D. K. Sarkarab,
V. Selvanathan*c,
M. Mottakinad,
A. K. Mahmud Hasana,
Md. Ariful Islama,
Hamad Almohamadi*e,
Nabeel H. Alharthifg and
Md. Akhtaruzzaman*ah
aSolar Energy Research Institute, Universiti Kebangsaan Malaysia, Bangi, Selangor Darul Ehsan 43600, Malaysia. E-mail: akhtar@ukm.edu.my; vidhya@ukm.edu.my
bDepartment of Applied Chemistry and Chemical Engineering, Rajshahi University, Rajshahi-6205, Bangladesh
cInstitute of Sustainable Energy, Universiti Tenaga Nasional (The Energy University), Jalan Ikram-Uniten, Kajang 43000, Selangor, Malaysia
dDepartment of Applied Chemistry and Chemical Engineering, Bangabandhu Sheikh Mujibur Rahman Science and Technology University, Gopalganj-8100, Bangladesh
eDepartment of Chemical Engineering, Faculty of Engineering, Islamic University of Madinah, Madinah, Saudi Arabia. E-mail: hha@iu.edu.sa
fDepartment of Mechanical Engineering, Faculty of Engineering, Islamic University of Madinah, Madinah, Saudi Arabia
gDepartment of Mechanical Engineering, College of Engineering, King Saud University, Riyadh, 11421, Saudi Arabia
hGraduate School of Pure and Applied Sciences, University of Tsukuba, Tsukuba, Ibaraki 305-8573, Japan
First published on 23rd June 2023
This study represents a green synthesis method for fabricating an oxygen evolution reaction (OER) electrode by depositing two-dimensional CuFeOx on nickel foam (NF). Two-dimensional CuFeOx was deposited on NF using in situ hydrothermal synthesis in the presence of Aloe vera extract. This phytochemical-assisted synthesis of CuFeOx resulted in a unique nano-rose-like morphology (petal diameter 30–70 nm), which significantly improved the electrochemical surface area of the electrode. The synthesized electrode was analyzed for its OER electrocatalytic activity and it was observed that using 75% Aloe vera extract in the phytochemical-assisted synthesis of CuFeOx resulted in improved OER electrocatalytic performance by attaining an overpotential of 310 mV for 50 mA cm−2 and 410 mV for 100 mA cm−2. The electrode also sustained robust stability throughout the 50 h of chronopotentiometry studies under alkaline electrolyte conditions, demonstrating its potential as an efficient OER electrode material. This study highlights the promising use of Aloe vera extract as a green and cost-effective way to synthesize efficient OER electrode materials.
Water splitting (WS) takes place through the OER and hydrogen evolution reaction (HER), where the OER part is most challenging because of its complexity and sluggish chemical kinetics process.5 The practically higher overpotential (OP) of the OER step consumes a large amount of energy for the operation, which is related to the high production cost of hydrogen. Without catalytic support, the whole production is not economically viable. Therefore, searching for suitable electrocatalysts is an attractive point to the researchers. The best-performing electrocatalysts (ECs) are noble metal-based materials such as RuO2, IrO2,7,8 AgPtOx for OER9 and Pt, Au, Pd, Ag, AgPtOx, AgAuO, AgPdOx, etc. for HER.10–12 The abovementioned ECs regarded as state-of-the-art and exhibit extraordinary performance in catalytic water splitting. But their extreme scarcity due to less natural resources, high cost, and inferior stability make a barrier to use them in large-scale hydrogen production.6 Nowadays, various earth-abundant transitional metals like Cu, Fe, Co, Ni, etc.13–18 with their alloys, oxides,19–21 chalcogenides,14 oxyhydroxides, hydroxides,18,21 phosphides,22 metal organic framework (MOF)23 have been reported as very much promising ECs owing to their performance and excellent stability. Recent reports discuss significant advancements in modifying catalyst surfaces and developing their morphology to yield electrocatalyst with expanded surface area.24,25 Generally, the use of organic molecules as capping and complexing agents are commonly employed strategies to achieve interesting morphologies of electrocatalysts directly grown on conductive substrates.26–28 The decoration of foreign materials in such nanostructures play a vital role in the OER catalysis process.29 Introducing inexpensive and easily accessible transitional metals (TMs) into a multiphase system is considered more beneficial for creating more active points that enhance electrical conductivity and ultimately lead to improved performance.21,30,31
Particularly, the combination of copper and iron for electrocatalytic material have shown encouraging results. In 2020, Xu et al. developed selenium enriched copper-iron selenide on copper foam as highly active catalysts for oxygen evolution through the optimization of surface morphology.32 The unique composition recorded ultralow overpotential of 200 mV at 10 mA cm−2. Inspired by these findings, in this work, a combination of copper-iron oxides synthesized via hydrothermal in different chemical environment is explored for catalyzing the OER. Some of the commonly used complexing agent used in hydrothermal synthesis of binder-free metal oxides-based electrocatalyst include urea, ammonium fluoride (NH4F), cetyltrimethylammonium bromide (CTAB), sodium dodecyl sulfate (STS) and polyvinylpyrrolidone (PVP).33–37 However, in line with the recent awareness towards green chemistry, these synthetic additives can be substituted with green complexing agent derived from phytochemicals which in fact have been used commonly in synthesis of metal oxide nanoparticles. The selection of plant extract utilized during the synthesis procedure plays a vital role in determining how ions are reduced, as well as in capping and stabilizing metal oxide nanostructures. Ultimately, this choice influences the overall morphology of the nanostructure.38,39
Therefore, in this work, the potential of using Aloe vera extract complexing agent for efficient morphology engineering is studied. Aloe vera gel primarily consists of water and polysaccharides like pectins, cellulose, hemicellulose, glucomannan, and acemannan. Additionally, the aloe latex contains hydroxyanthracenic derivatives, anthraquinone glycosides, and emodin.40,41 The unique chemical cocktail in the Aloe vera extract is expected to facilitate the formation of distinct nanostructured morphology that will boost its electrocatalytic activity. The catalysts were prepared through in situ hydrothermal and solvothermal processes, and five samples were produced, namely CuFeOx-A, CuFeOx-B, CuFeOx-C, CuFeOx-D, and CuFeOx-E, which varied based on the percentage of Aloe vera extract used. To evaluate the impact of the natural complexing agent on the catalysts' performance, the samples underwent characterization using X-ray diffraction (XRD), field effect scanning electron microscopy (FESEM), energy dispersive X-ray (EDX), and the necessary electrochemical tests. The study provides valuable information on the use of green synthesis of bimetallic heterostructure electrocatalysts for efficient water splitting.
ERHE = EAg/AgCl + 0.695 + 0.0591 × pH |
OER performance was examined using linear sweep voltammetry (LSV) at 2 mV s−1. Before final LSV measurements, the electrodes were scanned from 0.3 to 0.8 V (vs. SAACE) 20 mV S−1 until the electrode became stabilized. Tafel slopes were calculated from the nearest static LSV data. Overpotential (OP) is derived from the recorded LSV results based on the equation below.
OP = ERHE − 1.23 V |
The electrochemical active surface area (ECSA) was derived based on the double-layer capacitance (Cdl) by conducting a series of CV at the scan rate of 30 to 80 mV s−1 in a non-faradaic region (0.157–0.357 V vs. SAACE). The electrochemical impedance spectra (EIS) were evaluated within the frequency range from 100 mHz to 100 kHz. Finally, the stability measurement test was carried out by chronopotentiometry (CP) at the applied voltage of 0.35 V to achieve a constant current density (J) of 1.0 mA cm−2.
Fig. 2 XRD pattern of the synthesized CuFeOx-based electrode samples and NF; (a) from 10° to 80°, (b) from 42° to 46°, (c) from 50° to 55°, and (d) from 70° to 80°. |
These all are assigned signals of CuFeOx-based material (CuFeO2, Ni:CuFeO2, NiFeO2, etc.) on the NF, but their peak intensity is very diffused due to the presence of very high intensities of NF crystallinity.51 In the preparation process, DI water and Aloe vera–ethanol extract and their mixture are used as the solvent. Phytochemicals in Aloe vera–ethanol extract (AE) influences the crystallinity and material composition of the prepared catalyst samples in presence of solvent polarity. In Fig. 2(b), it can be found that the highest peak signal of all (CuFeOx-A, CuFeOx-B, CuFeOx-C, CuFeOx-D, and CuFeOx-E) samples on the plan (111) has shifted toward a lower diffraction angle from pure NF's peak. The main reasons for the peak shifting are the development of larger crystalline defects and the diffusion of Ni2+/Ni3+ in CuFeOx phases.52,53 Similar cases have been seen for the peaks in indexed planes (200), and (220) assigning Fig. 2(c) and (d). The characteristic XRD signals of CuFeOx-based delafossite crystal are observed in plans (015), (018), and (012) which are too defused compared to the peaks in plan (111) as shown in Fig. 2(b)–(d). The diffraction peaks of the agglomerated nanosheets (as seen in the FESEM image) of the CuFeOx-C sample are clearly distinguishable from those of the other samples, except for NF. The lower XRD intensity seen in samples B and C in Fig. 2. Lower crystallinity, smaller particle size, the presence of structural defects or impurities, and various crystallographic phases are all potential causes. These variables may affect the overall crystalline order, which can result in decreased XRD intensity. Besides these, some other unidentified diffraction peaks in planes (yzx), (zyx), (yxz) are found in the lower 2θ angle which required more studies. The overall XRD study of the prepared catalyst materials illustrates that the deposited films are very thin (low peak intensity) with mostly polycrystalline, and multiphases.
The FESEM images in the series shown in Fig. 3(d) illustrate that numerous highly porous interconnected nanosheets have grown vertically on the 3D NF. The diameter of the nanosheets is 30–70 nm. The vertically linked nanosheets have made funnel-types of hollow space insides themselves. As a result, electrode–electrolyte contact surface area increases many folds that enhance the sluggish OER kinetics leading to the lowest OP for WS. A solvent mixture of 25% DI water and 75% AE was used in the synthesis process, where the phytochemicals in AE played the most effective role in creating the optimal composition and morphology of the product. It may be hypothesised that the particular solvent composition has an impact on the rose-like morphology seen with a 75% AE and 25% DIW solvent combination in the FESEM investigation. The formation of unique patterns may result from the favourable nucleation and growth circumstances this particular solvent ratio. The rose-like shape formation may be greatly influenced by the variable surface tensions, reaction rates, and templating effects caused by the solvent combination. Separately, 25% and 50% AE were mixed with DIW and were applied to the samples CuFeOx-B, and CuFeOx-C. In both cases, nanostructures have changed distinctly seen in Fig. 3(b) and (c) but activities are limited for smaller active areas due to the large aggregation of nanoparticles in sample CuFeOx-B, and that of solid nanosheets in the sample CuFeOx-C. Their OER performances, as shown in Table 1, are higher than those of samples CuFeOx-A and CuFeOx-E, in which no mixed solvent was used. The attributed OP and relevant nano structuring morphology of the FESEM study ensured that the chemicals in AE explored their best activities by the polarity of the solvent. The exact mechanism of the mixed solvent demands extensive investigation of experiments.
Sample | % (DIW + AE) | % Ni | % Cu | % Fe | % O | (OP)50 | (OP)100 |
---|---|---|---|---|---|---|---|
a Elemental composition of synthesized electrode samples based on EDS and OP for a J of 50 mA cm−2 and 100 mA cm−2. | |||||||
CuFeOx-A | 100 + 0 | 8 | 69 | 1.7 | 22 | 530 | 690 |
CuFeOx-B | 75 + 25 | 25 | 53 | 2.8 | 18 | 400 | 530 |
CuFeOx-C | 50 + 50 | 56 | 22 | 3 | 19 | 340 | 450 |
CuFeOx-D | 25 + 75 | 35 | 24 | 18 | 22 | 310 | 410 |
CuFeOx-E | 0 + 100 | 25 | 53 | 1 | 16 | 420 | 510 |
Table 1 shows that sample CuFeOx-D with 75% AE is shown to be the best electrocatalyst. Based on the elemental compositions presented in Fig. 4, it is evident that the composition of solvent plays a vital role in the elemental content of the samples. The samples were prepared in different solvent medium comprising of various amounts of water and ethanol based-Aloe vera extract. It is evident that as the solvent content is increased from 0–75% Aloe vera extract, the composition of iron content increases and copper content decreases in the sample, reaching 18% of iron and 24% copper in CuFeOx-D. Realizing that the composition of oxygen remain fairly consistent for all the samples, it is proposed that in samples CuFeOx-A, CuFeOx-B and CuFeOx-D, the significantly high copper percentages indicate formation of copper rich-CuFeOx. In CuFeOx-C and CuFeOx-D, the percentages of copper and iron content in comparison to oxygen suggests the formation of CuFeOx with ideal stoichiometry which favours catalytic properties. In addition to the morphological property, the synergistic effect of both copper and iron in CuFeOx-D assist to improve the overall catalytic performance of the sample. At 100% Aloe vera extract, the amount of iron dropped again significantly, indicating that in the absence of water and presence of only Aloe vera, the incorporation of iron into copper iron oxide is not favoured. We suspect that the higher viscosity of solvent comprising only Aloe vera extract as well as the presence of ethanol as the primary solvent in CuFeOx-E is not conducive for the uniform formation of copper iron oxide on nickel foam substrate. The nickel content in the samples mainly originated from the nickel foam substrate and thus the higher percentages of nickel in CuFeOx-C is indicative of the formation of a thin layer of copper iron oxide layer on the nickel foam surface and possibility of exposed nickel foam surface. Fig. S1† displays the EDX spectra of the prepared samples.
With EDX mapping, it is possible to map a sample's elemental distribution. It focused electron beam across the sample and detecting the distinctive X-rays given off by the elements. The distribution of each element throughout the sample surface is displayed on the resultant map.
According to Fig. 5, the CuFeOx-based electrocatalyst seems to have a regular distribution of components over the surface of the electrode. The distribution of Cu, Fe, and O on the surface appears to be uniform according to the elemental map, which means that there are no obvious regions of high or low element concentration. The distribution of elements is found to be most homogeneous but the metal content in each sample is dominated by the amount of AE present in the mixed solvent. In addition, the electrode's surface seems rough rather than smooth, indicating that the electrode has a larger surface area. Electrocatalysts with greater surface areas are preferred because they can offer more active locations where catalytic reactions can take place. This study enhanced the structural integrity and porosity of the electrodes, by depositing very thin dendritic and porous Fe, Ni, and Cu nanostructures on the NF. This was accomplished by adopting a quick and easy in situ solvothermal process in which O2 bubbles serve as templates for the development of pores. Here, four different catalysts based on AE phytochemicals were prepared on high-surface-area support, and in each case, they should generate larger double-layer capacitances, indicating that the active sites are more accessible.
Fig. 6 XPS analysis of CuFeOx-D sample; (a) CuFeOx-D survey peak, (b) Cu 2p, (c) Ni 2p, (d) Fe 2p, (e) O 1s. |
The electrode CuFeOx-D has achieved high performance due to its material composition and morphological features. The compositional aspects have been discussed in the EDS section. CuFeOx-D shows superior performances because of its large surface area and high Fe content, as it is found in FESEM image and EDS results (Table 1). A self-supported conductive network was created by the interconnected nanosheets. This network aided in the transmission of electrons between the catalyst's surface-active sites and the current collector. As the OP increased, a lot of oxygen bubbles were produced on the electrode surface. It prevented the electrolyte from making direct contact with the active sites, which had an impact on how long the reaction would last. As a result, the electrodes' ability to transfer mass and catalysed reactions depended on the release of bubbles. The surface of the Pt electrode contained hydrogen bubbles that had grown and accumulated, as shown in a digital photograph. Moreover, the CuFeOx-D/NF-produced bubbles quickly disappeared from the electrode surface (Fig. 8), suggesting that the active sites may be quickly re-exposed to the electrolyte. The rapid surface bubble ejection demonstrated that the interconnected nanosheets design of CuFeOx-D/NF effectively boosted reaction kinetics and facilitated mass transfer.
Fig. 8 During water splitting in 1.0 M KOH electrolyte; O2 evolves at CuFeOx-D and H2 at Pt electrodes. |
The Tafel equation (OP = a + blogJ) is used to determine the Tafel slopes of the samples, which are displayed in Fig. 7(c). Here, a is the intercept-constant and b is the Tafel slope. Fig. 7(d) displays the corresponding values of the calculated slopes. The Tafel slope for the sample CuFeOx-D is the lowest (46 mV dec−1) and the respective values of other samples are 129 (CuFeOx-A), 72 (CuFeOx-B), 83 (CuFeOx-C), and 117 mV dec−1 (CuFeOx-E). A higher value of Tafel slope indicates faster electron transport, which in turn suggests a favourable kinetic barrier for the OER.63 Therefore, the smaller Tafel slope value denotes the better OER catalytic reaction.
Fig. 7(e) displays the Nyquist plot of the samples, which was derived from the EIS data. EIS is carried out to investigate the ohmic resistance (Rohm) of the solution/electrode surface and the faradaic charge transfer resistance (Rct). The semicircle part in the Nyquist plots represents that the faradaic charge transfer process is the rate-determining step for OER.49 In the high-frequency range, the intercept of the real axis (Z′) with the semicircle section that affects Rohm represents the sum of electrode resistance (Re) and solution resistance (Rs) connected in series (i.e., Rohm = Re + Rs).50 Fig. 7(e) shows that a small semicircle with a 0.12 Ω Rct is observed. All the prepared samples contain small semicircle and a straight line, indicating better charge transfer through the electrode. CuFeOx-D has the lowest Rohm value of 1.08 Ω cm−2 among all the prepared catalyst electrodes, while CuFeOx-A, CuFeOx-B, CuFeOx-C, and CuFeOx-E have significantly higher values of 2.84, 1.52, 1.43, and 1.91 Ω cm−2, respectively, which indicating a quicker reaction kinetics and superior low resistance for the CuFeOx-D electrode.
The catalyst's ECSA addresses the intrinsic activity and performance of OER ECs. Data from cyclic voltammetry (CV) tests carried out in the non-faradaic potential range were used to compute the ECSA of the samples based on the double-layer capacitance (Cdl). As shown in Fig. 9, the CV tests were carried out on all samples at potentials ranging from 0.15 V to 0.30 V vs. SAACE and at scan speeds of 30, 40, 50, 60, 70, and 80 mV s−1. The Cdl was then calculated from the slope of the linear plot of J = (Ja − Jc, Ja and Jc represent anodic and cathodic current densities) at 0.25 V vs. SAACE as a function of the sweep rate.48 The Cdl value were used to compute the ECSA. ECSA is equal to Cdl/Cs, where Cs is the specific capacitance, which is typically equal to 0.040 mF cm−2 for metal electrodes in KOH solution. Fig. 9(e) shows the ECSA values of the Cdl curves of the samples. The calculated ECSA of the samples are 46.75 mF cm−2 (CuFeOx-A), 93.25 mF cm−2 (CuFeOx-B), 117.15 mF cm−2 (CuFeOx-C), 162.7 mF cm−2 (CuFeOx-D), and 136.5 mF cm−2 (CuFeOx-E). The ECSA is directly proportional to active sites so greater ECSA means a larger area of catalyst surface is exposed to reactants species as a result, OER becomes faster.24 The performance of the synthesized materials with previously reported similar Cu-based polymetallic OER catalysts are listed in Table 2. In comparison to the literature, the CuFeOx-D material has a commendable efficiency with overpotential of 175 mV at 10 mA cm−2, implying that it could be a promising candidate for electrocatalytic applications. Furthermore, the reported stability of 50 hours demonstrates its robustness and durability under operational conditions.
Fig. 9 CV polarization curve at various scan rates; (a) CuFeOx-A, (b) CuFeOx-B, (c) CuFeOx-C, (d) CuFeOx-D, and (e) CuFeOx-E. |
No. | Material | η (mV)/Jsc (mA cm−2) | Stability | Preparation process | Ref. |
---|---|---|---|---|---|
1 | Cu–NiFe LDH/NF | 199/10 | 24 h | Chemical oxidation | 64 |
2 | Cu(OH)2:Fe(OH)3/CF | 365/10 | 12 h | In situ hydrothermal | 65 |
3 | NiFe/Cu2O NWs/CF | 215/10 | 25 h | In situ hydrothermal | 66 |
4 | FeCoNiMnCu | 280/10 | 40 h | Cathodic plasma electrolysis deposition | 67 |
5 | Cu(OH)2–NiFe LDH | 283/10 | 10 h | Unipolar pulse electro-deposition | 68 |
6 | CuFe2O4/NF | 340/10 | 1000 c | Electro-spume | 69 |
7 | CaCu3Fe4O12 | 450/05 | 11 h | Ion assisted solvothermal | 70 |
8 | CuO@Ni/NiFe (OH)x | 230/10 | 16 h | Chemical oxidation–calcination | 71 |
9 | Cu0.3Ir0.7Oδ | 150/100 | 6000 s | Hydrothermal doping | 72 |
10 | RuO2·NiO/NF | 144/10 | 72 h | In situ grown | 73 |
11 | RuO2·Ru | 172/10 | 10000 cycle | Laser ablation | 74 |
12 | RuO2 | 320/10 | 20 h | Electro-chemical | 75 |
13 | RuO2/CeO2 | 350/10 | 70 h | In situ solution | 76 |
14 | Ru–RuO2/CNT | 210/10 | 30 h | In situ hybrid | 77 |
15 | CuFeOx-D | 175/10 | 50 h | In situ solvothermal | This work |
Fig. 10 (a) Chronopotentiometry (CP) for the sample CuFeOx-D at a constant J of 100 mA cm−2 for 50 hours, (b) XRD diffractogram and (c) FESEM image of CuFeOx-D after 50 hours stability studies. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ra02512h |
This journal is © The Royal Society of Chemistry 2023 |