assisted Synthesis of NiS2 Nanostructure by Using Nickel (II)-Salen Precursor: Optical and Magnetic Properties. Dalton Trans. 2014, 43, 16745–16753. [Google Scholar] [CrossRef]
  36. Dong, J.; Cheng, Z.; Zha, S.; Liu, M. Identification of Nickel Sulfides on Ni-YSZ Cermet Exposed to H2 Fuel Containing H2S Using Raman spectroscopy. J. Power Sources 2006, 156, 461–465. [Google Scholar] [CrossRef]
  37. Li, H.; Chai, L.; Wang, X.; Wu, X.; Xi, G.; Liu, Y.; Qian, Y. Hydrothermal Growth and Morphology Modification of β-NiS Three-Dimensional Flowerlike Architectures. Cryst. Growth Des. 2007, 7, 1919–1920. [Google Scholar] [CrossRef]
  38. Lin, T.; Shuan, C.; Hung, K. High Energy Density Asymmetric supercapacitor Based on NiOOH/Ni3S2/3D Graphene and Fe3O4/Graphene Composite Electrodes. Sci. Rep. 2014, 4, 7274. [Google Scholar] [CrossRef] [Green Version]
  39. Jing, F.; Lv, Q.; Xiao, J.; Wang, Q.; Wang, S. Highly Active and Dual-function Self-Supported Multiphase NiS–NiS2–Ni3S2/NF Electrodes for Overall Water Splitting. J. Mater. Chem. A 2018, 6, 14207–14214. [Google Scholar] [CrossRef]
  40. Kang, D.; Hu, C.; Zhu, Q. Morphology controlled synthesis of hierarchical structured Fe2O3 from natural ilmenite and its high performance for dyes adsorption. App. Sur. Sci. 2018, 459, 327–335. [Google Scholar] [CrossRef]
  41. Deng, J.; Ren, P.; Deng, D.; Yu, L.; Yang, F.; Bao, X. Highly active and durable non-precious-metal catalysts encapsulated in carbon nanotubes for hydrogen evolution reaction. Energy Environ. Sci. 2014, 7, 1919–1923. [Google Scholar] [CrossRef]
  42. Suryawanshi, U.P.; Ghorpade, U.V.; Lee, D.M.; He, M.; Shin, S.W.; Kumar, P.V.; Jang, J.S.; Jung, H.R.; Suryawanshi, M.P.; Kim, J.H. Colloidal Ni2P Nanocrystals Encapsulated in Heteroatom-Doped Graphene Nanosheets: A Synergy of 0D@ 2D Heterostructure Toward Overall Water Splitting. Chem. Mater. 2020, 33, 234–245. [Google Scholar] [CrossRef]
  43. Babar, P.; Lokhande, A.; Karade, V.; Pawar, B.; Gang, M.G.; Pawar, S.; Kim, J.H. Bifunctional 2D electrocatalysts of transition metal hydroxide nanosheet arrays for water splitting and urea electrolysis. ACS Sustain. Chem. Eng. 2019, 7, 10035–10043. [Google Scholar] [CrossRef]
  44. Xiao, X.; Huang, D.; Fu, Y.; Wen, M.; Jiang, X.; Lv, X.; Li, M.; Gao, L.; Liu, S.; Wang, M.; et al. Engineering NiS/Ni2P heterostructures for efficient electrocatalytic water splitting. ACS Appl. Mater. Interfaces 2018, 10, 4689–4696. [Google Scholar] [CrossRef]
  45. Li, Q.; Wang, D.; Han, C.; Ma, X.; Lu, Q.; Xing, Z.; Yang, X. Construction of amorphous interface in an interwoven NiS/NiS2 structure for enhanced overall water splitting. J. Mater. Chem. A 2018, 6, 8233–8237. [Google Scholar] [CrossRef]
  46. Wang, T.; Guo, X.; Zhang, J.; Xiao, W.; Xi, P.; Peng, S.; Gao, D. Electronic structure modulation of NiS2 by transition metal doping for accelerating the hydrogen evolution reaction. J. Mater. Chem. A 2019, 7, 4971–4976. [Google Scholar] [CrossRef]
  47. Tian, T.; Huang, L.; Ai, L.; Jiang, J. Surface anion-rich NiS2 hollow microspheres derived from metal–organic frameworks as a robust electrocatalyst for the hydrogen evolution reaction. J. Mater. Chem. A 2017, 5, 20985–20992. [Google Scholar] [CrossRef]
  48. Wu, X.; Yang, B.; Li, Z.; Lei, L.; Zhang, X. Synthesis of supported vertical NiS2 nanosheets for hydrogen evolution reaction in acidic and alkaline solution. RSC Adv. 2015, 5, 32976–32982. [Google Scholar] [CrossRef]
  49. Fan, M.; Li, X.; Wei, D.; Wang, Y.; Li, M. Fabrication of Te@ NiTe2/NiS heterostructures for electrocatalytic hydrogen evolution reaction. Electrochim. Acta 2019, 328, 135075. [Google Scholar]
  50. Chia, X.; Sofer, Z.; Luxa, J.; Pumera, M. Unconventionally layered CoTe2 and NiTe2 as electrocatalysts for hydrogen evolution. Chem. A Eur. J. 2017, 23, 11719–11726. [Google Scholar] [CrossRef]
  51. De Silva, U.; Masud, J.; Zhang, N.; Hong, Y.; Liyanage, W.P.; Zaeem, M.A.; Nath, M. Nickel telluride as a bifunctional electrocatalyst for efficient water splitting in alkaline medium. J. Mater. Chem. A 2018, 6, 7608–7622. [Google Scholar] [CrossRef]
  52. Al-Enizi, A.M.; Shaikh, S.F.; Ubaidullah, M.; Ghanem, M.A.; Mane, R.S. Self-grown one-dimensional nickel sulfo-selenide nanostructured electrocatalysts for water splitting reactions. Int. J. Hydrog. Energy 2020, 45, 15904–15914. [Google Scholar] [CrossRef]
  53. Lu, J.; Xiong, T.; Zhou, W.; Yang, L.; Tang, Z.; Chen, S. Metal nickel foam as an efficient and stable electrode for hydrogen evolution reaction in acidic electrolyte under reasonable overpotentials. ACS Appl. Mater. Interfaces 2016, 8, 5065–5069. [Google Scholar] [CrossRef] [PubMed]
Figure 1. [(a,a-1,b,b-1)] False FE-SEM images of NiCl2 obtained with and without glycerol at different magnifications, and (c) TEM image; (c-1) HR-TEM (inset shows SEAD pattern) of NiCl2 electrode prepared with glycerol.
Figure 1. [(a,a-1,b,b-1)] False FE-SEM images of NiCl2 obtained with and without glycerol at different magnifications, and (c) TEM image; (c-1) HR-TEM (inset shows SEAD pattern) of NiCl2 electrode prepared with glycerol.
Catalysts 13 00177 g001
Scheme 1. Self-growth of NiCl2@NiF nanostructures with and without glycerol surfactant.
Scheme 1. Self-growth of NiCl2@NiF nanostructures with and without glycerol surfactant.
Catalysts 13 00177 sch001
Figure 2. (a) XRD patterns (A and B) electrode with and without glycerol NiCl2 and C (i.e., NiF), (b) full XPS spectrum survey, (c) Ni 2p, (d) Cl, (e) BET and (f) pore size distribution of electrode with glycerol, respectively.
Figure 2. (a) XRD patterns (A and B) electrode with and without glycerol NiCl2 and C (i.e., NiF), (b) full XPS spectrum survey, (c) Ni 2p, (d) Cl, (e) BET and (f) pore size distribution of electrode with glycerol, respectively.
Catalysts 13 00177 g002
Figure 3. (a) HER polarization curves, (b) Tafel plots, and (c) EIS spectroscopy measurements of NiF and NiCl2 (with and without glycerol) electrode materials. (d) Cyclic stability test (inset shows the real-time image of H2 evolution reaction) of dandelion flower-type NiCl2.
Figure 3. (a) HER polarization curves, (b) Tafel plots, and (c) EIS spectroscopy measurements of NiF and NiCl2 (with and without glycerol) electrode materials. (d) Cyclic stability test (inset shows the real-time image of H2 evolution reaction) of dandelion flower-type NiCl2.
Catalysts 13 00177 g003
Scheme 2. Schematic presentation showing various causes responsible for the enhanced catalytic performance of dandelion flower-type NiCl2 catalyst towards the HER.
Scheme 2. Schematic presentation showing various causes responsible for the enhanced catalytic performance of dandelion flower-type NiCl2 catalyst towards the HER.
Catalysts 13 00177 sch002
Table 1. The comparative set up showing HER activities of present work with reported data.
Table 1. The comparative set up showing HER activities of present work with reported data.
CatalystsElectrolyteɳ (mV)Tafel Slope (mV dec−1)J
(mA cm−2)
Ref.
Ni2P@NSG1.0 M KOH1104310[42]
NiFeCo1.0 M KOH1086410[43]
NiS/Ni2P/carbon cloth1.0 M KOH11178.120[44]
NiS/NiS21.0 M KOH248142.310[45]
Fe- NiS20.5 M H2SO41984210[46]
NiS21.0 M KOH21915710[47]
NiS2/graphite substrate0.5 M H2SO42404110[48]
Te@NiTe2/NiS/acetylene black1.0 M KOH10111810[49]
NiTe20.5 M H2SO45604410[50]
Ni3Te21.0 M KOH30494.210[51]
NiSSe1.0 M KOH15412510[52]
NiF0.5 M H2SO42101006.610[53]
NiF20.1 M KOH1724710[12]
NiCl21.0 M KOH1764110Current 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

Shinde, N.M.; Raut, S.D.; Ghule, B.G.; Deokate, R.J.; Narwade, S.H.; Mane, R.S.; Xia, Q.; Pak, J.J.; Kim, J.-S. Hydrogen Evolution Reaction Activities of Room-Temperature Self-Grown Glycerol-Assisted Nickel Chloride Nanostructures. Catalysts 2023, 13, 177. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13010177

AMA Style

Shinde NM, Raut SD, Ghule BG, Deokate RJ, Narwade SH, Mane RS, Xia Q, Pak JJ, Kim J-S. Hydrogen Evolution Reaction Activities of Room-Temperature Self-Grown Glycerol-Assisted Nickel Chloride Nanostructures. Catalysts. 2023; 13(1):177. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13010177

Chicago/Turabian Style

Shinde, Nanasaheb M., Siddheshwar D. Raut, Balaji G. Ghule, Ramesh J. Deokate, Sandesh H. Narwade, Rajaram S. Mane, Qixun Xia, James J. Pak, and Jeom-Soo Kim. 2023. "Hydrogen Evolution Reaction Activities of Room-Temperature Self-Grown Glycerol-Assisted Nickel Chloride Nanostructures" Catalysts 13, no. 1: 177. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13010177

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