Next Article in Journal
Validation of a Fixed Bed Reactor Model for Dimethyl Ether Synthesis Using Pilot-Scale Plant Data
Next Article in Special Issue
Investigation of Photocatalysis by Mesoporous Titanium Dioxide Supported on Glass Fibers as an Integrated Technology for Water Remediation
Previous Article in Journal
Application of Metal-Based Nanocatalysts for Addressing Environmental Issues and Energy Demand
Previous Article in Special Issue
Ceramized Fabrics and Their Integration in a Semi-Pilot Plant for the Photodegradation of Water Pollutants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phytochemical-Assisted Green Synthesis of Nickel Oxide Nanoparticles for Application as Electrocatalysts in Oxygen Evolution Reaction

by
Vidhya Selvanathan
1,
M. Shahinuzzaman
2,*,
Shankary Selvanathan
3,
Dilip Kumar Sarkar
1,
Norah Algethami
4,
Hend I. Alkhammash
5,
Farah Hannan Anuar
6,
Zalita Zainuddin
7,
Mohammod Aminuzzaman
8,9,
Huda Abdullah
10 and
Md. Akhtaruzzaman
1,*
1
Solar Energy Research Institute (SERI), Universiti Kebangsaan Malaysia (UKM), Bangi 43600, Selangor, Malaysia
2
School of Computer Science and Information Technology, Central University of Science and Technology, Dhaka 1216, Bangladesh
3
Department of Chemistry, Faculty of Science, University of Malaya, Kuala Lumpur 50603, Malaysia
4
Department of Physics, Faculty of Science, Taif University, Taif 21944, Saudi Arabia
5
Department of Electrical Engineering, College of Engineering, Taif University, Taif 21944, Saudi Arabia
6
Department of Chemistry, Faculty of Science and Technology, Universiti Kebangsaan Malaysia (UKM), Bangi 43600, Selangor, Malaysia
7
Department of Physics, Faculty of Science and Technology, Universiti Kebangsaan Malaysia (UKM), Bangi 43600, Selangor, Malaysia
8
Department of Chemical Science, Faculty of Science, Universiti Tunku Abdul Rahman (UTAR), Perak Campus, Jalan Universiti, Bandar Barat, Kampar 31900, Perak, Malaysia
9
Centre for Photonics and Advanced Materials Research (CPAMR), Universiti Tunku Abdul Rahman (UTAR), Jalan Sungai Long, Bandar Sungai Long, Kajang 43000, Selangor, Malaysia
10
Department of Electrical, Electronic & Systems Engineering, Faculty of Engineering & Built Environment, Universiti Kebangsaan Malaysia (UKM), Bangi 43600, Selangor, Malaysia
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(12), 1523; https://doi.org/10.3390/catal11121523
Submission received: 20 November 2021 / Revised: 7 December 2021 / Accepted: 11 December 2021 / Published: 15 December 2021
(This article belongs to the Special Issue Heterogeneous Photocatalysis: A Solution for a Greener Earth)

Abstract

:
Electrocatalytic water splitting is a promising solution to resolve the global energy crisis. Tuning the morphology and particle size is a crucial aspect in designing a highly efficient nanomaterials-based electrocatalyst for water splitting. Herein, green synthesis of nickel oxide nanoparticles using phytochemicals from three different sources was employed to synthesize nickel oxide nanoparticles (NiOx NPs). Nickel (II) acetate tetrahydrate was reacted in presence of aloe vera leaves extract, papaya peel extract and dragon fruit peel extract, respectively, and the physicochemical properties of the biosynthesized NPs were compared to sodium hydroxide (NaOH)-mediated NiOx. Based on the average particle size calculation from Scherrer’s equation, using X-ray diffractograms and field-emission scanning electron microscope analysis revealed that all three biosynthesized NiOx NPs have smaller particle size than that synthesized using the base. Aloe-vera-mediated NiOx NPs exhibited the best electrocatalytic performance with an overpotential of 413 mV at 10 mA cm−2 and a Tafel slope of 95 mV dec−1. Electrochemical surface area (ECSA) measurement and electrochemical impedance spectroscopic analysis verified that the high surface area, efficient charge-transfer kinetics and higher conductivity of aloe-vera-mediated NiOx NPs contribute to its low overpotential values.

1. Introduction

One of the greatest challenges of the technological era is the inflating global energy demand as electricity becomes a necessity in every facet of life. The non-renewable nature of fossil-fuel-based energy production has stimulated the pursuit of sustainable energy sources such as solar and wind energy. To overcome the intermittency issues of these renewable energy sources, production of hydrogen fuel from electrocatalytic water splitting is considered a strategic approach. Under a standard environment, 1.23 V of thermodynamic potential is required to drive electrochemical water splitting [1]. Nonetheless, in practical conditions, the sluggish kinetics of oxygen evolution reaction (OER) mandates high overpotential to generate the required current density. Hence, most practical electrolyzers consume more than 1.23 V to drive the desired reaction.
Electrocatalysts are usually employed to attain satisfactory water-splitting efficiency, but by far best electrocatalytic performance was demonstrated by noble-metal-based materials (e.g., RuO2 and IrO2). The high cost and low abundancy of noble metals are not conducive for large-scale water-splitting applications. Therefore, non-precious transition metal oxides and hydroxides were explored as alternative electrocatalysts. Nickel oxide (NiOx) is one such promising material for electrocatalytic water splitting due to its high electron transport capacity, good chemical stability and facile method of synthesis. Particularly, in the case of NiOx NPs, the synthetic route adapted influences the electrical and structural properties of the material.
Traditionally, metal oxide nanoparticles are synthesized using a strong base as a reducing agent, followed by reaction with a capping agent or reducing agent. Sometimes, additional organic solvents are required to dissolve these capping agents. Alternatively, a green synthesis approach can be utilized to reduce the usage of unnecessary toxic chemicals [2]. In green synthesis, primary and secondary metabolites of plants are used to aid the synthesis of metal oxide nanoparticles [3]. These metabolites form a coating layer or stabilizing layer on the surface of the metal oxide nanoparticles and prevent intensive aggregation, thus yielding well-defined morphologies and tunable particle size [4]. The choice of plant species and its respective parts play an important role in determining the morphology and particle size of the NPs as they contain distinct phytochemical profiles [5].
One of the major advantages of chemical capping agents is the tunability of the species to yield nanoparticles with different particle size and morphology. For example, in the case of synthetic-polymers-based capping agents, factors such as the nature of polymer (functional group, branching), molecular weight and concentration were found to affect its function as a stabilizer [6]. However, in the context of green synthesis, much work is dedicated to exploring the effect of using a single plant extract formulation. Since each plant species comes with its unique cocktail of phytochemicals, it is useful to compare the effect of using different plant extracts on the overall size and morphology of the nanoparticles. Thus, in this study we have explored three different plant extracts, namely, aloe vera leaves extract, papaya fruit peel extract and dragon fruit peel extract, as capping agents for NiOx NPs.
Aloe vera gel contains mostly water and polysaccharides, such as pectins, cellulose, hemicellulose, glucomannan and acemannan [7,8]. In addition to that, the aloe latex is made up of hydroxyanthracenic derivatives and anthraquinone glycosides, along with emodin. Papaya peel is rich in a variety of vitamins including β-carotene, vitamin B (thiamine, riboflavin, niacin and folate), vitamin C and vitamin E [9]. On the other hand, dragon fruit peel is reported to contain wide compositions of antioxidants such as betasianin, flavonoids, phenols, terpenoids, thiamine, niacin, pyridoxine and cobalamin [10]. The effect of these different phytochemical compositions of extracts on the structural and morphological properties of the NiOx NPs and, consequently, their electrocatalytic activities for OER were studied and reported in detail.

2. Results and Discussion

2.1. Structural Properties

Figure 1 shows the structural properties of the synthesized NiOx nanoparticles using different extracts after calcination at 300 °C and 600 °C. As shown in the diffractograms, all the samples exhibited diffraction peaks at 2θ values of 37.2, 43.2, 62.8, 75.4 and 79.4° attributed to (101), (012), (104), (113) and (006) crystal planes of cubic NiOx [11,12]. These peaks were in agreement with the standard JCPDS card no. 00-004-1159. Significant differences in the peak intensities were evident between samples calcined at 300 °C and 600 °C. At lower calcination temperature, the presence of organic residues from the plant extract used during NiOx synthesis reduced the crystallinity of the nanoparticles. Upon calcination at 600 °C, high purity NiOx nanoparticles were produced, leading to sharp and intense XRD peaks.
The average crystal sizes (D) were calculated based on the width of the peak due to the (200) planes by using the Scherrer’s formula [13]:
D = 0.94 λ β C o s θ
From Table 1, it could be seen that the average crystallite size of the nanoparticles synthesized with only NaOH was 15.9 nm, whereas aloe vera, papaya and dragon fruit extracts was 11.5 nm, 8.1 nm and 10.2 nm, respectively. The reduced crystallite size of NiOx nanoparticles synthesized in presence of plant extracts emphasizes the role of phytochemicals that are able to chelate with metal ions, thus acting as stabilizers or capping agents. Figure 2 illustrates the plausible mechanism for formation of NiOx nanoparticles in this study. Upon dissolution, nickel acetate tetrahydrate dissociated to release Ni2+ ions which then combined with hydroxide ions from the base to form Ni(OH)2. The continuous stirring under a heated environment enabled Ni(OH)2 to thermally decompose into NiO nuclei [14]. The phytochemicals present in each plant extract served as capping ligands which formed interactions with the nanoparticles’ surface, leading to steric hindrance and providing eventual stability to the nanocomposite. Hence, the presence of biomass-derived extracts controlled the particle size and minimized agglomeration of the nanoparticles.

2.2. Morphological Properties

The morphological properties of the synthesized NiOx nanoparticles are shown in Figure 3. From the images, it is clear that the NiOx nanoparticles were spherical-shaped, densely packed and intercalated with each other. Dragon fruit and papaya-extract-mediated NiOx NPs depict well-defined nanostructures with fewer intercalations. The particle size distribution for the NiOx NPs synthesized without extract ranged between 40 and 60 nm, whereas plant-extract-mediated NiOx NPs were between 20 and 30 nm in size. The trend in particle size coincided with the trend in crystallite size derived from the XRD analysis. For all the samples, the particle size revealed in morphological analysis was higher than average crystallite size estimated from XRD analysis, hence indicating that the single NiOx particle is made up of few crystallites.
The elemental composition of the nanoparticles was verified from EDX analysis. As shown in Figure 4, the choice of plant extract influenced the composition of nickel and oxygen in the samples. NiOx NPs synthesized using aloe vera were more oxygen deficient compared to papaya and dragon fruit. Aloe-vera-mediated nickel oxide also exhibited the highest amount of residual carbon, despite calcining the samples at 600 °C. It is deduced that this could be due to the more viscous nature of the aloe vera extract, which adheres to the surface of the nanoparticles during synthesis. The absence of residual carbon originating from the phytochemicals was evident in dragon-fruit- and papaya-extract-mediated NiOx NPs, which could explain the occurrence of fewer intercalations, as shown in the FESEM images.

2.3. Optical Properties

The absorption spectrum of the NiOx nanoparticles, as shown in Figure 5a, was recorded at room temperature within the range of 200–800 nm. There was a peak observed at around 300–350 nm for all synthesized nanopowders, which can be assigned to the exciton transitions [15]. The band gap energies were estimated by extrapolation using Tauc plots, and the energy curves are shown in Figure 5. It was seen that the band gap of NiOx nanoparticle synthesis in different samples were ranging from 3.09 to 3.22 eV and the band gap was slightly increased for plant-extract-assisted synthesized nanoparticles. Particularly, the decrease in particle size, as well as the increase in band gap energy of the as-prepared nanoparticles, signified the size quantization effects [16]. When particle size reached the nanosized scale, the overlapping of adjacent energy levels minimized and, subsequently, the width of energy band widened [17,18].

2.4. Electrocatalytic Properties for Oxygen Evolution Reaction (OER)

The electrocatalytic activity of the synthesized NiOx NPs towards OER was explored by performing linear scanning voltammetry (LSV) in an aqueous solution of 1.0 M KOH. Generally, the oxygen evolution reaction in alkaline medium is represented as follows [19,20,21]:
4 O H O 2 + 2 H 2 O + 4 e
The four-electron reaction mechanism can be described sequentially via the following equations:
O H O H * + e
O H * + O H O * + H 2 O + e
O * + O H O O H * + e
O O H * + O H O 2 + 2 H 2 O + e
where OH*, O* and OOH* indicate the species adsorbed to the active site of the catalyst.
Based on the polarization curves shown in Figure 6a, to attain the current density of 10 mA cm2, aloe-vera- and papaya-extract-mediated NiOx NPs exhibited the overpotential values of 416 mV and 433 mV, respectively, which are lower than NaOH-based NiOx, which records 474 mV. Interestingly, dragon-fruit-extract-mediated NiOx showed the overpotential of 501 mV, the highest overpotential compared to the previous three samples. As listed in Table 2, the onset potential, which indicates the voltage at which current begins to rise, was lowest for aloe-vera-mediated NPs and highest for dragon fruit.
The plausible reaction mechanism of NiOx for OER is outlined as follows:
N i 2 + + 3 O H N i O O H + H 2 O + e
N i O O H + O H N i O ( O H ) 2 + e
N i O ( O H ) 2 + 2 O H N i O O 2 + 2 H 2 O + 2 e
N i O O 2 + 2 O H N i O O H + O 2 + e
Throughout the electrocatalytic process, Ni2+ ions that can be present on surface of the NiOx NPs will undergo oxidation to form Ni3+ ions which then aid to catalyze the oxidation of OH ions, leading to the formation of water [22].
To further comprehend the reaction kinetics of the nickel-oxide-based electrocatalysts, the Tafel plot was plotted (Figure 6c), and the linear section of Tafel slope was derived from the following equation:
η = b log j + a
where η is the overpotential, b is the Tafel slope and j refers to current density. NaOH-, aloe-vera-, papaya- and dragon-fruit-mediated NiOx recorded slope values of 191, 95, 122 and 224 mV/decade, respectively. The Tafel slope values indicate the rate-limiting step which is attributed to the reaction with first electron transfer. Thus, the lower Tafel slope value of aloe-vera-mediated NiOx NPs translates to more efficient reaction kinetics for OER. One of the proposed explanations for the superior electrocatalytic performance of aloe-vera-mediated NiOx NPs can be related to the relatively high oxygen deficiency in the samples, as evidenced by EDX measurements. The non-stoichiometric composition most likely led to the presence of surface defects, which have been found to enhance the electrocatalytic performance of metal oxides for OER [23,24].
Another key factor that governs the performance of electrocatalyst is the surface area available as active sites for catalytic activities. In OER, electrochemically active surface area (ECSA) is used as the parameter to assess the specific catalytic activity of an electrocatalyst, and it can be derived from electrochemical double-layer capacitance (Cdl). To calculate Cdl values, the samples were subjected to cyclic voltammetry (CV) at different scan rates in the non-Faradaic segment where the voltage was scanned from 1.46 to 1.56 V vs. RHE (Figure 7). The double-layer capacitance was then derived based on the gradient of linear plot of ΔJ = (Ja − Jc) (where Ja and Jc are anodic and cathodic current density) at 1.52 V vs. RHE as a function of the scan rate [25]. The ECSA was then calculated from the double-layer capacitance based on the following equation:
E C S A = C d l C s
where Cs is the specific capacitance which is equated to 0.040 mF cm−2 based on the conventional value for metal electrode in KOH solution. The roughness factor (Rf) can then be evaluated by dividing ECSA by the electrode area. Table 3 summarizes the ECSA and roughness factor for all the samples. Papaya-mediated NiOx NPs exhibit the highest value of ECSA and Rf of 5.50 mFcm2 and 1936, respectively, thus indicating that it has the highest catalytically active surface area. This could be the boosting factor that supports the low overpotential value recorded by papaya-mediated NiOx NPs despite its high Tafel slope. The variation of active surface area of the electrocatalysts prepared using different plant extracts further emphasizes the role of phytochemicals in fine-tuning the morphology and size distribution of nanoparticles during the synthesis process.
Electrochemical impedance spectroscopy (EIS) is a powerful technique that allows us to obtain insights on electrode–electrolyte interface dynamics. The EIS measurements for all the samples were performed with a frequency range from 0.1 Hz to 100 kHz, and the resulting Nyquist plots comprising of semicircles are presented in Figure 8. The starting point of the semicircle signifies the internal resistance (also known as solution resistance (RS)) of the catalyst, whereas the diameter of the semicircle denotes charge-transfer resistance between the electrode–electrolyte interfaces (RCT) [26]. In terms of equivalent circuit, the depressed semicircle is represented by RS connected in series to the capacitor (QDL) and RCT in a parallel arrangement, as shown in Figure 8. The following equation evaluates impedance values according to the described equivalent circuit [27]:
Z D S S C = R S + R C T 1 + ( j ω ) n C T   R C T   Q C T
The equation was used to fit the Nyquist plots and the solution resistance, charge-transfer resistance and double-layer capacitance based on the equivalent circuit fitting are tabulated in Table 4. Typically, charge-transfer resistance (RCT) is considered the most significant parameter as it dictates the extent of electronic transfers between the reactant and electrocatalyst surface. A lower RCT value also indicates higher conductivity of the electrocatalyst material which favors electron mobility during the electrocatalytic reaction [28]. In this study, aloe-vera-mediated NiOx NPs showed the lowest solution resistance and charge-transfer resistance. This can be attributed to better conductivity of the NiOx NPs due to its higher oxygen deficiency compared to other green-synthesized nanoparticles. Therefore, it is evident that efficient charge-transfer kinetics and improved conductivity is one of the crucial factors for the higher electrocatalytic performance of aloe-vera-mediated NiOx NPs compared to its dragon fruit and papaya counterparts.
To evaluate the stability of aloe-vera-mediated NiOx NPs, chronoamperometry analysis was performed for 24 h at an overpotential of 415 mV. Within the first 2 h, it was observed that there was a gradual increase in current density. This is attributed to the activation of nickel oxide into nickel oxyhydroxide (NiOOH) during the initial part of the analysis [22]. Higher conductivity of NiOOH compared to NiO rendered the gradual increase in current density. As the chronoamperometry was extended for 24 h, the current density remained stable without any significant decay.
The stability of the final product was further verified by subjecting the NiOx-modified glassy carbon electrode to cyclic voltammetry (CV) analysis for 500 continuous cycles, which were carried out between 1.2 and 1.8 V at a scan rate of 100 mVs−1. As shown in Figure 9, after 500 CV cycles, the overpotential of aloe-vera-mediated NiOx NPs increased from 416 mV to 428 mV with a difference of only 0.12 mV at the current density of 10 mA cm−2. This observation validates the high electrocatalytic stability of the green-synthesized nickel oxide nanoparticle for OER activity.

3. Materials and Methods

3.1. Chemicals

Nickel acetate tetrahydrate (Ni(CH3COO)2.4H2O), NaOH and KOH were purchased from Sigma Aldrich, Hamburg, Germany. Deionized (DI) water was used to prepare all types of solution. All the other chemicals were of analytical reagent grade and used without any further purification.

3.2. Preparation of Plant Extract

The NiOx nanoparticles were synthesized using various plant extracts such as aloe vera leaf extract, dragon fruit peel extract and papaya peel extract. NiOx NPs without extract were also synthesized. Different plant samples were collected from the local market at Hentian Kajang, Selangor, Malaysia. In order to prepare the plant extract, at first, 50 g of different plant samples were washed with DI water, cut into small pieces and were taken in three different 500 mL glass beakers and 250 mL DI water was added to all the beakers. Then the mixtures were heated at 60–70 °C for 30 min with continuous stirring. After heating, the extracts were filtered with vacuum filter and the filtrates were collected in three different 500 mL reagent bottles.

3.3. Synthesis of NiOx Nanoparticle

The NiOx nanoparticles were synthesized using the green synthesis technique. In this technique, at first, 30 mL (1 M) nickel acetate solutions were heated in four different 250 mL glass beakers at 60–70 °C temperature with continuous stirring. Then, 30 mL DI water was added to the first beaker (without extract) and 30 mL of three different plant extracts to the other three beakers was added in the precursor solution with constant heating (60–70 °C temperature) and stirring. Then, 1 M NaOH solution was added, dropwise, into the reaction mixture with vigorous stirring and heating for 2 h. After completion of the reaction, the mixtures were washed with DI water several times and filtered using Whatman 1 filter paper. Finally, the synthesized NPs were dried in an oven at 100 °C overnight, calcined at 300 °C and 600 °C for 6 h, powdered using mortar-pastel and stored in the glass vial for further analysis. Figure 10 represents the schematic view of green synthesis process of NiOx NPs.

3.4. Characterization of Materials

The morphological and structural properties of synthesized NiOx nanoparticles were examined using various characterization techniques such as X-ray diffraction (XRD), absorption spectra (UV-Vis) and field-emission scanning electron microscopy (FESEM), along with energy dispersive X-ray (EDX). The XRD patterns were taken in the 2θ ranging from 10° to 80° using BRUKER aXS-D8 Advance Cu-Ka diffractometer (Bruker, MA, USA). The FESEM model LEO 1450 Vp was used for investigating the surface morphology, as well as grain size and growth of the NiOx nanoparticles. The optoelectronic properties of the synthesized nanocatalyst, such as the optical transmittance, absorbance and optical band gap, were calculated using UV-Vis spectrometer Perkin Elmer Instruments Lambda35 (PerkinElmer, Waltham, MA, USA).

3.5. Electrochemical Characterization

Electrochemical characterization was performed using Metrohm Autolab workstation (Metrohm, Herisau, Switzerland) in a three-electrode setup in which platinum electrode and silver/silver chloride were used as the counter electrode and reference electrode, respectively. The NiOx NPs were drop casted onto glassy carbon electrode with an active area of 0.07 cm2 to function as the working electrode. A 5 mg amount of the NiOx NPs was mixed with 250 μL ethanol, 250 μL deionized water and 50 μL of Nafion (5 wt%). The dispersion was sonicated for an hour and 10 μL of the dispersion was drop casted on top of the glassy carbon electrode surface followed by drying. Linear sweep voltammetry between 0 and 1 V with scan rate of 5 mV s−1 was performed to evaluate the oxygen evolution reaction. The measured potential vs. Ag/AgCl was converted to a reversible hydrogen electrode (RHE) based on the following equation:
E R H E = E A g / A g C l + 0.059   p H + E A g / A g C l 0
where E A g / A g C l 0 is 0.1976 V at 25 °C and E A g / A g C l is measured potential vs. Ag/AgCl. The overpotential is calculated by measuring the difference between measured potential vs. RHE at current density of 10 mA cm−2 and the standard value of 1.23 V.

4. Conclusions

Utilization of different plant extracts as capping agents affects the structural and morphological properties of NiOx NPs. FESEM results show that NiOx NPs synthesized without extract had a particle size between 40 and 60 nm, whereas plant-extract-mediated NiOx were between 20 and 30 nm. The reduction in particle size upon introducing plant extracts was further supported by the shifting of the band gap to a higher value based on optical characterization. EDX analysis reveal that aloe vera, papaya and dragon fruit extract yielded NiOx NPs with different oxygen and nitrogen stoichiometry. The deficiency of oxygen is higher in aloe vera-mediated NiOx NPs, which might elevate its conductivity and thus lead to lower charge-transfer resistance in EIS analysis. ECSA calculation based on cyclic voltammetry indicated that the highest electrocatalytic surface area was possessed by papaya-extract-mediated NiOx NPs, followed by aloe vera and dragon fruit extract. Overall, the best electrocatalytic performance is attributed to aloe vera-extract-mediated NiOx NPs, which only required 433 mV overpotential to reach 10 mAcm−2 with a Tafel slope of 95 mVdec−1.

Author Contributions

Conceptualization, V.S, M.S. and S.S.; methodology, M.S., D.K.S. and M.A. (Mohammod Aminuzzaman); formal analysis, F.H.A. and Z.Z.; writing—original draft preparation, V.S. and S.S.; writing—review and editing, H.I.A. and N.A.; supervision, M.A. (Md. Akhtaruzzaman); project administration, H.A. and M.A. (Md. Akhtaruzzaman); funding acquisition, H.I.A. and N.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are highly grateful to Universiti Kebangsaan Malaysia for supporting this study through the Modal Insan fellowship (RFA1). Authors also extend their appreciation to The University Researchers Supporting Project Number (TURSP-2020/264), Taif University, Taif, Saudi Arabia. Authors are thankful to Universiti Tunku Abdul Rahman (UTAR) providing necessary research facilities to carry out this research work successfully.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are highly grateful to Universiti Kebangsaan Malaysia for sup-porting this study through the Modal Insan fellowship (RFA1). Authors also extend their appre-ciation to The University Researchers Supporting Project Number (TURSP-2020/264), Taif Uni-versity, Taif, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. You, B.; Sun, Y. Innovative strategies for electrocatalytic water splitting. Acc. Chem. Res. 2018, 51, 1571–1580. [Google Scholar] [CrossRef]
  2. Pal, G.; Rai, P.; Pandey, A. Chapter 1—Green synthesis of nanoparticles: A greener approach for a cleaner future. In Green Synthesis, Characterization and Applications of Nanoparticles; Shukla, A.K., Iravani, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–26. [Google Scholar] [CrossRef]
  3. Jadoun, S.; Arif, R.; Jangid, N.K.; Meena, R.K. Green synthesis of nanoparticles using plant extracts: A review. Environ. Chem. Lett. 2021, 19, 355–374. [Google Scholar] [CrossRef]
  4. Shafey, A.M.E. Green synthesis of metal and metal oxide nanoparticles from plant leaf extracts and their applications: A review. Green Process. Synth. 2020, 9, 304–339. [Google Scholar] [CrossRef]
  5. Singh, J.; Dutta, T.; Kim, K.-H.; Rawat, M.; Samddar, P.; Kumar, P. ‘Green’ synthesis of metals and their oxide nanoparticles: Applications for environmental remediation. J. Nanobiotechnol. 2018, 16, 84. [Google Scholar] [CrossRef]
  6. Madkour, M.; Bumajdad, A.; Al-Sagheer, F. To what extent do polymeric stabilizers affect nanoparticles characteristics? Adv. Colloid Interface Sci. 2019, 270, 38–53. [Google Scholar] [CrossRef]
  7. Tippayawat, P.; Phromviyo, N.; Boueroy, P.; Chompoosor, A. Green synthesis of silver nanoparticles in aloe vera plant extract prepared by a hydrothermal method and their synergistic antibacterial activity. PeerJ 2016, 4, e2589. [Google Scholar] [CrossRef]
  8. Quispe, C.; Villalobos, M.; Bórquez, J.; Simirgiotis, M. Chemical composition and antioxidant activity of Aloe vera from the Pica Oasis (Tarapacá, Chile) by UHPLC-Q/Orbitrap/MS/MS. J. Chem. 2018, 2018, 6123850. [Google Scholar] [CrossRef] [Green Version]
  9. Phang, Y.-K.; Aminuzzaman, M.; Akhtaruzzaman, M.; Muhammad, G.; Ogawa, S.; Watanabe, A.; Tey, L.-H. Green synthesis and characterization of CuO nanoparticles derived from papaya peel extract for the photocatalytic degradation of palm oil mill effluent (POME). Sustainability 2021, 13, 796. [Google Scholar] [CrossRef]
  10. Hendra, R.; Masdeatresa, L.; Abdulah, R.; Haryani, Y. Red dragon peel (Hylocereus polyrhizus) as antioxidant source. AIP Conf. Proc. 2020, 2243, 030007. [Google Scholar] [CrossRef]
  11. Manigandan, R.; Dhanasekaran, T.; Padmanaban, A.; Giribabu, K.; Suresh, R.; Narayanan, V. Bifunctional hexagonal Ni/NiO nanostructures: Influence of the core–shell phase on magnetism, electrochemical sensing of serotonin, and catalytic reduction of 4-nitrophenol. Nanoscale Adv. 2019, 1, 1531–1540. [Google Scholar] [CrossRef] [Green Version]
  12. Sekar, S.; Kim, D.Y.; Lee, S. Excellent oxygen evolution reaction of activated carbon-anchored NiO nanotablets prepared by green routes. Nanomaterials 2020, 10, 1382. [Google Scholar] [CrossRef]
  13. Munna, F.T.; Selvanathan, V.; Sobayel, K.; Muhammad, G.; Asim, N.; Amin, N.; Sopian, K.; Akhtaruzzaman, M. Diluted chemical bath deposition of CdZnS as prospective buffer layer in CIGS solar cell. Ceram. Int. 2021, 47, 11003–11009. [Google Scholar] [CrossRef]
  14. Tripathi, R.M.; Bhadwal, A.S.; Gupta, R.K.; Singh, P.; Shrivastav, A.; Shrivastav, B.R. ZnO nanoflowers: Novel biogenic synthesis and enhanced photocatalytic activity. J. Photochem. Photobiol. B Biol. 2014, 141, 288–295. [Google Scholar] [CrossRef]
  15. Lamba, P.; Singh, P.; Singh, P.; Kumar, A.; Singh, P.; Bharti; Kumar, Y.; Gupta, M. Bioinspired synthesis of nickel oxide nanoparticles as electrode material for supercapacitor applications. Ionics 2021, 27, 5263–5276. [Google Scholar] [CrossRef]
  16. Cao, Y.; Hu, P.; Jia, D. Solvothermal synthesis, growth mechanism, and photoluminescence property of sub-micrometer PbS anisotropic structures. Nanoscale Res. Lett. 2012, 7, 668. [Google Scholar] [CrossRef] [Green Version]
  17. Talluri, B.; Prasad, E.; Thomas, T. Ultra-small (r < 2 nm), stable (>1 year) copper oxide quantum dots with wide band gap. Superlattices Microstruct. 2018, 113, 600–607. [Google Scholar] [CrossRef] [Green Version]
  18. Selvanathan, V.; Aminuzzaman, M.; Tey, L.-H.; Razali, S.A.; Althubeiti, K.; Alkhammash, H.I.; Guha, S.K.; Ogawa, S.; Watanabe, A.; Shahiduzzaman, M.; et al. Muntingia calabura leaves mediated green synthesis of CuO nanorods: Exploiting phytochemicals for unique morphology. Materials 2021, 14, 6379. [Google Scholar] [CrossRef]
  19. Liang, Q.; Brocks, G.; Bieberle-Hütter, A. Oxygen evolution reaction (OER) mechanism under alkaline and acidic conditions. J. Phys. Energy 2021, 3, 026001. [Google Scholar] [CrossRef]
  20. Jin, H.; Wang, X.; Tang, C.; Vasileff, A.; Li, L.; Slattery, A.; Qiao, S.-Z. Stable and highly efficient hydrogen evolution from seawater enabled by an unsaturated nickel surface nitride. Adv. Mater. 2021, 33, 2007508. [Google Scholar] [CrossRef]
  21. Wang, C.; Qi, L. Hollow nanosheet arrays assembled by ultrafine ruthenium–cobalt phosphide nanocrystals for exceptional pH-universal hydrogen evolution. ACS Mater. Lett. 2021, 3, 1695–1701. [Google Scholar] [CrossRef]
  22. Rani, B.J.; Ravi, G.; Yuvakkumar, R.; Ravichandran, S.; Ameen, F.; Al-Sabri, A. Efficient, highly stable Zn-doped NiO nanocluster electrocatalysts for electrochemical water splitting applications. J. Sol-Gel Sci. Technol. 2019, 89, 500–510. [Google Scholar] [CrossRef]
  23. Zhu, K.; Shi, F.; Zhu, X.; Yang, W. The roles of oxygen vacancies in electrocatalytic oxygen evolution reaction. Nano Energy 2020, 73, 104761. [Google Scholar] [CrossRef]
  24. Wang, Y.; Liang, Z.; Zheng, H.; Cao, R. Recent progress on defect-rich transition metal oxides and their energy-related applications. Chemistry 2020, 15, 3717–3736. [Google Scholar] [CrossRef]
  25. Wang, H.; Xie, A.; Li, X.; Wang, Q.; Zhang, W.; Zhu, Z.; Wei, J.; Chen, D.; Peng, Y.; Luo, S. Three-dimensional petal-like graphene Co3.0Cu1.0 metal organic framework for oxygen evolution reaction. J. Alloys Comp. 2021, 884, 161144. [Google Scholar] [CrossRef]
  26. Santos, H.L.S.; Corradini, P.G.; Medina, M.; Dias, J.A.; Mascaro, L.H. NiMo–NiCu Inexpensive Composite with High Activity for Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2020, 12, 17492–17501. [Google Scholar] [CrossRef]
  27. Selvanathan, V.; Ruslan, M.H.; Aminuzzaman, M.; Muhammad, G.; Amin, N.; Sopian, K.; Akhtaruzzaman, M. Resorcinol-Formaldehyde (RF) as a Novel Plasticizer for Starch-Based Solid Biopolymer Electrolyte. Polymers 2020, 12, 2170. [Google Scholar] [CrossRef]
  28. Akbayrak, M.; Önal, A.M. Metal oxides supported cobalt nanoparticles: Active electrocatalysts for oxygen evolution reaction. Electrochim. Acta 2021, 393, 139053. [Google Scholar] [CrossRef]
Figure 1. X-ray diffractogram of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles calcined at (i) 300 °C and (ii) 600 °C.
Figure 1. X-ray diffractogram of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles calcined at (i) 300 °C and (ii) 600 °C.
Catalysts 11 01523 g001
Figure 2. Plausible mechanism for green synthesis of NiOx nanoparticles.
Figure 2. Plausible mechanism for green synthesis of NiOx nanoparticles.
Catalysts 11 01523 g002
Figure 3. FESEM images of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles.
Figure 3. FESEM images of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles.
Catalysts 11 01523 g003
Figure 4. EDX spectrum of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles.
Figure 4. EDX spectrum of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles.
Catalysts 11 01523 g004
Figure 5. (a) UV-Vis absorption spectrum and the corresponding Tauc plot of (b) NaOH-, (c) aloe-vera-, (d) dragon-fruit- and (e) papaya-extract-mediated NiOx nanoparticles.
Figure 5. (a) UV-Vis absorption spectrum and the corresponding Tauc plot of (b) NaOH-, (c) aloe-vera-, (d) dragon-fruit- and (e) papaya-extract-mediated NiOx nanoparticles.
Catalysts 11 01523 g005
Figure 6. (a) Polarization curve by linear sweep voltammetry, (b) overpotential at 10 mA cm−2 and (c) Tafel plot.
Figure 6. (a) Polarization curve by linear sweep voltammetry, (b) overpotential at 10 mA cm−2 and (c) Tafel plot.
Catalysts 11 01523 g006
Figure 7. Cyclic voltammogram of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles and (e) linear plot between current density and scan rate.
Figure 7. Cyclic voltammogram of (a) NaOH-, (b) aloe-vera-, (c) dragon-fruit- and (d) papaya-extract-mediated NiOx nanoparticles and (e) linear plot between current density and scan rate.
Catalysts 11 01523 g007
Figure 8. Nyquist plots from EIS analysis.
Figure 8. Nyquist plots from EIS analysis.
Catalysts 11 01523 g008
Figure 9. (a) Chronoamperometric stability tests at 1.65 V (vs the RHE) and (b) LSV curves obtained before and after 500 CV cycles of aloe-vera-mediated NiOx NPs.
Figure 9. (a) Chronoamperometric stability tests at 1.65 V (vs the RHE) and (b) LSV curves obtained before and after 500 CV cycles of aloe-vera-mediated NiOx NPs.
Catalysts 11 01523 g009
Figure 10. Reaction scheme for green synthesis process of NiOx NPs.
Figure 10. Reaction scheme for green synthesis process of NiOx NPs.
Catalysts 11 01523 g010
Table 1. Particle size calculation based on Debye–Scherrer’s equation.
Table 1. Particle size calculation based on Debye–Scherrer’s equation.
MaterialFWHMβcosθCrystallite
Size (nm)
Average
Crystallite
Size (nm)
NiO–NaOH37.40.510.008516.315.9
43.20.510.008316.8
62.80.630.009414.7
NiO–Aloe37.40.710.011711.811.5
43.20.630.010113.6
62.81.030.01539.0
NiO–Papaya37.40.900.01499.310.2
43.20.680.010912.6
62.81.080.01608.6
NiO–Dragon37.41.620.02695.28.1
43.20.680.010912.6
62.81.450.02166.4
Table 2. Electrocatalytic parameters of NiOx nanoparticles.
Table 2. Electrocatalytic parameters of NiOx nanoparticles.
MaterialOnset Potential
(V)
Overpotential
at 10 mA cm−2 (mV)
Tafel Slope
(mV dec−1)
NiO–NaOH 1.57474191
NiO–Aloe1.5441695
NiO–Papaya1.56433122
NiO–Dragon1.59501224
Table 3. Electrochemical active surface area and double-layer capacitance of NiOx nanoparticles.
Table 3. Electrochemical active surface area and double-layer capacitance of NiOx nanoparticles.
MaterialCdl (mF cm−2)ECSA (cm2)Rf
NiO–NaOH 4.50112.501584
NiO–Aloe4.80120.001690
NiO–Papaya5.50137.501936
NiO–Dragon2.2556.25792
Table 4. Equivalent circuit fitting parameters for the EIS analysis.
Table 4. Equivalent circuit fitting parameters for the EIS analysis.
MaterialRS (Ω)RCT (Ω)QDL (F)
NiO–NaOH 44.5339.93.04 × 10−4
NiO–Aloe17.0187.93.59 × 10−4
NiO–Papaya29.2265.62.72 × 10−4
NiO–Dragon29.4544.51.54 × 10−4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Selvanathan, V.; Shahinuzzaman, M.; Selvanathan, S.; Sarkar, D.K.; Algethami, N.; Alkhammash, H.I.; Anuar, F.H.; Zainuddin, Z.; Aminuzzaman, M.; Abdullah, H.; et al. Phytochemical-Assisted Green Synthesis of Nickel Oxide Nanoparticles for Application as Electrocatalysts in Oxygen Evolution Reaction. Catalysts 2021, 11, 1523. https://doi.org/10.3390/catal11121523

AMA Style

Selvanathan V, Shahinuzzaman M, Selvanathan S, Sarkar DK, Algethami N, Alkhammash HI, Anuar FH, Zainuddin Z, Aminuzzaman M, Abdullah H, et al. Phytochemical-Assisted Green Synthesis of Nickel Oxide Nanoparticles for Application as Electrocatalysts in Oxygen Evolution Reaction. Catalysts. 2021; 11(12):1523. https://doi.org/10.3390/catal11121523

Chicago/Turabian Style

Selvanathan, Vidhya, M. Shahinuzzaman, Shankary Selvanathan, Dilip Kumar Sarkar, Norah Algethami, Hend I. Alkhammash, Farah Hannan Anuar, Zalita Zainuddin, Mohammod Aminuzzaman, Huda Abdullah, and et al. 2021. "Phytochemical-Assisted Green Synthesis of Nickel Oxide Nanoparticles for Application as Electrocatalysts in Oxygen Evolution Reaction" Catalysts 11, no. 12: 1523. https://doi.org/10.3390/catal11121523

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