Next Article in Journal
Enhancement of Photocatalytic and Anticancer Properties in Y2O3 Nanocomposites Embedded in Reduced Graphene Oxide and Carbon Nanotubes
Previous Article in Journal
The Key Role of Carbon Materials in the Biological and Photocatalytic Reduction of Nitrates for the Sustainable Management of Wastewaters
Previous Article in Special Issue
Engineered NiCo2O4 Spinel Nanostructures for Enhanced Electrocatalytic Performance in Energy Storage and Non-Enzymatic Glucose Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of NiCu–Polymeric Membranes for Electro-Oxidizing Ethylene Glycol Molecules in Alkaline Medium

1
Department of Chemical Engineering, College of Engineering and Computer Sciences, Jazan University, Jazan 45142, Saudi Arabia
2
Chemistry Department, Faculty of Science, Cairo University, Giza 12613, Egypt
3
Engineering and Technology Research Center, Jazan University, P.O. Box 114, Jazan 82817, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(10), 959; https://doi.org/10.3390/catal15100959
Submission received: 28 August 2025 / Revised: 24 September 2025 / Accepted: 25 September 2025 / Published: 6 October 2025

Abstract

Binary metallic nickel–copper nanocatalysts were anchored onto a polyvinylidene fluoride-co-hexafluoropropylene membrane [NiCu/PVdF–HFP] using the electrospinning technique, followed by the chemical reduction of the relevant precursor salts by introducing sodium borohydride to the synthesis mixture. A series of varied Ni:Cu weight % proportions was developed in order to optimize the electroactivity of this binary nanocomposite towards the investigated oxidation process. A number of physicochemical tools were used to ascertain the morphology and chemical structure of the formed metallic species on polymeric films. Cyclic voltammetric studies revealed a satisfactory performance of altered NiCu/PVdF–HFP membranes in alkaline solution. Ethylene glycol molecules were successfully electro-oxidized at their surfaces, showing the highest current intensity [564.88 μA cm−2] at the one with Ni:Cu weight ratios of 5:5. The dependence of these metallic membranes’ behavior on the added alcohol concentration to the reaction electrolyte and the adjusted scan rate during the electrochemical measurement was carefully investigated. One hundred repeated scans did not significantly deteriorate the NiCu/PVdF–HFP nanostructures’ durability. Decay percentages of 76.90–87.95% were monitored at their surfaces, supporting the stabilized performance for prolonged periods. A much-decreased Rct value was estimated at Ni5Cu5/PVdF–HFP [392.6 Ohm cm2] as a consequence of the feasibility of the electron transfer step for the electro-catalyzing oxidation process of alcohol molecules. These enhanced study results will hopefully motivate the interested workers to explore the behavior of many binary and ternary combinations of metallic nanomaterials after their deposition onto convenient polymeric films for vital electrochemical reactions.

Graphical Abstract

1. Introduction

The continuous use of fossil fuels has led to many environmental problems arising. This has directed many research scientists to seek out suitable energy sources that have good stability and increased yield [1,2,3,4]. Direct alcohol fuel cells have many advantages such as being simple to handle, promising efficiency for energy transformation, increased power density and good safety [5,6,7,8,9]. Among different alcohols, ethylene glycol appears as a promising fuel with its encouraging features of reduced toxicity, nonvolatile nature, increased energy density and boiling constant [10,11,12,13,14]. Platinum-containing nanomaterials are considered the most popular anodes for direct alcohol fuel cells. However, their increased cost and reduced reaction kinetics might significantly retard their widespread application. These problems account for the need to explore inexpensive nanocatalysts with satisfactory electrocatalytic performance [15,16].
Numerous preparation schemes have been proposed to design nickel and nickel oxide-containing nanostructures as a result of their enhanced activity and stability towards many catalytic processes [17]. Active surfaces of nickel species were sufficient to improve antibacterial efficiency [18], alcohol oxidation [19,20], hydrogen evolution [21], oxygen evolution [22] reactions and energy storage applications [23,24]. For example, Nagajyothi et al. [25] have examined the electroactivity of hydrothermally prepared NiFe2O4 nanoparticles for electro-oxidizing a number of alcohols such as ethanol, methanol and propanol in a basic solution. Significant stability was shown during the operation of this nickel ferrite nanocatalyst within 50,000 s. A remarkable reduction in the measured Eonset value was also demonstrated, especially when methanol molecules were introduced to the investigated medium. Loading nickel–molybdenum oxide onto nickel foam after its doping with nitrogen displayed a promoted alcohol oxidation current density of 100 mA cm−2 at a decreased potential of 1.338 V (RHE). An increased faradic efficiency for electro-oxidizing benzyl alcohol species was also measured [98.7%] [26]. Incorporating tungsten into the chemical structure of electrodeposited nickel nanoparticles onto a carbon steel surface appreciably facilitated the surpassing of the accompanying energy barrier with the cleavage step of C–C bonds. An ethanol oxidation current density of 160 mA cm−2 was monitored at 1.7 V (RHE) with a superior durability performance within 2000 cycles [27]. Electrodeposited Ni0.8Co0.2(OH)2 electrodes exhibited an increased conversion rate of glycerol molecules as 99.8% after 4 h through an energy consumption of 36.38 W mol−1 [28]. Ag/NiO/CNTs nanocomposites in different proportions showed encouraging activities towards the ethanol conversion process—the presence of 25 wt.% Ag was sufficient to increase the oxidation current density at the formed nanomaterial by about five times [29]. NiMn/CNFs nanocatalysts were synthesized by the electrospinning process for the relevant acetates of nickel and manganese in different weight percentages at 20 kV with further heating for 300 min at 850 °C. Fruitful activity towards oxidizing ethylene glycol molecules was measured while increasing the incorporated Mn content in the fabricated NiMn/CNFs nanomaterials. The electrode that had 60Ni:40Mn composition showed an increased oxidation current density and good stability behavior [30]. Mixing spinel NiCo2O4 nanopowders with nitrogen-doped carbon nanotubes and graphene effectively motivated their oxidation activity for ethylene glycol molecules with Eonset value of 1.08 V (RHE) [31].
Conducting polymers have been recently employed in numerous fields such as sensors [32], energy storage [33,34], photocatalysis [35] and medical purposes [36]. Their specific mechanical, optical, electronic and magnetic features elect polymer-based nanostructures as promising catalysts with increased surface areas and surprising reactions [37,38]. Polyaniline [39,40], polypyrrole [41], poly(3,4-ethylene dioxythiophene) [42], polydiphenylamine [43] and poly(vinyl alcohol) [44] significantly improve the electrocatalytic behavior of carbonaceous supports [45,46]. Polyvinylidene fluoride-co-hexafluoropropylene is a linear polymer chain that contains H, C and F atoms [47]. Its satisfactory chemical and heat resistances, as well as increased piezoelectricity and flexibility, account for the ability to exploit PVDF–HFP films in energy harvesting, sensors and antibacterial media [48,49]. Supporting PVdF–HFP membrane with montmorillonite effectively decreases its porosity while improving the thermal stability of the fabricated composite when applied for direct alcohol fuel cells. The dissociation degree of sulfonic acid was extensively promoted by the action of adsorbed water molecules that, in turn, increased the ionic conductivity of the formed film [50]. Activated PVdF–HFP surface with p-toluene sulfonic acid/phosphoric acid/sulfuric acid displayed an ionic conductivity of 27.27 mS cm−1 and decreased permeability of methanol molecules as 9.7 × 10−8 cm2 s−1 [51]. Alloyed Ni-Co nanoparticles were deposited onto PVdF–HFP membrane and examined for glucose electro-oxidation reaction in a basic medium. A surprising detection performance was monitored towards the sugar molecules within a linear concentration region as 1 μM–7 mM [52]. Anchoring copper nanoparticles onto this film also motivated its activity to determine glucose concentration down to 11 nM with an excellent sensitivity of 506.62 μA mM−1 cm−2 [53]. Ren et al. [54] have built a nanocomposite of nitrogen-doped carbon quantum dots with Bi2WO6 and loaded it onto PVDF–HFP film surface. About 85.3% of oxytetracycline was adsorbed by this modified membrane with water flow of 34.4% and visible light of 63.1%. An acceptable reusability was also observed for this studied film when operated for numerous cycles.
Modifying the morphological structures of nickel-dependent nanocatalysts by alloying with numerous metals provides increased catalytic active sites for improved reaction rates. Copper is a low-cost transition metal that sufficiently participates in valuable purposes such as oxygen evolution [55], sensors [56], water splitting [57] and electro-oxidation processes [58,59,60]. The introduction of copper into NiFeCu oxohydroxides structure was beneficial in modifying the d-band center of nickel atoms and reducing the strength of adsorbed byproducts onto the fabricated electrode surface [61]. Promoted mechanical stability and electrical conductivity were also developed after doping NiO membranes with copper species [62]. Cao et al. [63] have measured an increased electrochemical active surface area at electrodeposited NiCuO nanoflowers as 85.9 m2/g. It accounted for their surprising activity towards glucose molecules oxidation reaction with a mass activity of 376.80 A/g and facile charge transfer characteristics. Hierarchical NiCu arrays were effectively employed to produce adipic acid and hydrogen with valuable stability within 200 h and increased yield. In situ spectroscopic investigation indicated that Cu0.81Ni0.19 nanoalloy motivated the conversion reaction kinetics with enhanced free energy during the intermediate adsorption step [64]. Ni4Cu1 deposits onto polyphenylenediamine activated ethanol oxidation process with reduced onset potential of 390 mV and effective current density of 70.5 mA/cm2 [65]. By considering these fruitful results, a series of NixCu1–x nanodeposits onto polyvinylidene fluoride-co-hexafluoropropylene membrane was evaluated for electro-catalyzing ethylene glycol oxidation reaction in basic solution. These polymeric films were first synthesized using the electrospinning method, followed by the formation of binary nickel–copper coatings via the chemical reduction of their precursor salts by sodium borohydride. These metallic polymeric films were characterized through scanning electron microscopy (SEM), energy dispersive X-ray analysis (EDX), X-ray diffraction (XRD) and Fourier Transform Infrared spectroscopy (FTIR). Various electrochemical tools involving cyclic voltammetry, chronoamperometry and electrochemical impedance spectroscopy measurements were conducted to measure the electroactivity of these polymeric nanostructures and optimize their performance.

2. Results and Discussion

Numerous advantageous characteristics were gained when the electrospinning technique was employed to fabricate nanofibrous membranes. Promoted porosity, increased interconnectivity, satisfied flexibility and outperformed surface-to-volume ratios were provided by electrospun films [66]. PVdF–HFP, as one of these promising polymeric surfaces, displays suitable features including increased dielectric constant, good piezoelectric and pyroelectric properties, enhanced thermal stability and encouraging hydrophobicity. This might rationalize the motivated studies to create metallic nanofibers using this polymeric membrane [50,67]. SEM image in Figure 1a revealed the formation of homogeneous and bead-free fibrous mats of PVdF–HFP. The fast evaporation of acetone as the solvent, along with the jet movement from the positively charged syringe needle to the negatively charged collector, participated in creating nanoporous architecture during the electrospinning process. This might help to provide a suitable microenvironment for metal ions to settle. Incorporating hydrated precursor salts (nickel and copper acetates) also contributed to the phase separation process (liquid–liquid demixing) inside the hydrophobic polymeric matrix. This might increase the pores’ density and provide a chance for the analyte molecules to trap easily with much-reduced diffusion barriers [68]. The rate of the hydrogen evolution reaction was consistently enhanced, and the formed membrane conductivity was promoted as a result of inserting varied metallic species. The possibility of developing a gel state also facilitated the elongation of the axial gel as much as possible [53]. SEM pictures of different NiCu/PVdF–HFP membranes were presented in Figure 1b–e. The sample with Ni:Cu ratio of 3:7 [see Figure 1e] displayed rough and interwoven surfaces. The porous nature of these fibers enabled the distribution of some metallic deposits onto their external sides. The lowered reduction potential of copper ions relative to those of nickel favored their chemical reduction using sodium borohydride, especially when they were present in increased amounts [3Ni:7Cu]. Therefore, a rapid chemical reduction process was conducted, resulting in uncontrolled deposition of localized copper nuclei on the polymeric film surface. This irregular growth of metallic nanoparticles resulted in the formation of rough clusters. The EDX chart of Ni3Cu7/PVdF–HFP film was constructed at certain surface points as indicated in Figure S1. Definite peaks of nickel and copper were distributed along with the investigated EDX chart. Additionally, carbon and fluorine peaks appeared to characterize the fabricated polymeric surface. The XRD pattern of Ni3Cu7/PVdF–HFP membrane in Figure S2 demonstrated a number of defined peaks at 2θ values of 18.20°, 20.00°, 26.60°, 36.15°, 43.40° and 50.40°. These peaks match well those of Ni and Cu crystallites with the respective diffraction planes of (100), (020), (110), (021), (111) and (200) based on JCPDS cards of Ni(04-0850) and Cu(04-0836) [51,69]. Accordingly, a face-centered cubic structure was expected with comparable lattice parameters as 3.523 Å for nickel and 3.615 Å for copper. FTIR charts of PVdF–HFP membrane and its relevant metallic forms displayed some defined peaks as shown in Figure 2. The α and β crystalline phases were revealed through their corresponding peaks at 749 and 837 cm−1 [38]. Another two bands at 672 and 872 cm−1 were created due to CF and CH2 wiggling vibrations in the amorphous polymeric surface. On the other hand, symmetric C–F stretching, CF2 stretching and bending vibrations were clarified via characteristic bands at 1071, 1175 and 1400 cm−1 [70]. An extra band at 1561 cm−1 indicated that NiO and CuO species were formed in the studied nanofibrous network [71].
The cyclic voltammetric behavior of varied NiCu/PVdF–HFP electrodes in 0.5 M NaOH solution was investigated in Figure 3 at 5 mV s−1. All studied nanocomposites exhibited a defined redox couple for the electrochemical transformation of nickel ions between (2+) and (3+) oxidation states [73,74]. Ni5Cu5/PVdF–HFP membrane in Figure 3a displayed this Ni(OH)2/NiOOH conversion at potential values of 541 and 425 mV in the forward and backward directions, respectively. A more positive potential shift of 89 and 91 mV was monitored for the film that contained 90 wt.% Ni-10 wt.% Cu [see section (b)] for the corresponding oxidation and reduction processes of this redox transfer. By calculating ΔE of Ni(II)/Ni(III) redox reaction for different NiCu/PVdF–HFP films, values of 121, 116, 114 and 95 mV were estimated for those with Ni:Cu weight ratios of 3:7, 5:5, 9:1 and 7:3, respectively [see sections (c, d)]. These ΔE values were much lower than those reported at NiC/CNFs [231 mV], NiC-MnC/CNFs (95:5) [155 mV] and NiC-MnC/CNFs (90:10) [139 mV] nanomaterials to reveal the facile electron transfer reaction at NiCu-polymeric films [75]. Moreover, promoted current densities of 47.70 and −33.71 μA cm−2 were shown by Ni5Cu5/PVdF–HFP nanostructure in the forward and backward routes of the examined transformation to clarify the availability of an increased number of active surface sites with an acceptable reaction rate. However, the other samples with Ni:Cu weight ratios of 7:3, 3:7 and 9:1 had much lower oxidation current densities of 36.03, 32.28 and 29.56 μA cm−2, correspondingly, compared to values of −12.84, −12.49 and −9.31 μA cm−2 at these electrodes surfaces during the reduction step of Ni(OH)2/NiOOH conversion. This study result illustrates the preferential performance of Ni5Cu5/PVdF–HFP membrane during the formation of multiple layers of nickel hydroxide/oxyhydroxide species.
The electrochemical response of Ni5Cu5/PVdF–HFP nanomaterial was investigated in an alkaline electrolyte at varied scan rates in Figure 4a. The Ni(II)/Ni(III) transformation peaks were clearly shown at all chosen scan rates. Adjusting the scanned measurement at higher sweep rates fairly shifted the oxidation and reduction peaks of this redox couple towards more positive and negative potentials, respectively. The charge transfer step was greatly affected by the contact between the film surface and its surrounding medium [76,77,78,79]. This might account for this potential change in Ni(OH)2/NiOOH transition with altering the sweep rate, as inferred from Figure 4b. Additionally, a continuous current intensity enhancement was noticed for runs at higher sweep rates when the forward and backward directions of Ni(OH)2/NiOOH conversion were considered. These current densities were plotted against υ in Figure 4c. Straight lines were attained on both sides of Ni(II)/Ni(III) transformation. The coverage of film surface with the deposited metallic species (Γ*) was derived from the slopes of (I–υ) relations [80] as 1.928 × 10−12 mol cm−2. Another linear relation was also extracted when these currents were varied against υ0.5 in Figure 4d. This result supported the contribution of diffusion rules during nickel hydroxide/oxyhydroxide transition. Moreover, the Randles–Sevick equation was employed to calculate the diffusion coefficient of alcohol molecules within the Ni5Cu5/PVdF–HFP membrane as 1.954 × 10−4 cm2 s−1 [81,82]. This value greatly exceeded the reported one at deposited nickel oxide species onto activated carbon (3.79 × 10−10 cm2 s−1) and multi-walled carbon nanotubes (7.69 × 10−10 cm2 s−1) [83] and calcined NiCo2O4 (20 wt.%) nanoparticles onto graphite (1.2558 × 10−8 cm2 s−1) [84].
The cyclic voltammogram of Ni9Cu1/PVdF–HFP nanocatalyst in (1.0 M EG + 0.5 M NaOH) solution was presented in Figure 5a at 10 mV s−1. The alcohol oxidation process was started at 532 mV and a rapid current density increase was noticed to display an oxidation peak at 730 mV with a current density of 67.81 μA cm−2. The reduction step of NiOOH species was further detected in the reverse scan at 500 mV. The relevant activities of Ni7Cu3/PVdF–HFP, Ni5Cu5/PVdF–HFP and Ni3Cu7/PVdF–HFP films were demonstrated in Figure 5b–d. Their respective Eonset values were gradually shifted in the negative direction by 44, 24 and 72 mV when compared to the measured one at Ni9Cu1/PVdF–HFP electrode to clarify the feasibility of NiCu/PVdF–HFP film surfaces in different Ni:Cu proportions to electro-oxidize ethylene glycol molecules. The alcohol oxidation peak also revealed promising current densities for the membranes with Ni:Cu wt.% ratios of 7:3 [323.15 μA cm−2] and 5:5 [564.88 μA cm−2] showing 4.77 and 8.33 folds activity promotion when related to that at Ni9Cu1/PVdF–HFP. However, the insertion of 30%Ni:70%Cu in the fabricated polymeric membrane resulted in a decreased oxidation current density [50.17 μA cm−2] to suggest the suitability of moderate loadings of nickel and copper for better performance towards ethylene glycol molecules. A volcano-shaped plot was drawn when representing the alcohol oxidation current density at each chosen Cu wt.% value inside the formed membranes as indicated in Figure 5e. The loading of varied metallic ingredients into the examined nanostructures significantly controlled their activity. For example, Rodriguez et al. [85] have fabricated a series of Pt10−xFex/CNTs [x = 1, 2, 3, 4 and 5] nanocatalysts through the Brust–Schiffrin protocol and investigated their behavior for electro-oxidizing methanol molecules. These nanomaterials observed a volcano-shaped performance while altering the incorporated Fe:Pt ratio during the synthesis step. The presence of 30% Fe or less was sufficient for achieving enhanced alcohol electro-oxidation current density due to the favored rearrangement of Pt atoms by iron species with their promoted electronic environment. On the contrary, higher contents of Fe atoms [40% or more] might delay the rate of the alcohol dehydrogenation reaction, leading to decreased nanocatalyst activity. The mass activity of Pd4Au3 aerogels onto carbon support for oxygen reduction reaction was 11.6 and 17.4 folds higher than those at Pt/C and Pd/C, respectively [86]. Pd6Bi1 nanowires exhibited mass activities of 1231, 2066 and 3047 mA mgPd−1 for the corresponding electro-oxidation reactions of glycerol, ethanol and ethylene glycol molecules. Density functional theory studies clarified that altered d-band positions were created based on the introduced Pd:Bi stoichiometric ratio in the synthesized nanocomposite. This might vary the formed coordination environment of palladium species in order to get a balanced ratio between OH* and CO* radicals [87].
The influence of varying the alcohol concentration during the evaluation of the metallic membranes’ activity was studied using the cyclic voltammetry technique. The obtained cyclic voltammograms at Ni5Cu5/PVdF–HFP electrode in 0.5 M NaOH solution are demonstrated in Figure 6a at 10 mV s−1 after inserting 0.1, 0.3, 0.4, 0.5, 0.7 and 1.0 M EG molecules. A subsequent activity enhancement of this investigated nanocatalyst was noticed with increasing the added alcohol concentration in the supporting electrolyte. Recording this oxidation peak current density as a function of the selected alcohol concentration in Figure 6b confirmed this strong dependence in a linear tendency. The order of the oxidation process with respect to ethylene glycol concentration was derived from elucidating the slope of log (I)–log (C) plot in Figure 6c. An order value of 0.809 was deduced.
The durability of varied NiCu/PVdF–HFP nanocatalysts for prolonged working to electro-oxidize EG molecules in 0.5 M NaOH solution was evaluated using chronoamperometry and repeated cyclization techniques. Chronoamperograms of these membranes were operated at 600 mV for 1800 s in Figure 7a. The presence of appreciable contents of alcohol molecules at the electrode surfaces resulted in increased oxidation current densities at the first few moments of this study. They rapidly decreased, reaching a steady response after about 100–400 s, with a decaying oxidation rate by collecting its byproducts at the available reaction sites [88,89]. By considering the steady state oxidation current densities of NiCu/PVdF–HFP electrodes after 1800 s, introducing equal ratios of nickel and copper species into the polymeric film composition [50:50] demonstrated the highest stability within the prepared electrodes [126.47 μA cm−2]. This current intensity value outperformed the obtained one at Ni7Cu3/PVdF–HFP, Ni9Cu1/PVdF–HFP and Ni3Cu7/PVdF–HFP by 1.52, 4.56 and 6.60 folds, respectively. The superiority of Ni5Cu5/PVdF–HFP nanomaterial was also supported by estimating the decay percentages of altered electrodes at the end of the chronoamperometry test. The related oxidation current density at Ni3Cu7/PVdF–HFP, Ni9Cu1/PVdF–HFP, Ni7Cu3/PVdF–HFP and Ni5Cu5/PVdF–HFP decreased by 71.36%, 77.73%, 82.98% and 87.42% when compared to the recorded magnitudes at the start of the stability region. The cyclic voltammograms of varied NiCu/PVdF–HFP nanocomposites were also scanned for repeated 100 cycles at 10 mV s−1 in (1.0 M EG + 0.5 M NaOH) solution. The obtained result was demonstrated in Figure S3 to reveal the consecutive activity decline of all NiCu/PVdF–HFP electrodes upon cyclization for increased runs. The oxidation current densities of alcohol molecules were followed at different cycles and drawn in Figure 7b. After 100 cycles, satisfactory stability performances were achieved by the fabricated NiCu/PVdF–HFP films with percentages ranging between 76.90% and 87.95%. The incorporation of 50%Ni:50%Cu into the prepared polymeric film was recommended for its encouraging activity and durability to be optimized for further catalytic processes. The contribution of variable wt.% values of copper oxide species to the chemical composition of the investigated nanocatalyst appreciably affected its durability during the alcohol oxidation process. Our earlier studies demonstrated that the insertion of moderate proportions of copper oxide into the nanocomposite skeleton efficiently promoted its stability behavior. Adding 5, 7.5, 10, 12.5 and 15 wt.% copper oxide to nickel oxide nanostructure onto graphite substrate yielded respective steady state current densities of 1.51, 2.56, 2.80, 3.26 and 4.29 mA cm−2 when ethanol molecules were added to basic solution. However, higher metallic oxide contents might exhibit an adverse effect on the durability test, as concluded from the chronoamperometric studies [90]. Here, in this present work, a gradual improvement in the stability performance of NiCu/PVdF–HFP film was shown by introducing an increased copper oxide wt.% in the prepared polymeric surface up to 50%. The irregular distribution of catalytic nanoparticles for the nanostructures with higher oxide content (70 wt.% copper oxide), as revealed from the relevant SEM image, might reduce the available number of active sites for operating the oxidation reaction and impede the electron transfer process, resulting in decayed activity and stability performances.
Varying the scan rate while conducting the linear sweep voltammetric runs at Ni5Cu5/PVdF–HFP film in (1.0 M EG + 0.5 M NaOH) solution is shown in Figure 8a. This measurement parameter significantly affects the obtained oxidation peak current density and potential values. Increased membrane activities were monitored for the related voltammetric scans at higher rates. Sketching the obtained current intensities as a function of υ0.5 in Figure 8b resulted in a linear region as a consequence of the prevalence of diffusion rules when the alcohol molecules were electro-oxidized at the membrane surface. A continuous potential shift of the oxidation peak towards positive values was also observed at Ni5Cu5/PVdF–HFP film as its linear sweep voltammetric studies were working at faster rates. Drawing this oxidation peak potential against log υ in Figure 8c reflected this observation via a straight-line plot. The electron transfer coefficient (α) magnitude was extracted from this line slope that equals [2.303RT/(1–α)nF] [91]. It was found to be 0.410, as a comparable value to that reported when ethanol molecules were electro-oxidized at Ni/PVdF–HFP membrane [0.468] [92]. The linear sweep voltammogram at 1 mV s−1 was exploited to calculate some kinetic parameters, such as Tafel slope and the exchange current density magnitudes. The varied potentials against log I were followed for this curve in Figure 8d. A linear trend was observed with a Tafel slope value of 132.35 mV dec−1. Additionally, Ni5Cu5/PVdF–HFP film exhibited an exchange current density of 4.232 × 10−9 A cm−2. This result demonstrated an enhanced alcohol oxidation reaction rate at NiCu-polymeric surface when contrasting its Tafel slope to that reported at the relevant Ni/PVdF–HFP membrane [369 mV dec−1] [92].
The phase shift plots of varied NiCu/PVdF–HFP nanostructures were studied in basic solution after inserting subsequent alcohol concentrations in Figure 9a–d for the films with Ni:Cu weight ratios of 9:1, 7:3, 5:5 and 3:7, respectively. They were scanned within the frequency region of 10,000–0.1 Hz at 575 mV. Steady phase angle values were monitored irrespective of ethylene glycol concentration within the high-frequency region for all investigated NiCu/PVdF–HFP nanomaterials. This resistive performance was recorded at the outer electrode layer. A gradual phase angle increase was shown as the scanned frequency decreased. This phase shift might be linear when reduced ethylene glycol contents were present in the chosen medium to reflect the controlled electrode behavior by diffusion-limited kinetics. However, increased alcohol concentrations in NaOH solution were responsible for lowering the obtained phase angles, displaying a characteristic maximum within the intermediate and decreased frequency sections. This phase angle maximum was formed at increased frequency values as the amount of alcohol introduced to the basic electrolyte increased. This was clearly apparent for the metallic polymer surfaces with Ni:Cu weight ratios of 7:3 [see Figure 9b] and 5:5 [see Figure 9c]. This phase angle peak is a transition to a capacitive attitude within the electrode double layer. The influence of altering the metallic Ni:Cu proportions during the membrane’s fabrication onto their reported phase angles within the scanned frequency region was concluded from the corresponding diagrams of NiCu/PVdF–HFP electrodes in (0.3 M EG + 0.5 M NaOH) solution in Figure 9e. The increased copper content in the formed film significantly improved its activity as revealed from continuous phase angle decrease in the lowered frequency section for electrodes with Ni:Cu weight ratios of 9:1, 7:3 and 5:5. The highest catalytic behavior was assigned for the Ni5Cu5/PVdF–HFP membrane. At the same time, a linear trend was then shown for Ni3Cu7/PVdF–HFP electrode with increased phase angles to illustrate its lowered activity.
On the other hand, bode plots of Ni9Cu1/PVdF–HFP, Ni7Cu3/PVdF–HFP, Ni5Cu5/PVdF–HFP and Ni3Cu7/PVdF–HFP nanocatalysts were examined in an alkaline medium adding ethylene glycol molecules in increased proportions till 1.0 M; the obtained results are presented in Figure 10a–d. A silent electrochemical response was recorded for all prepared metallic membranes when higher frequencies were scanned. However, a fast increase in the measured impedance values was significantly observed when medium and low frequencies were operating. Moreover, log Z values were consecutively decreased for electrolytes with higher alcohol concentrations to monitor the lowest resistance magnitudes when 1.0 M ethylene glycol was inserted. This result significantly supported the promising activity of the fabricated nanocatalysts for electro-oxidizing ethylene glycol species. The Ni:Cu weight ratio in the evaluated NiCu/PVdF–HFP electrodes appreciably controlled the charge transfer rate of the oxidation process across their surfaces. This was derived from contrasting the related bode curves of altered NiCu/PVdF–HFP nanocomposites in (1.0 M alcohol + 0.5 M NaOH) in Figure 10e. A remarkable decrease in the noticed impedance values was shown as the copper content in the synthesized membrane increased to exhibit the lowest log Z magnitudes for the one with Ni:Cu weight ratio of 0.5:0.5. On contrary, the metallic film with a higher Cu content (70 wt.%) demonstrated much increased resistance when compared to the recorded values for those with reduced copper contribution. This situation was similar to that reported by our research group when studying the response of deposited NiCu oxide [90] and NiCo2O4 [84] nanoparticles onto graphite surface during ethanol oxidation reaction in an alkaline solution using EIS experiments. Including different proportions of transition metal oxides in these nanocomposites appreciably affected their measured resistance values. As concluded from the related results for copper- and cobalt-containing binary oxides, a gradual reduction in the obtained resistance value was monitored as the content of copper and cobalt in the synthesized nanostructures increased up to 15 and 20 wt.%, respectively. Above these proportions, the activities of formed nanomaterials declined, showing increased resistance.
The respective Nyquist curves of different NiCu/PVdF–HFP films were drawn using a basic solution with a gradual addition of increased volumes of ethylene glycol. The achieved curves are clarified in Figure 11 for the membranes with metallic weight % magnitudes of Ni:Cu as 9:1 1 [section (a)], 7:3 [section (b)], 5:5 [section (c)] and 3:7 [section (d)]. They all exhibited two depressed semicircles in the high and low frequency regions in the presence of altered alcohol concentrations; however, the measured resistances subsequently decayed as more ethylene glycol molecules were incorporated in the alkaline electrolyte. The diameter of the formed semicircle in the decreased frequency zone clearly clarified this conclusion. It continuously exhibited a reduced curvature reflecting lowered charge transfer resistances at the studied electrode surface. This Nyquist diagram was contrasted for the films with different Ni:Cu weight % ratios in Figure 11e when 1.0 M ethylene glycol molecules were present in NaOH solution. The Ni9Cu1/PVdF–HFP electrode had the largest curvature of recorded semicircles in the low-frequency section among the examined nanomaterials. This result supported its increased impedance for a facile charge transfer step during the alcohol oxidation reaction. Ni7Cu3/PVdF–HFP and Ni5Cu5/PVdF–HFP electrodes had much lower resistance values. Favored reaction kinetics were expected at the Ni5Cu5/PVdF–HFP film surface. Further increase in the copper amount in the fabricated membrane was not suggested for a valuable electro-oxidation process as a result of its increased resistance to an efficient electron transfer mechanism. This result was further confirmed by proposing a convenient electric circuit structure with specific components in appropriate connection to simulate the theoretical results with the obtained experimental data. The presented model in Figure 11f has two (R–CPE) sections: They are in a serial arrangement with a solution resistance component (Rs). Here, R1 and R2 reflect a numerical representation of the measured impedance values in the outer and inner regions of the electrode surface, respectively. Moreover, CPE expresses the related constant phase elements in these two sections with a good replacement for the capacitive role in the constructed circuit. The unfair distribution of the electrolyte impedance and double-layer capacitance was handled through inserting this CPE component instead of a capacitor [93]. By employing this electric circuit skeleton to the measured curves of varied NiCu/PVdF–HFP electrodes in NaOH solution with altered alcohol concentrations, the simulated values of these model constituents are presented in Table 1. A consecutive decay of R2 magnitudes was reported for each investigated metallic polymeric membrane as the content of alcohol molecules in the selected basic electrolyte increased. Varied electrode surfaces displayed the lowest resistance values in the presence of 1.0 M ethylene glycol. Under such experimental conditions, R2 value gradually decreased when Ni:Cu weight % ratio in the synthesized film was in the order: 9:1 [4230.0 Ohm cm2] > 3:7 [1954.0 Ohm cm2] > 7:3 [1056.0 Ohm cm2] > 5:5 [392.6 Ohm cm2] [see the listed results in Table 2].
The potential value at which the EIS experiment operated significantly affected the obtained charge transfer rate. This was clarified by conducting a series of EIS measurements for Ni5Cu5/PVdF–HFP film in (0.3 M ethylene glycol + 0.5 M NaOH) solution at a number of potentials (350–700 mV) in Figure 12. The phase shift of this membrane showed a gradual decrease in its phase angles in the middle- and low-frequency sections as the examined potential was shifted towards more positive values. A lowered phase maximum was also detected at increased potentials, reaching its most reduced value at 700 mV [see section (a)]. Moreover, the bode diagrams of Ni5Cu5/PVdF–HFP film at altered potentials presented in Figure 12b demonstrate decreased impedance values for the working measurements at more positive potentials. Within the detected potentials, the lowest resistance was attained at 700 mV. Nyquist curves in Figure 12c displayed subsequent reduction in the semicircles’ diameters as the swept potential was positively moved. The smallest resistance value was monitored for the metallic membrane surface at 700 mV. The constructed electric circuit in Figure 11f was applied for fitting these experimental EIS results with their corresponding theoretical data, and the derived elements’ magnitudes were shown in Table 3. A consecutive decrease in the charge transfer resistance was observed by positively adjusting the potential value during the EIS test. It exhibited 1.885 folds reduction for the operated experiment at 700 mV in relation to that at 350 mV.
Ni–Cu alloys were selected in this study instead of Co–Cu or Fe–Cu analogues. This choice was motivated by the well-established synergistic interaction between Ni and Cu, where Ni provides abundant Ni(II)/Ni(III) redox sites for electro-catalyzing the alcohol oxidation process. At the same time, Cu enhances the electrical conductivity of fabricated nanostructures and modulates the electronic structure of Ni. This in turn weakens the bonding interactions of the metallic sites with accumulated poisoning species [CO-like intermediates]. Accordingly, nickel–copper-containing nanocatalysts significantly display greater resistance to surface deactivation and maintain increased activity retention when applied for long-term electro-oxidation processes [94,95,96,97]. Moreover, reduced onset potentials and promoted current densities were measured at this binary system during alcohol oxidation reaction when contrasted to the reported results at Co- or Fe-incorporated nanomaterials. Integrating nickel–copper nanoparticles into the synthesized PVdF–HFP films significantly combines these enhanced features of metallic counterparts with the feasible ionic transport characteristics of polymeric supports to have a highly efficient electrocatalyst platform with a satisfactory durability. The synergy between nickel and copper oxides in the formed nanostructures originates from the availability of redox active sites of nickel hydroxide/oxyhydroxide species for promoting the alcohol electro-oxidation process rate, while the electronic conductivity of these nanocomposites was greatly enhanced by introducing copper oxide nanoparticles in appreciable contents, showing the optimum performance for the electrode surfaces with 50%Ni:50%Cu. A satisfactory durability behavior was also demonstrated by NiCu/PVdF–HFP membranes as a result of modifying the electronic environment of nickel sites via their d-band centers to reduce the accumulation of CO-like poisoning intermediates. The shortage of copper oxide content in the Ni9Cu1/PVdF–HFP electrode might account for its decreased steady state oxidation current density. In contrast, the inhomogeneous distribution of deposited copper oxide sites for membranes with excessive copper oxide wt.% values [Ni3Cu7/PVdF–HFP] [as demonstrated by the morphological studies] dissipated the active reaction centers. This explained the decayed activity and durability of electrodes with higher copper oxide loadings [above 50 wt.%]. EIS measurements of the alcohol oxidation reaction at varied NiCu/PVdF–HFP films supported the obtained results by cyclic voltammetry and chronoamperometry. Among all examined membranes, Ni5Cu5/PVdF–HFP displayed the lowest Rct magnitudes to illustrate the pronounced electron transfer rate at its surface with rapid reaction kinetics. The encouraging performance of deposited metallic species onto polymeric films might direct numerous interested research groups to fabricate comparable nanomaterials for different electrocatalytic fields.

3. Experimental Section

3.1. Chemicals

Poly(vinylidene fluoride-co-hexafluoropropene) with M.wt. of 65,000 g mol−1, Ni(II) acetate tetrahydrate, Cu(II) acetate tetrahydrate and sodium borohydride were delivered from Sigma-Aldrich (St. Louis, MO, USA). N,N-dimethylformamide (DMF) and acetone were purchased from Fluka. All chemicals were employed as received without additional treatment.

3.2. The Formation of Metallic PVdF–HFP Membrane Surfaces in Altered Ni:Cu Proportions

The polymeric film (15 wt.%) was firstly formed by inserting 1.5 g of PVdF–HFP powder inside a mixed solution of acetone and DMF in a volume proportion of 1:4. The metallic acetate precursors in varied weight percentage ratios [NixCu10−x, x = 3, 5, 7 and 9] were subsequently dissolved into DMF, followed by sufficient stirring to finally have a transparent mixture. The metallic salt solutions and polymeric sol gel were homogenized together by stirring overnight. A plastic capillary syringe was then filled with this mixture and fixed inside a lab-scale electrospinner. One section of this syringe contained a copper rod, while the other counterpart was connected to a power supply acting as the positive pole under an increased voltage. The negative electrode was joined to a ground iron drum with an Al sheet to collect the formed fibers. The electrospinning process was conducted at 20 kV with consecutive drying under vacuum at room temperature. The reduction step of incorporating metallic ions inside PVdF–HFP films was carried out using sodium borohydride solution with an increased molar ratio of five times compared to that of mixed metallic precursor salts. The polymeric films were immersed in this reduction solution until hydrogen molecules were completely evolved without further deposition of black spots onto the film’s surfaces. They were finally cleaned with double-distilled water and ethanol, and dried in an air oven within 300 min at 60 °C.

3.3. The Physical Characterization of NiCu/PVdF–HFP Membranes Using Convenient Analysis Tools

The crystallinity and the size of the formed metallic particles were investigated using an X-ray diffractometer (Rigaku Co., Tokyo, Japan). Cu Kα radiation with a wavelength of 1.54056 Å was employed. ATR-FTIR model “Nicolet iS 10” (Thermo Fisher Scientific, Waltham, MA, USA) was also connected to a specular reflectance for evaluating the incorporation of binary metallic species into the selected PVdF–HFP membrane. A wavenumber range extending from 500 to 4000 cm−1 was scanned. The morphology of fabricated nanomaterials was studied through scanning electron microscopy (SEM, Hitachi S-7400, Tokyo, Japan) in conjunction with energy-dispersive X-ray (EDX) equipment for the determination of numerous nanocatalyst elements distribution and their relative weight percentages.

3.4. The Electrochemical Investigation of Varied NiCu/PVdF–HFP Membranes Activity for Electro-Oxidizing Ethylene Glycol Molecules

The electrocatalytic activity of altered NiCu/PVdF–HFP films towards ethylene glycol electro-oxidation reaction was tested in an alkaline medium. Different electrochemical tools, including cyclic voltammetry, chronoamperometry, linear sweep voltammetry and electrochemical impedance spectroscopy were operated via Gamry potentiostat. A three-electrodes cell was constituted using NiCu/PVdF–HFP membrane [2.0 × 1.0 cm2] as the working electrode. It was fixed to a suitable copper wire to be connected to the potentiostat workstation. On the other hand, a Pt wire and a silver/silver chloride electrode were inserted in the electrochemical system as the counter and reference electrodes, respectively. A stabilization of the electrochemical performance of examined nanocatalysts was achieved after their continuous cyclization in 0.5 M NaOH solution at 50 mV s−1 for 40 cycles using a potential region from 0 up to 900 mV (Ag/AgCl). The optimum alcohol concentration was predicted by investigating the influence of changing ethylene glycol content in the supporting electrolyte on the nanomaterials’ activity. Cyclic voltammetric studies were also conducted at numerous scan rates during the oxidation process. The durability of NiCu/PVdF–HFP electrodes for efficient activity after more extended working periods was checked using the respective chronoamperograms at 575 mV for 30 min. A number of electrochemical impedance spectroscopy experiments were conducted at the fabricated NiCu/PVdF–HFP membranes’ surfaces to explore their impedance behavior towards differing alcohol concentrations and potential magnitudes during the electro-oxidation process. This whole study was carried out at 25 °C using an aerated electrochemical cell.

4. Conclusions

This work demonstrated the possibility of applying anchored NiCu nanoparticles onto electrospun PVdF–HFP membranes for a highly efficient and durable platform during the alkaline electro-oxidation process of ethylene glycol molecules. Among the investigated compositions, the Ni5Cu5/PVdF–HFP membrane achieved the best balance of activity and stability performances. It delivered an oxidation current density of 564.88 μA cm−2 and a reaction order of 0.809 with respect to ethylene glycol concentration. After 100 repeated oxidation cycles, this Ni5Cu5/PVdF–HFP film retained 87.95% of its initial activity. A decreased Rct value [392.6 Ohm cm2] was also measured to ascertain the rapid charge transfer kinetics at these metallic polymeric surfaces. This surprising activity of 50 wt.%Ni:50 wt.% Cu composition was explained by the synergistic interplay between the distribution of sufficient Ni(II)/Ni(III) redox sites within the membrane surface and the improved conducting environment by providing copper-based species in appreciable contents. They efficiently weakened the adsorption tendency of poisoning intermediates to the catalytic surface leaving highly active areas for prolonged oxidation reaction. On the other hand, the insufficient and irregular arrangements of these metallic sites for films with Ni:Cu weight ratios of 9:1 and 3:7 might result in lowered steady state oxidation current densities and increased Rct values during their electrochemical activity investigations. This preparation strategy of modified polymeric membranes significantly presents a cost-effective alternative to noble metals containing nanocatalysts, extending their applicability to other valuable electrocatalytic reactions in biosensing and energy storage fields.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15100959/s1, Figure S1: SEM image (a) and EDX chart (b) of Ni3Cu7/PVdF-HFP membrane at different surface points; Figure S2: XRD chart of Ni3Cu7/PVdF-HFP membrane; Figure S3: Subsequent cyclic voltammograms of (a) Ni9Cu1/PVdF-HFP, (b) Ni7Cu3/PVdF-HFP, (c) Ni5Cu5/PVdF-HFP and (d) Ni3Cu7/PVdF-HFP nanocatalysts in (1.0 M EG + 0.5 M NaOH) solution for 100 cycles at 10 mV s−1.

Author Contributions

Conceptualization, R.M.A.H.; methodology, A.Y., R.M.A.H., and I.M.M.; software, I.M.M.; validation, R.M.A.H.; formal analysis, R.M.A.H. and A.Y.; investigation, I.M.M.; resources, A.Y. and I.M.M.; data curation, R.M.A.H. and A.Y.; writing—original draft preparation, R.M.A.H.; writing—review and editing, I.M.M. and A.A.; visualization, A.Y.; supervision; R.M.A.H.; project administration, I.M.M.; funding acquisition, A.Y. and I.M.M. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the funding of the Deanship of Graduate Studies and Scientific Research, Jazan University, Saudi Arabia, through project number: (RG24-M038).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors gratefully acknowledge the funding of the Deanship of Graduate Studies and Scientific Research, Jazan University, Saudi Arabia, through project number: (RG24-M038).

Conflicts of Interest

The authors have no relevant financial or non-financial interests to disclose.

References

  1. Antolini, E.; Gonzalez, E.R. Alkaline direct alcohol fuel cells. J. Power Sources 2010, 195, 3431–3450. [Google Scholar] [CrossRef]
  2. Govindarajan, D.; Murugadoss, G.; Kirubaharan, K.; Manavalan, R.K.; Manibalan, G.; Shaikh, J.; Etesami, M.; Kheawhom, S. Improved electrochemical supercapacitive properties of CuO-Ni(OH)2 nanocomposites by eco-friendly low-temperature synthesis. J. Alloys Compd. 2023, 942, 169130. [Google Scholar] [CrossRef]
  3. Meng, Z.-Y.; Li, Y.; Wang, S.-X.; Qiu, Z.-M.; Suen, N.-T.; Guo, X.-T.; Pi, Y.-C.; Pang, H. Boosting hydrogen oxidation performance of bimetallic telluride by electrochemical surface engineering. Rare Met. 2025, 44, 4669–4678. [Google Scholar] [CrossRef]
  4. Ren, X.; Qiu, L.; Li, M.; Tian, F.; He, L.; Guo, X.; Wu, F.; Liu, Y.; Sheng, J.; Yang, W.; et al. Hierarchically interfacial sulfide/phosphides achieve industry-level water electrolyzer in alkaline conditions at 3 A cm−2. Appl. Catal. B Environ. Energy 2025, 378, 125622. [Google Scholar] [CrossRef]
  5. Yan, X.; Yu, S.; Tang, Y.; Sun, D.; Xu, L.; Xue, C. Triangular AgAu/Pt core–shell nanoframes with a dendritic Pt shell and enhanced electrocatalytic performance toward the methanol oxidation reaction. Nanoscale 2018, 10, 2231–2235. [Google Scholar] [CrossRef]
  6. Qiu, X.; Li, T.; Deng, S.; Cen, K.; Xu, L.; Tang, Y. A general strategy for the synthesis of PtM (M=Fe, Co, Ni) decorated three-dimensional hollow graphene nanospheres for efficient methanol electrooxidation. Chem. A Eur. J. 2018, 24, 1246–1252. [Google Scholar] [CrossRef] [PubMed]
  7. Cermenek, B.; Ranninger, J.; Feketeföldi, B.; Letofsky-Papst, I.; Kienzl, N.; Bitschnau, B.; Hacker, V. Novel highly active carbon supported ternary PdNiBi nanoparticles as anode catalyst for the alkaline direct ethanol fuel cell. Nano Res. 2019, 12, 683–693. [Google Scholar] [CrossRef]
  8. Yang, Y.; Ma, Y.; Lu, C.; Li, S.; Zhu, M. Molten salt technique for the synthesis of carbon-based materials for supercapacitors. Green Chem. 2023, 25, 10209–10234. [Google Scholar] [CrossRef]
  9. Zikhali, M.; Matthews, T.; Selepe, C.T.; Adegoke, K.A.; Mugadza, K.; Gwebu, S.S.; Maxakato, N.W. Electrocatalytic performance of Pd-based electro-catalysts supported on CNTs for isopropanol and ethanol electro-oxidation in alkaline medium. Mater. Today Chem. 2023, 30, 101484. [Google Scholar] [CrossRef]
  10. Ren, D.; Deng, Y.; Handoko, A.D.; Chen, C.S.; Malkhandi, S.; Yeo, B.S. Selective electrochemical reduction of carbon dioxide to ethylene and ethanol on copper(I) oxide catalysts. ACS Catal. 2015, 5, 2814–2821. [Google Scholar] [CrossRef]
  11. An, L.; Chen, R. Recent progress in alkaline direct ethylene glycol fuel cells for sustainable energy production. J. Power Sources 2016, 329, 484–501. [Google Scholar] [CrossRef]
  12. Ma, S.; Sadakiyo, M.; Luo, R.; Heima, M.; Yamauchi, M.; Kenis, P.J.A. One-step electrosynthesis of ethylene and ethanol from CO2 in an alkaline electrolyzer. J. Power Sources 2016, 301, 219–228. [Google Scholar] [CrossRef]
  13. Hu, T.; Wang, Y.; Xiao, H.; Chen, W.; Zhao, M.; Jia, J. Shape-control of super-branched Pd–Cu alloys with enhanced electrocatalytic performance for ethylene glycol oxidation. Chem. Commun. 2018, 54, 13363–13366. [Google Scholar] [CrossRef]
  14. Zheng, X.; Wang, G.; Zhao, Y.; Wu, L.; Wang, Y.; Song, Y.; Tian, P.; Wang, X. Controllable morphology of Pd-loaded potassium tantalates with high catalytic performance for ethylene glycol electrooxidation. Electrochim. Acta 2021, 376, 137978. [Google Scholar] [CrossRef]
  15. Bi, P.; Hong, W.; Shang, C.; Wang, J.; Wang, E. Bimetallic PdRu nanosponges with a tunable composition for ethylene glycol oxidation. RSC Adv. 2016, 6, 12486–12490. [Google Scholar] [CrossRef]
  16. Fu, X.; Wan, C.; Huang, Y.; Duan, X. Noble metal based electrocatalysts for alcohol oxidation reactions in alkaline media. Adv. Funct. Mater. 2022, 32, 2106401. [Google Scholar] [CrossRef]
  17. Barakat, N.A.M.; Saadawi, A.; Madih, K.; Hefny, R. Multiphase nickel-based carbon nanofibers for enhanced electrooxidation of urea and glycerol. Next Mater. 2025, 9, 101036. [Google Scholar] [CrossRef]
  18. Chowdhury, M.A.S.; Islam, M.M.; Jamal, M. Green synthesis of nickel oxide nanoparticles using Allium cepa stalks and investigation of their antibacterial activity. Results Chem. 2025, 16, 102328. [Google Scholar] [CrossRef]
  19. Raveendran, A.; Chandran, M.; Wabaidur, S.M.; Islam, M.A.; Dhanusuraman, R.; Ponnusamy, V.K. Facile electrochemical fabrication of nickel-coated polydiphenylamine (Ni/PDPA) nanocomposite material as efficient anode catalyst for direct alcohol fuel cell application. Fuel 2022, 324 Part B, 124424. [Google Scholar] [CrossRef]
  20. Khatun, E.; Ramona, G.; Sobota, L.; Bondue, C.J.; Tschulik, K. Au-Ni rotating ring-disk electrode (RRDE) study: A new and rapid screening technique for quantitative detecting aldehyde during electrochemical alcohol oxidation reaction. Electrochim. Acta 2025, 519, 145803. [Google Scholar] [CrossRef]
  21. Guo, C.; Ren, X.; Guo, C.; Akram, N.; Mansha, M.S.; Ahmad, A.; Khawaja, S.; Wang, J. Fabrication of CdS/Ni-MOF bifunctional catalyst for the photocatalytic hydrogen evolution coupled with benzyl alcohol’s oxidation. J. Photochem. Photobiol. A Chem. 2025, 462, 116181. [Google Scholar] [CrossRef]
  22. Ma, M.; Yang, H.; Li, Q.; Ma, J.; Wu, W.; Zhang, Y.; Cao, X.; Huang, Y. Graphite encapsulated high active nickel-molybdenum/nickel oxide porous nanosheet composites as enhanced oxygen evolution reaction electrocatalysts. J. Alloys Compd. 2025, 1017, 179073. [Google Scholar] [CrossRef]
  23. Bera, A.; Mondal, A.; Maity, P.; Das, A.; Hati, G.; Khatua, B.B. The construction of a flexible solid-state asymmetric supercapacitor device using nickel-copper oxide nanoplates and carbon nanosphere serving as positive and negative electrode, respectively with excellent electrochemical capabilities. Mater. Sci. Eng. B 2024, 305, 117404. [Google Scholar] [CrossRef]
  24. Subedi, B.; Khatoon, N.; Gaire, M.; Majed, A.; He, J.; Zhang, X.; Chrisey, D.B. Rapid fabrication of binder free nickel cobalt oxide electrodes with dendritic nanostructure for electrochemical energy storage applications. Colloids Surf. A Physicochem. Eng. Asp. 2024, 695, 134251. [Google Scholar] [CrossRef]
  25. Nagajyothi, P.C.; Muniratnam, K.; Shim, J. Nickel ferrite as an efficient electrocatalyst for alcohol oxidation reactions. Mater. Lett. 2025, 381, 137752. [Google Scholar] [CrossRef]
  26. Wan, J.; Mu, X.; Jin, Y.; Zhu, J.; Xiong, Y.; Li, T.; Li, R. Nitrogen-doped nickel-molybdenum oxide as a highly efficient electrocatalyst for benzyl alcohol oxidation. Green Chem. 2022, 24, 4870–4876. [Google Scholar] [CrossRef]
  27. Hedayatinasab, F.; Ghaffarinejad, A. Enhanced ethanol oxidation using nickel-tungsten alloy: Durable catalysts for direct ethanol fuel cells. J. Environ. Chem. Eng. 2025, 13, 115228. [Google Scholar] [CrossRef]
  28. Massaneiro, J.; Valério, T.L.; Pellosi, D.S.; da Silva, B.J.G.; Vidotti, M. Electrocatalytic oxidation of glycerol performed by nickel/cobalt alloys: Adding value to a common subproduct of chemical industry. Electrochim. Acta 2024, 506, 145013. [Google Scholar] [CrossRef]
  29. ElSayed, N.G.; Farghali, A.A.; El Rouby, W.M.A.; Hmamm, M.F.M. Silver decorated nickel oxide nanoflake/carbon nanotube nanocomposite as an efficient electrocatalyst for ethanol oxidation. Nanoscale Adv. 2024, 6, 5133–5144. [Google Scholar] [CrossRef] [PubMed]
  30. Abdel Hameed, R.M.; Abutaleb, A.; Zouli, N.; Yousef, A. Facile synthesis of electrospun transition metallic nanofibrous mats with outstanding activity for ethylene glycol electro-oxidation in alkaline solution. Mol. Catal. 2022, 522, 112186. [Google Scholar] [CrossRef]
  31. Suárez-Barajas, A.; Ramos-Castillo, C.M.; Olivas, A.; Guerra-Balcázar, M.; Álvarez-Contreras, L.; Arjona, N. Oxygen vacancy-enriched NiCo2O4 spinels/N-doped carbon nanotubes-graphene composites for the ethylene glycol electro-oxidation. Fuel 2024, 360, 130371. [Google Scholar] [CrossRef]
  32. Su, X.; Ren, J.; Meng, X.; Ren, X.; Tang, F. A novel platform for enhanced biosensing based on the synergy effects of electrospun polymer nanofibers and graphene oxides. Analyst 2013, 138, 1459–1466. [Google Scholar] [CrossRef]
  33. Kumar, G.G.; Kim, A.R.; Nahm, K.S.; Yoo, D.J. High proton conductivity and low fuel crossover of polyvinylidene fluoride–hexafluoro propylene–silica sulfuric acid composite membranes for direct methanol fuel cells. Curr. Appl. Phys. 2011, 11, 896–902. [Google Scholar] [CrossRef]
  34. Ali Shah, A.U.H.; Yasmeen, N.; Rahman, G.; Bilal, S. High electrocatalytic behavior of Ni impregnated conducting polymer coated platinum and graphite electrodes for electrooxidation of methanol. Electrochim. Acta 2017, 224, 468–474. [Google Scholar] [CrossRef]
  35. Kazim, H.; Sabri, M.; Al-Othman, A.; Tawalbeh, M. Functionalized conducting polymers in photocatalysis and opportunities for artificial intelligence applications. Nano-Struct. Nano-Objects 2024, 40, 101371. [Google Scholar] [CrossRef]
  36. Bharatish, A.; Kumar, A.; Siddhanth, K.S.; Manikant, V.; Jagdish, P.; Sharma, A.; Solaiachari, S. On optimizing wettability, surface roughness and swelling behaviour of laser-polished 3D-printed PETG polymer for bio-medical implants. Polymer 2025, 330, 128482. [Google Scholar] [CrossRef]
  37. Panthi, G.; Barakat, N.A.M.; Khalil, K.A.; Yousef, A.; Jeon, K.-S.; Kim, H.Y. Encapsulation of CoS nanoparticles in PAN electrospun nanofibers: Effective and reusable catalyst for ammonia borane hydrolysis and dyes photodegradation. Ceram. Int. 2013, 39, 1469–1476. [Google Scholar] [CrossRef]
  38. Mandal, B.P.; Vasundhara, K.; Abdelhamid, E.; Lawes, G.; Salunke, H.G.; Tyagi, A.K. Improvement of magnetodielectric coupling by surface functionalization of nickel nanoparticles in Ni and polyvinylidene fluoride nanohybrids. J. Phys. Chem. C 2014, 118, 20819–20825. [Google Scholar] [CrossRef]
  39. Kumar, R.V.; Mastai, Y.; Diamant, Y.; Gedanken, A. Sonochemical synthesis of amorphous Cu and nanocrystalline Cu2O embedded in a polyaniline matrix. J. Mater. Chem. 2001, 11, 1209–1213. [Google Scholar] [CrossRef]
  40. Nagashree, K.L.; Ahmed, M.F. Electrocatalytic oxidation of methanol on Ni modified polyaniline electrode in alkaline medium. J. Solid State Electrochem. 2010, 14, 2307–2320. [Google Scholar] [CrossRef]
  41. Salman, M.; Aslam, M.; Mustafa, Z.; Ahmad, M.; Shehzad, A. Flexible polypyrrole/polyvinyl alcohol/titanium dioxide/graphene quantum dots nanocomposite films for energy storage devices. Diam. Relat. Mater. 2025, 152, 111996. [Google Scholar] [CrossRef]
  42. Lei, Z.-L.; Wan, L.; Lü, Q.-F. Carbon nanotubes modified V-Ti3C2Tx/poly(3,4-ethylenedioxythiophene) composite as a high-performance electrode for supercapacitor. Diam. Relat. Mater. 2024, 150, 111696. [Google Scholar] [CrossRef]
  43. Sundararajan, P.B.; Muthukrishnan, F.P.; Murugesan, A.; Chuang, H.-C.; Pandiyarajan, S. Nitrogen-sulphur dual doped reduced graphene oxide-polydiphenylamine supported PdSn nanocomposite: A bimetallic electrocatalyst for methanol electro-oxidation in alkaline solution. Mater. Sci. Eng. B 2024, 304, 117363. [Google Scholar] [CrossRef]
  44. Herranz, D.; Escudero-Cid, R.; Montiel, M.; Palacio, C.; Fatás, E.; Ocón, P. Poly(vinyl alcohol) and poly(benzimidazole) blend membranes for high performance alkaline direct ethanol fuel cells. Renew. Energy 2018, 127, 883–895. [Google Scholar] [CrossRef]
  45. Wring, S.A.; Hart, J.P. Chemically modified, carbon-based electrodes and their application as electrochemical sensors for the analysis of biologically important compounds. A review. Analyst 1992, 117, 1215–1229. [Google Scholar] [CrossRef]
  46. Nguyen, N.S.; Das, G.; Yoon, H.H. Nickel/cobalt-decorated 3D graphene nanocomposite electrode for enhanced electrochemical detection of urea. Biosens. Bioelectron. 2016, 77, 372–377. [Google Scholar] [CrossRef]
  47. Li, M.; Sun, J.; Chen, G.; Yao, S.; Cong, B. Construction double electric field of sulphur vacancies as medium ZnS/Bi2S3-PVDF self-supported recoverable piezoelectric film photocatalyst for enhanced photocatalytic performance. Appl. Catal. B Environ. 2022, 301, 120792. [Google Scholar] [CrossRef]
  48. Su, P.; Zhang, D.; Zhu, M.; Liang, T.; Yang, N.; Zhao, H.; Zhang, D.; Liu, J.; Cai, P.; Pu, X. Re-usable Cd0.9Zn0.1S-ZnO-C/PVDF piezo-photocatalytic film with exceptional hydrogen evolution capability triggered by the synergetic advantages of piezoelectricity and S-Scheme heterojunction. J. Energy Chem. 2024, 96, 164–176. [Google Scholar] [CrossRef]
  49. Huang, Q.; Zhu, F.; Xiao, F.; Zhang, G.; Hou, H.; Bi, J.; Yan, S.; Hao, H. Construction of PVDF-HKUST-1/BiVO4 hybrid electrospun flexible fiber membrane for enhanced piezo-photocatalytic reduction performance of Cr(VI). Colloids Surf. A Physicochem. Eng. Asp. 2024, 693, 134046. [Google Scholar] [CrossRef]
  50. Kumar, G.G. Irradiated PVdF-HFP-montmorillonite composite membranes for the application of direct ethanol fuel cells. J. Mater. Chem. 2011, 21, 17382–17391. [Google Scholar] [CrossRef]
  51. Kumar, G.G.; Nahm, K.S.; Elizabeth, R.N. Electro chemical properties of porous PVdF-HFP membranes prepared with different nonsolvents. J. Membr. Sci. 2008, 325, 117–124. [Google Scholar] [CrossRef]
  52. Senthilkumar, N.; Babu, K.J.; Kumar, G.G.; Kim, A.R.; Yoo, D.J. Flexible electrospun PVdF-HFP/Ni/Co membranes for efficient and highly selective enzyme free glucose detection. Ind. Eng. Chem. Res. 2014, 53, 10347–10357. [Google Scholar] [CrossRef]
  53. Kumar, T.R.; Babu, K.J.; Yoo, D.J.; Kim, A.R.; Kumar, G.G. Binder free and free-standing electrospun membrane architecture for sensitive and selective non-enzymatic glucose sensors. RSC Adv. 2015, 5, 41457–41467. [Google Scholar] [CrossRef]
  54. Ren, H.; Labidi, A.; Sial, A.; Miao, Z.; Zhao, Y.; Feng, X.; Luo, J.; Wang, C. Fluid-induced piezoelectric field enhanced photocatalytic antibiotic removal over nitrogen-doped carbon quantum dots/Bi2WO6-polyvinylidene fluoride-co-hexafluoropropylene membrane in aqueous environments. J. Colloid Interface Sci. 2025, 691, 137412. [Google Scholar] [CrossRef]
  55. Li, J.; Du, L.; Guo, S.; Chang, J.; Wu, D.; Jiang, K.; Gao, Z. Molybdenum iron carbide-copper hybrid as efficient electrooxidation catalyst for oxygen evolution reaction and synthesis of cinnamaldehyde/benzalacetone. J. Colloid Interface Sci. 2024, 673, 616–627. [Google Scholar] [CrossRef]
  56. Castro, A.S.; de Azevedo, L.S.; Bruni, A.T.; de Menezes, M.M.T.; Dockal, E.R.; de Oliveira, O.V.; Ferreira, B.; de Oliveira, M.F. Cocaine electrooxidation behavior, mechanism, and kinetics on a carbon paste electrode chemically modified with a cobalt or copper Schiff base complex. Forensic Chem. 2021, 26, 100363. [Google Scholar] [CrossRef]
  57. Joshi, K.K.; Chauhan, S.V.; Pataniya, P.M.; Sumesh, C.K. Efficient hybrid water splitting and direct electrooxidation of organic dye from wastewater using copper cobalt sulphide nanosheets. Int. J. Hydrogen Energy 2024, 58, 1562–1575. [Google Scholar] [CrossRef]
  58. Chen, C.; Jin, L.; Shang, H.; Song, T.; Gao, F.; Zhang, Y.; Wang, C.; Wang, C.; Du, Y. Monodispersed bimetallic platinum-copper alloy nanospheres as efficient catalysts for ethylene glycol electrooxidation. J. Colloid Interface Sci. 2019, 551, 81–88. [Google Scholar] [CrossRef]
  59. Meng, K.; Fan, Z.; Wen, H.; Hu, Y.; Zhang, W.; Chen, Z. Construction of a Cu3+–OH–Pt interface for enhancing glycerol electrooxidation coupled with hydrogen evolution. Green Chem. 2025, 27, 5376–5387. [Google Scholar] [CrossRef]
  60. Sumbal, S.; Łuczak, J.; Hassan, A.; Lieder, M. Role of Cu and Ni in dual sites layered electrocatalysts for ammonia electrooxidation. Appl. Catal. A Gen. 2025, 696, 120181. [Google Scholar] [CrossRef]
  61. Han, M.; Wang, N.; Zhang, B.; Xia, Y.; Li, J.; Han, J.; Yao, K.; Gao, C.; He, C.; Liu, Y.; et al. High-valent nickel promoted by atomically embedded copper for efficient water oxidation. ACS Catal. 2020, 10, 9725–9734. [Google Scholar] [CrossRef]
  62. He, Q.; Yao, K.; Wang, X.; Xia, X.; Leng, S.; Li, F. Room-temperature and solution-processable Cu-doped nickel oxide nanoparticles for efficient hole-transport layers of flexible large-area perovskite solar cells. ACS Appl. Mater. Interfaces 2017, 9, 41887–41897. [Google Scholar] [CrossRef] [PubMed]
  63. Cao, M.; Cao, H.; Meng, W.; Wang, Q.; Bi, Y.; Liang, X.; Yang, H.; Zhang, L.; Lang, M.-F.; Sun, J. Nickel-copper oxide nanoflowers for highly efficient glucose electrooxidation. Int. J. Hydrogen Energy 2021, 46, 28527–28536. [Google Scholar] [CrossRef]
  64. Ren, X.; Zhang, Q.; Tong, Y.; Zhou, G.; Lin, C.; Zhao, Y.; Chen, P. Nickel-copper alloying arrays realizing efficient co-electrosynthesis of adipic acid and hydrogen. J. Energy Chem. 2025, 101, 7–15. [Google Scholar] [CrossRef]
  65. Chemchoub, S.; El Attar, A.; Oularbi, L.; Younssi, S.A.; Bentiss, F.; Jama, C.; El Rhazi, M. Electrosynthesis of eco-friendly electrocatalyst based nickel-copper bimetallic nanoparticles supported on poly-phenylenediamine with highest current density and early ethanol oxidation onset potential. Int. J. Hydrogen Energy 2022, 47, 39081–39096. [Google Scholar] [CrossRef]
  66. Gibson, P.; Schreuder-Gibson, H.; Rivin, D. Transport properties of porous membranes based on electrospun nanofibers. Colloids Surf. A Physicochem. Eng. Asp. 2001, 187–188, 469–481. [Google Scholar] [CrossRef]
  67. Kumar, G.G.; Shin, J.; Nho, Y.-C.; Hwang, I.S.; Fei, G.; Kim, A.R.; Nahm, K.S. Irradiated PVdF-HFP–tin oxide composite membranes for the applications of direct methanol fuel cells. J. Membr. Sci. 2010, 350, 92–100. [Google Scholar] [CrossRef]
  68. Chen, Y.; Shi, Y.; Wang, Y. Preparation of hollow poly(vinylidene fluoride) capsules containing nickel catalyst for hydrogen storage and production. Int. J. Energy Res. 2015, 39, 634–642. [Google Scholar] [CrossRef]
  69. Babu, K.J.; Senthilkumar, N.; Kim, A.R.; Kumar, G.G. Freestanding and binder free PVdF-HFP/Ni-Co nanofiber membrane as a versatile platform for the electrocatalytic oxidation and non-enzymatic detection of urea. Sens. Actuators B Chem. 2017, 241, 541–551. [Google Scholar] [CrossRef]
  70. Nath, A.K.; Kumar, A. Swift heavy ion irradiation induced enhancement in electrochemical properties of ionic liquid based PVdF-HFP-layered silicate nanocomposite electrolyte membranes. J. Membr. Sci. 2014, 453, 192–201. [Google Scholar] [CrossRef]
  71. Ozay, O.; Aktas, N.; Inger, E.; Sahiner, N. Hydrogel assisted nickel nanoparticle synthesis and their use in hydrogen production from sodium boron hydride. Int. J. Hydrogen Energy 2011, 36, 1998–2006. [Google Scholar] [CrossRef]
  72. Abutaleb, A. CuNi alloy NPs anchored on electrospun PVDF-HFP NFs catalyst for H2 production from sodium borohydride. Polymers 2023, 15, 474. [Google Scholar] [CrossRef] [PubMed]
  73. Hu, C.; Lv, C.; Zeng, N.; Liu, A.; Liu, Y.; Hu, L.; Li, P.; Yao, Y.; Cai, J.; Tang, T. Recent advances in Ni-based electrocatalysts for hydrogen evolution reaction. Energy Technol. 2023, 11, 2201048. [Google Scholar] [CrossRef]
  74. Li, Z.; Zheng, Y.; Guo, H.; Cheng, X.; Huang, Y.; Liu, C.; Zang, J.; Dong, L. Ni-based electrocatalysts for urea oxidation reaction: Mechanistic insights and recent advancements. J. Alloys Compd. 2024, 1008, 176591. [Google Scholar] [CrossRef]
  75. Abbas, M.; Abdel Hameed, R.M.; Al-Enizi, A.M.; Thamer, B.M.; Yousef, A.; El-Newehy, M.H. Decorated carbon nanofibers with mixed nickel–manganese carbides for methanol electro-oxidation in alkaline solution. Int. J. Hydrogen Energy 2021, 46, 6494–6512. [Google Scholar] [CrossRef]
  76. Wang, W.; Li, R.; Hua, X.; Zhang, R. Methanol electrooxidation on glassy carbon electrode modified with bimetallic Ni(II)Co(II) salen complexes encapsulated in mesoporous zeolite A. Electrochim. Acta 2015, 163, 48–56. [Google Scholar] [CrossRef]
  77. Rostami, T.; Jafarian, M.; Miandari, S.; Mahjani, M.G.; Gobal, F. Synergistic effect of cobalt and copper on a nickel-based modified graphite electrode during methanol electro-oxidation in NaOH solution. Chin. J. Catal. 2015, 36, 1867–1874. [Google Scholar] [CrossRef]
  78. Rahmani, K.; Habibi, B. NiCo alloy nanoparticles electrodeposited on an electrochemically reduced nitrogen-doped graphene oxide/carbon-ceramic electrode: A low cost electrocatalyst towards methanol and ethanol oxidation. RSC Adv. 2019, 9, 34050–34064. [Google Scholar] [CrossRef]
  79. Urbańczyk, E.; Wala, M.; Blacha-Grzechnik, A.; Stolarczyk, A.; Maciej, A.; Simka, W. Electrocatalytic methanol oxidation using Ni–Co–graphene composite electrodes. Int. J. Hydrogen Energy 2021, 46, 33272–33286. [Google Scholar] [CrossRef]
  80. Bard, A.J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and Applications, 2nd ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2001. [Google Scholar]
  81. Zhang, J.; Tse, Y.-H.; Pietro, W.J.; Lever, A.B.P. Electrocatalytic activity of N, N′, N′′, N′′′-tetramethyl-tetra-3,4-pyridoporphyrazinocobalt(II) adsorbed on a graphite electrode towards the oxidation of hydrazine and hydroxylamine. J. Electroanal. Chem. 1996, 406, 203–211. [Google Scholar] [CrossRef]
  82. Dantas, L.M.F.; dos Reis, A.P.; Tanaka, S.M.C.N.; Zagal, J.H.; Chen, Y.-Y.; Tanaka, A.A. Electrocatalytic oxidation of hydrazine in alkaline media promoted by iron tetrapyridinoporphyrazine adsorbed on graphite surface. J. Braz. Chem. Soc. 2008, 19, 720–726. [Google Scholar] [CrossRef]
  83. Abdel Hameed, R.M.; Medany, S.S. Influence of support material on the electrocatalytic activity of nickel oxide nanoparticles for urea electro-oxidation reaction. J. Colloid Interface Sci. 2018, 513, 536–548. [Google Scholar] [CrossRef]
  84. Afify, D.G.; Abdel Hameed, R.M.; Mohamed, A.M.; Ghayad, I.M. Outstanding activity and durability of supported NiCo2O4 nanoparticles for direct ethanol fuel cells. Appl. Organomet. Chem. 2025, 39, e7802. [Google Scholar] [CrossRef]
  85. Rodriguez, J.R.; Fuentes-Moyado, S.; Zepeda, T.A.; Díaz de León, J.N.; Cruz-Reyes, J.; Oropeza-Guzman, M.T.; Berhault, G.; Alonso-Núňez, G. Methanol electro-oxidation with alloy nanoparticles of Pt10−x–Fex supported on CNTs. Fuel 2016, 182, 1–7. [Google Scholar] [CrossRef]
  86. Dong, K.; Zhang, Y.; Yang, X.; Yuan, Q. Au enables PdAu aerogel efficient catalysts for oxygen reduction and C2 alcohol oxidation. Int. J. Hydrogen Energy 2024, 78, 443–451. [Google Scholar] [CrossRef]
  87. Lao, X.; Li, Z.; Yang, L.; Zhang, B.; Ye, W.; Fu, A.; Guo, P. Monodispersed ultrathin twisty PdBi alloys nanowires assemblies with tensile strain enhance C2+ alcohols electrooxidation. J. Energy Chem. 2023, 79, 279–290. [Google Scholar] [CrossRef]
  88. El-Shafei, A.A. Electrocatalytic oxidation of methanol at a nickel hydroxide/glassy carbon modified electrode in alkaline medium. J. Electroanal. Chem. 1999, 471, 89–95. [Google Scholar] [CrossRef]
  89. Sun, H.; Ye, Y.; Liu, J.; Tian, Z.; Cai, Y.; Li, P.; Liang, C. Pure Ni nanocrystallines anchored on rGO present ultrahigh electrocatalytic activity and stability in methanol oxidation. Chem. Commun. 2018, 54, 1563–1566. [Google Scholar] [CrossRef]
  90. Afify, D.G.; Abdel Hameed, R.M.; Ghayad, I.M. Binary NiCu oxide nanoparticles onto graphite as promoting nanocatalysts for ethanol electro-oxidation process. Appl. Organomet. Chem. 2024, 38, e7418. [Google Scholar] [CrossRef]
  91. Laviron, E. General expression of the linear potential sweep voltammogram in the case of diffusionless electrochemical systems. J. Electroanal. Chem. Interfacial Electrochem. 1979, 101, 19–28. [Google Scholar] [CrossRef]
  92. Abdel Hameed, R.M.; Al-Enizi, A.M.; Karim, A.; Yousef, A. Free-standing and binder-free nickel polymeric nanofiber membrane for electrocatalytic oxidation of ethanol in alkaline solution. J. Mater. Res. Technol. 2022, 21, 4591–4606. [Google Scholar] [CrossRef]
  93. Maritan, A.; Toigo, F. On skewed ARC plots of impedance of electrodes with an irreversible electrode process. Electrochim. Acta 1990, 35, 141–145. [Google Scholar] [CrossRef]
  94. Mao, Y.-H.; Chen, C.-Y.; Fu, J.-X.; Lai, T.-Y.; Lu, F.-H.; Tsai, Y.-C. Electrodeposition of nickel-copper on titanium nitride for methanol electrooxidation. Surf. Coat. Technol. 2018, 350, 949–953. [Google Scholar] [CrossRef]
  95. Shahrokhian, S.; Rezaee, S. Vertically standing Cu2O nanosheets promoted flower-like PtPd nanostructures supported on reduced graphene oxide for methanol electro-oxidation. Electrochim. Acta 2018, 259, 36–47. [Google Scholar] [CrossRef]
  96. Askari, M.B.; Salarizadeh, P. Superior catalytic performance of NiCo2O4 nanorods loaded rGO towards methanol electro-oxidation and hydrogen evolution reaction. J. Mol. Liq. 2019, 291, 111306. [Google Scholar] [CrossRef]
  97. Sheikhi, S.; Jalali, F. Hierarchical NiCo2O4/CuO-C nanocomposite derived from copper-based metal organic framework and Ni/Co hydroxides: Excellent electrocatalytic activity towards methanol oxidation. J. Alloys Compd. 2022, 907, 164510. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a) PVdF–HFP, (b) Ni9Cu1/PVdF–HFP, (c) Ni7Cu3/PVdF–HFP, (d) Ni5Cu5/PVdF–HFP and (e) Ni3Cu7/PVdF–HFP films [72].
Figure 1. SEM images of (a) PVdF–HFP, (b) Ni9Cu1/PVdF–HFP, (c) Ni7Cu3/PVdF–HFP, (d) Ni5Cu5/PVdF–HFP and (e) Ni3Cu7/PVdF–HFP films [72].
Catalysts 15 00959 g001
Figure 2. FTIR charts of PVdF–HFP and its metallic NixCu1–x films [72].
Figure 2. FTIR charts of PVdF–HFP and its metallic NixCu1–x films [72].
Catalysts 15 00959 g002
Figure 3. Cyclic voltammetric curves of (a) Ni5Cu5/PVdF–HFP, (b) Ni9Cu1/PVdF–HFP, (c) Ni7Cu3/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP membranes in 0.5 M NaOH solution at 5 mV s−1.
Figure 3. Cyclic voltammetric curves of (a) Ni5Cu5/PVdF–HFP, (b) Ni9Cu1/PVdF–HFP, (c) Ni7Cu3/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP membranes in 0.5 M NaOH solution at 5 mV s−1.
Catalysts 15 00959 g003
Figure 4. (a) Variation of the cyclic voltammetric curves of Ni5Cu5/PVdF–HFP electrode in 0.5 M NaOH solution using scan rates of 5, 10, 20, 30, 40, 50, 75, 100 and 200 mV s−1. The respective potentials and current densities of Ni(II)/Ni(III) redox couple were drawn against υ and υ0.5 in sections (bd).
Figure 4. (a) Variation of the cyclic voltammetric curves of Ni5Cu5/PVdF–HFP electrode in 0.5 M NaOH solution using scan rates of 5, 10, 20, 30, 40, 50, 75, 100 and 200 mV s−1. The respective potentials and current densities of Ni(II)/Ni(III) redox couple were drawn against υ and υ0.5 in sections (bd).
Catalysts 15 00959 g004
Figure 5. Cyclic voltammograms of (a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP films in (1.0 M EG + 0.5 M NaOH) solution at 10 mV s−1. The measured alcohol oxidation current densities at these electrodes were followed by their copper content in section (e).
Figure 5. Cyclic voltammograms of (a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP films in (1.0 M EG + 0.5 M NaOH) solution at 10 mV s−1. The measured alcohol oxidation current densities at these electrodes were followed by their copper content in section (e).
Catalysts 15 00959 g005
Figure 6. (a) Cyclic voltammograms of Ni5Cu5/PVdF–HFP nanomaterial in 0.5 M NaOH solution that contained altered alcohol concentrations as 0.1, 0.3, 0.4, 0.5, 0.7 and 1.0 M at 10 mV s−1. The alcohol oxidation current density was measured at different concentrations in section (b), while log (I)–log (C) relation was illustrated in section (c).
Figure 6. (a) Cyclic voltammograms of Ni5Cu5/PVdF–HFP nanomaterial in 0.5 M NaOH solution that contained altered alcohol concentrations as 0.1, 0.3, 0.4, 0.5, 0.7 and 1.0 M at 10 mV s−1. The alcohol oxidation current density was measured at different concentrations in section (b), while log (I)–log (C) relation was illustrated in section (c).
Catalysts 15 00959 g006
Figure 7. (a) CA curves of different NiCu/PVdF–HFP membranes in 0.5 M NaOH solution and 1.0 M EG at 600 mV. Oxidation peak current densities at subsequent repeated cycles were monitored in section (b).
Figure 7. (a) CA curves of different NiCu/PVdF–HFP membranes in 0.5 M NaOH solution and 1.0 M EG at 600 mV. Oxidation peak current densities at subsequent repeated cycles were monitored in section (b).
Catalysts 15 00959 g007
Figure 8. (a) Linear sweep voltammograms of Ni5Cu5/PVdF–HFP film in (1.0 M EG + 0.5 M NaOH) solution at different rates. The corresponding (I–υ0.5), (E–log υ) and (E–log I) dependence relations were derived in sections (bd).
Figure 8. (a) Linear sweep voltammograms of Ni5Cu5/PVdF–HFP film in (1.0 M EG + 0.5 M NaOH) solution at different rates. The corresponding (I–υ0.5), (E–log υ) and (E–log I) dependence relations were derived in sections (bd).
Catalysts 15 00959 g008
Figure 9. Phase shift curves of (a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP nanocatalysts in 0.5 M NaOH solution with including increased concentrations of ethylene glycol up to 1.0 M. The corresponding plots of NiCu/PVdF–HFP films in different Ni:Cu proportions were distinguished in an alkaline solution that contained 0.3 M alcohol in section (e).
Figure 9. Phase shift curves of (a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP nanocatalysts in 0.5 M NaOH solution with including increased concentrations of ethylene glycol up to 1.0 M. The corresponding plots of NiCu/PVdF–HFP films in different Ni:Cu proportions were distinguished in an alkaline solution that contained 0.3 M alcohol in section (e).
Catalysts 15 00959 g009
Figure 10. Bode diagrams of (a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP nanomaterials in 0.5 M NaOH solution in the presence of increments of ethylene glycol content. Section (e) showed a valuable comparison of these curves for varied NiCu/PVdF–HFP films in a basic electrolyte after incorporating 1.0 M alcohol.
Figure 10. Bode diagrams of (a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP nanomaterials in 0.5 M NaOH solution in the presence of increments of ethylene glycol content. Section (e) showed a valuable comparison of these curves for varied NiCu/PVdF–HFP films in a basic electrolyte after incorporating 1.0 M alcohol.
Catalysts 15 00959 g010
Figure 11. Nyquist plots of NiCu/PVdF–HFP membranes that contained Ni:Cu weight ratios of (a) 9:1, (b) 7:3, (c) 5:5 and (d) 3:7 when immersed in a basic solution for alcohol electro-oxidation reaction in varied concentrations. The obtained curves for these electrodes in (1.0 M ethylene glycol + 0.5 M NaOH) solution are contrasted in section (e). The simulated electric circuit components and their connection routes are demonstrated in section (f).
Figure 11. Nyquist plots of NiCu/PVdF–HFP membranes that contained Ni:Cu weight ratios of (a) 9:1, (b) 7:3, (c) 5:5 and (d) 3:7 when immersed in a basic solution for alcohol electro-oxidation reaction in varied concentrations. The obtained curves for these electrodes in (1.0 M ethylene glycol + 0.5 M NaOH) solution are contrasted in section (e). The simulated electric circuit components and their connection routes are demonstrated in section (f).
Catalysts 15 00959 g011
Figure 12. (a) Phase shift, (b) Bode and (c) Nyquist curves of Ni5Cu5/PVdF–HFP membrane in (0.3 M ethylene glycol + 0.5 M NaOH) solution at varied potential values.
Figure 12. (a) Phase shift, (b) Bode and (c) Nyquist curves of Ni5Cu5/PVdF–HFP membrane in (0.3 M ethylene glycol + 0.5 M NaOH) solution at varied potential values.
Catalysts 15 00959 g012
Table 1. The magnitudes of altered electric circuit components after the fitting process of the experimental results to the theoretical ones at different NiCu/PVdF–HFP membranes [(a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP] in NaOH solution with inserting increments of ethylene glycol molecules. These measurements were conducted at 575 mV in the frequency range of 10,000–0.1 Hz.
Table 1. The magnitudes of altered electric circuit components after the fitting process of the experimental results to the theoretical ones at different NiCu/PVdF–HFP membranes [(a) Ni9Cu1/PVdF–HFP, (b) Ni7Cu3/PVdF–HFP, (c) Ni5Cu5/PVdF–HFP and (d) Ni3Cu7/PVdF–HFP] in NaOH solution with inserting increments of ethylene glycol molecules. These measurements were conducted at 575 mV in the frequency range of 10,000–0.1 Hz.
Ethylene Glycol Concentration/MRs/
Ohm
Yo1/
Ohm−1 sa cm−2 × 10−3
a1R1/
Ohm cm2
Yo2/
Ohm−1 sa cm−2 × 10−4
a2R2/
Ohm cm2
(a)
0.1649.01.0000.940370.01.0820.5938478.0
0.2597.01.5190.992384.41.1060.5798028.0
0.3653.01.9900.822380.01.1150.7027661.0
0.4624.82.5530.820374.61.1250.7017220.0
0.6646.83.4950.899363.41.1380.6845774.0
0.8623.14.5840.947366.21.1600.6924752.0
1.0618.45.3890.974368.81.2000.6254230.0
(b)
0.11463.11.4320.825134.31.5260.4704487.0
0.21461.72.1200.852137.71.8150.3403490.0
0.31478.52.9910.852158.52.3660.4572890.0
0.41460.13.8510.866141.43.1150.5931926.0
0.61438.25.7930.875143.44.1280.6441206.0
0.81420.37.4550.882157.75.1900.5641062.0
1.01406.09.5120.883160.46.1070.5001056.0
(c)
0.1412.73.7940.94079.02.5940.4851900.0
0.2409.54.5360.91872.02.9050.4891543.0
0.3407.15.1460.91567.43.2820.5201180.0
0.4404.75.8820.89574.53.8540.510894.5
0.6404.07.1710.89669.74.8180.568522.2
0.8405.48.5260.83659.85.8280.514408.1
1.0410.910.3660.85352.66.6000.489392.6
(d)
0.13906.01.2080.918400.71.3090.6725158.0
0.23912.01.7580.946373.51.3380.5444598.0
0.33960.02.4100.937357.41.3810.6334062.0
0.43960.02.9980.999337.61.4860.6853680.0
0.63916.04.4680.999384.81.5380.6212970.0
0.83872.05.5431.000331.61.7250.5892262.0
1.03878.06.8341.000364.21.8750.6651954.0
Table 2. A comparison of the simulated values of electric circuit constituents during the fitting step of obtained curves at NiCu/PVdF–HFP membranes, using varied Ni:Cu weight % ratios, in (1.0 M ethylene glycol + 0.5 M NaOH) solution.
Table 2. A comparison of the simulated values of electric circuit constituents during the fitting step of obtained curves at NiCu/PVdF–HFP membranes, using varied Ni:Cu weight % ratios, in (1.0 M ethylene glycol + 0.5 M NaOH) solution.
NanomaterialRs/
Ohm
Yo1/
Ohm−1 sa cm−2 × 10−3
a1R1/
Ohm cm2
Yo2/
Ohm−1 sa cm−2 × 10−4
a2R2/
Ohm cm2
Ni9Cu1/PVdF–HFP618.45.3890.974368.81.2000.6254230.0
Ni7Cu3/PVdF–HFP1406.09.5120.883160.46.1070.5001056.0
Ni5Cu5/PVdF–HFP410.910.3660.85352.66.6000.489392.6
Ni3Cu7/PVdF–HFP3878.06.8341.000364.21.8750.6651954.0
Table 3. The fitting results of the experimental data at Ni5Cu5/PVdF–HFP membrane with those of the theoretically predicted values in (0.3 M ethylene glycol + 0.5 M NaOH) solution at different potentials.
Table 3. The fitting results of the experimental data at Ni5Cu5/PVdF–HFP membrane with those of the theoretically predicted values in (0.3 M ethylene glycol + 0.5 M NaOH) solution at different potentials.
E/mVRs/
Ohm
Yo1/
Ohm−1 sa cm−2 × 10−3
a1R1/
Ohm cm2
Yo2/
Ohm−1 sa cm−2 × 10−4
a2R2/
Ohm cm2
350388.43.1990.83559.31.6030.5701976.0
400377.04.3960.87971.12.0540.5081644.0
450378.84.7560.90656.52.3400.5651536.0
500379.65.0020.93258.02.6540.5341318.0
550382.45.0760.93273.53.1300.5111249.0
600418.45.2770.93758.33.4380.5741158.0
650423.25.5570.92274.03.5540.5171088.0
700424.25.6640.92460.83.6270.5331048.0
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

Yousef, A.; Abdel Hameed, R.M.; Maafa, I.M.; Abutaleb, A. Synthesis of NiCu–Polymeric Membranes for Electro-Oxidizing Ethylene Glycol Molecules in Alkaline Medium. Catalysts 2025, 15, 959. https://doi.org/10.3390/catal15100959

AMA Style

Yousef A, Abdel Hameed RM, Maafa IM, Abutaleb A. Synthesis of NiCu–Polymeric Membranes for Electro-Oxidizing Ethylene Glycol Molecules in Alkaline Medium. Catalysts. 2025; 15(10):959. https://doi.org/10.3390/catal15100959

Chicago/Turabian Style

Yousef, Ayman, R. M. Abdel Hameed, Ibrahim M. Maafa, and Ahmed Abutaleb. 2025. "Synthesis of NiCu–Polymeric Membranes for Electro-Oxidizing Ethylene Glycol Molecules in Alkaline Medium" Catalysts 15, no. 10: 959. https://doi.org/10.3390/catal15100959

APA Style

Yousef, A., Abdel Hameed, R. M., Maafa, I. M., & Abutaleb, A. (2025). Synthesis of NiCu–Polymeric Membranes for Electro-Oxidizing Ethylene Glycol Molecules in Alkaline Medium. Catalysts, 15(10), 959. https://doi.org/10.3390/catal15100959

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