Next Article in Journal
Widely Targeted Metabolomics Analysis of the Roots, Stems, Leaves, Flowers, and Fruits of Camellia luteoflora, a Species with an Extremely Small Population
Next Article in Special Issue
The Influence of Thermal Treatment of Activated Carbon on Its Electrochemical, Corrosion, and Adsorption Characteristics
Previous Article in Journal
Artifacts and Anomalies in Raman Spectroscopy: A Review on Origins and Correction Procedures
Previous Article in Special Issue
Electrogeneration and Characterization of Poly(methylene blue) Thin Films on Stainless Steel 316 Electrodes—Effect of pH
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimizing NiFe-Modified Graphite for Enhanced Catalytic Performance in Alkaline Water Electrolysis: Influence of Substrate Geometry and Catalyst Loading

by
Mateusz Kuczyński
1,
Tomasz Mikołajczyk
1,
Bogusław Pierożyński
1,* and
Jakub Karczewski
2
1
Department of Chemistry, Faculty of Agriculture and Forestry, University of Warmia and Mazury in Olsztyn, Łódzki Square 4, 10-727 Olsztyn, Poland
2
Institute of Nanotechnology and Materials Engineering, Faculty of Applied Physics and Mathematics, Gdansk University of Technology, ul. G. Narutowicza 11/12, 80-233 Gdańsk, Poland
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(19), 4755; https://doi.org/10.3390/molecules29194755
Submission received: 7 September 2024 / Revised: 3 October 2024 / Accepted: 5 October 2024 / Published: 8 October 2024
(This article belongs to the Special Issue Exclusive Feature Papers in Electrochemistry)

Abstract

:
The oxygen evolution reaction (OER) and the hydrogen evolution reaction (HER) are critical processes in water splitting, yet achieving efficient performance with minimal overpotential remains a significant challenge. Although NiFe-based catalysts are widely studied, their performance can be further enhanced by optimizing the interaction between the catalyst and the substrate. Here, we present a detailed investigation of NiFe-modified graphite electrodes, comparing the effects of compressed and expanded graphite substrates on catalytic performance. Our study reveals that substrate geometry plays a pivotal role in catalyst distribution and activity, with expanded graphite facilitating more effective electron transfer and active site utilization. Additionally, we observe that increasing the NiFe loading leads to only modest gains in performance, due to catalyst agglomeration at higher loadings. The optimized NiFe–graphite composites exhibit superior stability and catalytic activity, achieving lower overpotentials and higher current densities, making them promising candidates for sustainable hydrogen production via alkaline electrolysis.

1. Introduction

The quest for sustainable and efficient energy sources has led to a significant focus on alkaline water splitting, a promising method for clean and renewable hydrogen production. Central to this process is the role of electrocatalysts, among which NiFe-based catalysts have emerged as a frontrunner. Their deposition on conductive substrates, such as expanded graphite and compressed graphite, enhances the performance of graphite for both hydrogen and oxygen evolution reactions (HER and OER, respectively), which are integral to the phenomenon of water electrolysis [1,2,3,4,5,6,7,8].
NiFe catalysts are distinguished by their unique structural and electronic properties. Combined investigations by means of X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) techniques have revealed square planar molecular structures in these catalysts, this being especially advantageous for their catalytic activity [9,10].
The electrochemical oxygen evolution and surface redox chemistry of NiFe layered double hydroxides (LDHs) have also been studied. The results produced by cyclic and linear sweep voltammetry showed that NiFe LDHs exhibited excellent catalytic activity under alkaline conditions, significantly outperforming Fe-free hydroxide analogs. Their low overpotentials for the OER indicate superior catalytic efficiency (with a current density of 10 mA cm−2 for an overpotential of 348 mV) [11].
The X-ray diffraction (XRD) analysis of NiFe nanoparticles (NPs) further elucidates their structural features, which are essential for understanding their catalytic behavior. The unique composition and formation of these NPs, as indicated by the presence of γ-Fe2O3 and the absence of nickel oxide diffraction peaks, suggest a specific structural configuration, leading to enhanced OER catalytic activity [12].
A key challenge in improving NiFe catalysts is the optimization of their interaction with the supporting substrate, which plays a crucial role in determining the overall catalytic performance. Substrate geometry and surface characteristics significantly influence the distribution of the catalyst, the accessibility of active sites, and the efficiency of electron transfer. While extensive research has focused on optimizing the composition and synthesis of NiFe catalysts, less attention has been paid to the physical properties of the substrates used and how these influence catalytic performance. Studies have shown that factors such as substrate morphology can dramatically affect catalytic activity, yet a systematic investigation into these effects, particularly in the context of NiFe-modified graphite, is missing.
This study focuses on the exploration of NiFe catalysts, specifically their deposition and behavior toward HER and OER processes of the expanded graphite/NiFe and compressed graphite/NiFe catalyst composites produced in this way. The choice of graphite as a substrate stems from its excellent conductivity, mechanical stability, and cost-effectiveness, making it an ideal candidate for large-scale applications [13]. Our investigation delves into the synthesis and characterization of NiFe catalysts on these substrates, aiming to understand the interaction between the catalyst and the substrate and how this impacts the overall efficiency of electrochemical water-splitting.
The novelty of this work lies in its dual focus on substrate geometry and catalyst loading, which together offer a more comprehensive understanding of how to optimize the content of NiFe-based catalysts for practical applications in hydrogen production. By addressing these critical factors, our study contributes to broader efforts to develop more efficient and scalable water electrolysis systems, aligning with the ongoing research aimed at overcoming the bottlenecks associated with current NiFe oxyhydroxide catalysts.

2. Materials and Methods

2.1. Solutions and Chemical Reagents

All solutions were prepared using ultra-pure water with a resistivity of 18.2 MΩ cm, obtained from a Spring 30 s water purification system (Hydrolab, Gdańsk, Poland). First, 0.1 and 1.0 M NaOH supporting solutions were prepared using sodium hydroxide pellets (99.9%, POCH, Gliwice, Poland). Then, a palladium reversible hydrogen electrode (RHE) was charged in a 0.5 M H2SO4 solution prepared from concentrated sulfuric acid (98%, Merck, Darmstadt, Germany).

2.2. Electrodes and Electrochemical Cell

Electrochemical experiments were conducted in a standard electrochemical cell made of Pyrex glass. The cell comprised three electrodes: a graphite-based working electrode (WE), a Pd RHE (Pd wire, 99.99% purity, 1.0 mm diameter, Sigma-Aldrich, Poznań, Poland) acting as a reference electrode (RE), and a Pt counter electrode (CE; Pt wire, 99.99% purity, 1.0 mm diameter, Sigma-Aldrich). Each electrode was placed in a separate compartment. Before carrying out any experiments, atmospheric air was removed from the cell by bubbling with argon (Ar 5.0 grade, Linde Gas Poland, Olsztyn, Poland), and an argon gas flow was continuously maintained above the solutions throughout the experiments.

2.3. Expanded Graphite and Compressed Graphite Electrodes

The expanded graphite and compressed graphite (see details in Table S2) plate electrodes, used as the working electrodes, had a size of 2 × 2 cm and corresponding thicknesses of 1.0 and 3.0 mm. Initially, they were ultrasonicated in acetone in order to remove any surface impurities that could otherwise be present.
The electrodes’ geometrical surface areas (GSA) for expanded graphite and compressed graphite were 8.6 and 9.8 cm2, respectively. All electrochemical parameters are reported based on these GSA estimations. However, results based on the electrochemically active surface area (ECSA) are also included to provide a more detailed understanding of their electrochemical behavior. The ECSA, estimated from a double-layer capacitance (Cdl), offers a more accurate measure of the active sites available for electrochemical reactions. The values of ECSA for the graphite electrodes are provided in the Supplementary Materials.
Electrodeposition of the NiFe catalyst on the graphite electrodes was carried out with four different loadings: 0.5 mg, 2.5 mg, 5.0 mg, and 7.5 mg of NiFe per 1 cm2. The conditions for the NiFe electrodeposition and composition of produced deposits are reported in Table S1 and Figure S1 of the Supplementary Materials.
Procedures for setting up the laboratory equipment and materials were consistent with those previously described in other works from this laboratory [14,15,16].

2.4. Experimental Methodology

Electrochemical measurements:
Electrochemical measurements were conducted, utilizing an AUTM 204 + FRA 32M Multi-Autolab potentiostat/galvanostat system in a 0.1 M NaOH solution as the supporting electrolyte. The choice of working solution enabled the best comparisons of the results with the existing literature.
Quasi-steady-state Tafel polarization: This technique was used to study the kinetics of the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) on the NiFe-modified expanded graphite and compressed graphite electrodes. Tafel polarization experiments were conducted at a scan rate of 0.5 mV s−1.
Electrochemical impedance spectroscopy (EIS): EIS measurements were conducted to investigate the electrode kinetics and charge transfer processes. The frequency range for EIS was typically between 1.0 × 105 and 0.5 × 10−1 Hz, with an output signal amplitude of 5 mV.
Cyclic voltammetry (CV): CV measurements were performed to assess the electrochemical behavior and stability of the electrodes. Three sweeps were always carried out over the potential span of −1.0 to 1.8 V vs. RHE at a scan rate of 5 and 50 mV s−1.
Surface characterization:
A combination of advanced spectroscopy techniques was employed to comprehensively characterize the surface properties of the produced electrodes.
Scanning electron microscopy/Energy-dispersive X-ray (SEM/EDX) surface spectroscopy characterization was performed for all examined samples, including the NiFe-modified expanded graphite and compressed graphite. These analyses were carried out using a Quanta FEG 250 (FEI Company, Brno, Czech Republic) microscope equipped with Bruker XFlash 6010 EDX (Berlin, Germany) instrumentation (with a resolution of 125 eV).
X-ray diffraction (XRD) analysis: The phase composition of the bulk samples was analyzed using an X-ray diffraction method by means of the X’Pert Pro MPD Philips (Panalytical, Almelo, Netherlands) diffractometer, using Cu Kα (1.542 Å) radiation at room temperature.
Atomic force microscopy (AFM) measurements were conducted using a Nanosurf CoreAFM (Liestal, Switzerland) system to characterize the surface topography and roughness of both unmodified and NiFe-modified graphite electrodes. For each sample, multiple areas were scanned to ensure representative results. The scan sizes were set to 1 × 1, 5 × 5, and 10 × 10 μm. The images were processed using the Nanosurf software v. 3.8 to analyze the surface roughness parameters.
Software and data analysis:
The electrochemical instruments were controlled by means of Nova 2.1 software (Metrohm Autolab B.V., Opacz-Kolonia, Poland). All impedance data were analyzed using the ZView 2.9 software package (Scribner Associates, Inc., Berwyn, PA, USA). The impedance spectra were fitted by the LEVM 6 program, a complex, non-linear, least-squares immittance fitting tool [17].
XT microscope control software (v. 2.3) ran the SEM microscope for image capture and analysis. Meanwhile, Bruker’s QUANTAX EDS has been integrated into the ESPRIT software (v. 2.1) suite for qualitative and quantitative elemental analyses.

3. Results and Discussion

3.1. Surface Characterization

3.1.1. SEM/EDS Analysis

Figure 1a–e and Figure 2a–e of this study reveal the structural characteristics of the NiFe alloy deposited on the expanded graphite and compressed graphite. The EDS images show that the NiFe structures on these substrates exhibit a regular and homogenous morphology for all NiFe loadings (see Figure 1b and Figure 2b). However, despite the overall uniformity, there were instances where the carbon element was visibly discernible within the structure, indicating areas of lesser NiFe coverage or integration.
The analysis determined that the grain size of the deposited NiFe alloy varied between 40 and 60 nm. Additionally, there were noticeable agglomerates of NiFe particles, with some reaching sizes of up to 150 nm. The weight of NiFe alloy for each loading amount (0.5, 2.5, 5.0, and 7.5 mg of NiFe per cm2) was confirmed by the weighting method, with an error of about 4.5%. Furthermore, some variations were observed in the elemental composition percentages in the SEM/EDX analysis of the NiFe alloy content, depending on the deposition time and the graphite substrate (for more details, see Figure S1).

3.1.2. AFM Analysis

Atomic force microscopy (AFM) was employed to examine the topographical characteristics of NiFe-modified graphite electrodes. The analysis focused on an assessment of the surface roughness and distribution of the NiFe catalyst on the substrates. The AFM images (Figure 3a,b) revealed a uniform deposit of NiFe particles with visible agglomeration and higher catalyst loadings contributing to an increased active surface area, this being essential for electrocatalytic processes. Furthermore, quantitative data on surface roughness parameters were obtained (Table S2) linking the microstructural features to the electrochemical performance, as observed by cyclic voltammetry and impedance spectroscopy studies. These findings highlight the correlation between the nanoscale surface attributes and the overall efficiency of the electrodes in electrolysis applications.

3.1.3. XRD Analysis

Figure 4 presents the X-ray diffractograms for samples deposited on respectively expanded graphite (Figure 4a) and compressed graphite (Figure 4b). In both cases, a graphite substrate with a hexagonal structure is clearly visible. In the case of compressed graphite, the structure is more oriented and only the planes of the 002 family are visible. The presence of the NiFe phase can be observed above surface concentrations of 2.5 mg cm−2, and the amount of this phase increases with rising NiFe concentration. The NiFe phase crystallizes in a regular Pm-3m structure, and the significant broadening of the observed peaks indicates its nanostructural character.

3.2. Electrochemical Characterizations

3.2.1. Cyclic Voltammetry

The cyclic voltammetry (CV) analysis detailed in Figure 5a,b of this study provides a comparative insight into the electrochemical behavior of all the examined electrodes (expanded graphite, compressed graphite, and all their NiFe modifications) in contact with a 0.1 M NaOH solution. The latter was achieved through CV sweeps across the potential range of −1.0 to 1.8 V vs. RHE, with a scan rate of 50 mV s−1, focusing on the last cycles to assess the voltammogram’s stability and reproducibility. The deposition of the NiFe alloy onto the graphite substrates led to a marked improvement in the catalysis of both the HER and OER processes. However, an increase in the NiFe loading did not proportionately enhance the catalytic properties. For the expanded graphite, the shape of the CV profiles suggested that while there is an initial improvement in catalytic behavior with the introduction of the NiFe alloy, further raising the NiFe loading value does not lead to a reasonable increase in the catalytic performance. This near-plateau in observed current density (Figure 5a) could most likely be attributed to the process of agglomeration of the catalyst particles at higher loadings, which may then inhibit the electron transfer process and active site utilization [18,19].
In contrast to the expanded graphite, the CV profiles for the compressed graphite electrodes displayed an inverse relationship between the NiFe loading and the catalytic performance. The obtained data suggest that the lowest NiFe loading yields the best catalytic behavior for the compressed graphite, outperforming those of composite electrodes with higher NiFe loadings. This unexpected trend could imply a more homogeneous distribution and utilization of active sites at lower catalyst loadings on this material, or else it might reflect differences in the interaction between the NiFe particles and the compressed graphite surface. Such a trend underscores the complex nature of catalyst–substrate interactions and the importance of optimizing loading levels for different substrate geometries to achieve the best possible electrochemical performance. Moreover, the characteristics of the NiFe oxidation/reduction peaks in the CV profiles were in line with the phenomena regarding the formation/transition of hydroxide and oxyhydroxide phases described in previous research from this laboratory [13].
Additionally, cyclic voltammetry measurements were used to determine the working electrodes’ double-layer capacitance, which is strongly correlated to the electrochemically active surface area (ECSA). In order to accurately assess the double-layer capacitance (Cdl) of electrodes, CV measurements were conducted around the potential of 0.7 V vs. RHE, within a potential sweep range of ±0.10 V. This specific potential range was chosen to ensure operations within the non-Faradaic CV region, focusing on capturing the capacitive behavior of the electrode–electrolyte interface without interference from redox processes. The sweep rates for these CV measurements varied between 5 and 100 mV s−1, allowing for an examination of the current response across a broad range of kinetic conditions. The Cdl was determined by analyzing the relationship between the capacitive current and the sweep rate. In the absence of Faradaic processes, the capacitive current at a given potential is expected to be directly proportional to the sweep rate, following the equation I = vC, where I is the current, v is the sweep rate, and C is the capacitance. The slope obtained from plotting the capacitive current against the sweep rate, as illustrated in Figure S2, directly quantifies the Cdl parameter [20,21]. The values derived from this analysis are presented in Table 1.
The change in the Cdl in terms of the function of the NiFe loading on expanded graphite and compressed graphite shows a clear trend. For the expanded graphite, as the loading of NiFe increases from 0 to 7.5 mg cm−2, the Cdl significantly decreases from 9000 µF to 2500 µF. This phenomenon is likely attributable to the NiFe catalyst layer’s tendency to cover the intrinsic microstructure of the expanded graphite, potentially obscuring some of its active sites.
On the compressed graphite, however, the trend is not as straightforward. Starting with an unmodified electrode with a Cdl of 2900 µF, there is an initial capacitance increase to 22,400 µF at a NiFe loading of 0.5 mg cm−2, which might indicate that small amounts of NiFe catalyst enhance the electroactive surface area or affect the surface roughness in a manner that considerably increases the capacitance. Beyond an initial capacitance increase at a low NiFe loading, the Cdl value demonstrates a decreasing trend with further increments of the NiFe catalyst on the compressed graphite. This reduction in the capacitance suggests that beyond a certain level, additional NiFe deposits may lead to a reduction in the electrode’s available surface area for charge accumulation, potentially due to particle agglomeration or changes in the surface microstructure, as evidenced by the SEM analysis (see Figure 1 and Figure 2) [22,23].
As presented in Table S2, the AFM analysis revealed that the compressed graphite had a more developed surface area than that of the expanded graphite, thus implying its higher potential for electroactivity. However, the electrochemical tests, such as cyclic voltammetry or AC impedance, indicated that the expanded graphite actually exhibited higher activity. This apparent contradiction could be explained by considering the active sites’ nature and activity. The AFM results imply that while the compressed graphite has a larger surface area (Table S2), not all of its area may consist of electrochemically active sites. Despite having a less developed surface, the expanded graphite presents a higher proportion of accessible active sites. However, when NiFe is deposited at a low loading level on the graphite substrates, the Cdl parameter for the compressed graphite shows a significant increase, opposite to the expanded graphite, based on the CV measurements. This suggests that the NiFe catalyst effectively utilizes the larger surface area of the compressed graphite at its low loadings, enhancing the electroactivity by interacting with the existing active sites and potentially creating new ones by the deposition of NiFe. However, as the NiFe loading increases, the subsequent decrease in the Cdl might indicate that the NiFe catalyst starts agglomerating (see Figure 1 and Figure 2) or commences blocking the substrate’s pores, thereby reducing the effective substrate’s surface area. This explanation aligns with the observed electrochemical behavior, whereby an initial deposition of NiFe improves the catalyst activity, as indicated by the increased Cdl values; further NiFe deposition leads to deterioration of the catalytic performance, which is likely due to overcrowding of the catalyst and a significant reduction in the number of accessible active sites. This subtly led to the use of the dimensional geometrical area as a means to better characterize the materials, thus enabling their more direct comparison. This approach takes into consideration the practical application of these materials in real-scale alkaline water electrolyzers.

3.2.2. AC Impedance—HER

Figure 6a and Table 2 of this article illustrate the selected impedance spectroscopy results (the full potential range is presented in Table S3) for electrodes modified with NiFe on the expanded graphite and the compressed graphite base material electrodes, all tested in 0.1 M of NaOH solution. The electrochemical parameters, such as charge transfer resistance (Rct), porosity resistance for reaction intermediates (Rp), double-layer capacitance (Cdl), and pseudo-capacitance (Cp), were derived from the data using equivalent circuit models, as depicted in Figure 6b–d.
For unmodified expanded graphite electrodes, the impedance measurements at low overpotentials (from −50 to −400 mV vs. RHE) demonstrated a single depressed semicircle at high frequencies, indicative of a porosity response, and a linear region at medium to low frequencies that corresponded to capacitive behavior. As the probing potential was displaced to more negative values, specifically, to the range between −450 and −700 mV vs. RHE, an additional semicircle appeared in the impedance spectrum, signifying the onset of the HER process at medium and low frequencies. For the expanded graphite electrodes, the Rp and Cp parameters were largely found to be potential-independent in reference to the intrinsic characteristics of the electrode’s porous entity. However, a significant increase in the Cdl parameter was observed from 7999 to 12,815 µF cm−2, as the cathodic overpotential increased from 50 to 700 mV. The latter may indicate that as the overpotential rises, more of the expanded graphite’s surface area becomes electrochemically active, which could be due to the electrode’s relatively poor catalytic properties and highly electrochemically non-uniform surface. Additionally, the Rct parameter associated with HER reactivity, which was monitored in the potential range of −450 to −700 mV, showed a notable decrease from 396.2 to 24.5 Ω cm2. This substantial reduction in the charge transfer resistance is characteristic of the kinetically controlled process.
Then, upon NiFe alloy deposition, the expanded graphite’s “porosity” response, indicated by the semicircle corresponding to the Rp and Cp parameters, was no longer discernible at more negative potentials than −150 mV. This change is attributed to the extended formation and accumulation of hydrogen (H2) bubbles during the HER process.
Additionally, the introduction of the NiFe catalyst onto the expanded graphite significantly enhanced the HER kinetics, as evidenced by the onset of the HER process at lower overpotentials (Figure 5a). Also, catalytic graphite modification caused a radical reduction in the Rct value by approximately 57 times for a NiFe loading of 0.5 mg cm−2, as compared to the Rct recorded for the unmodified expanded graphite at a potential of −450 mV. This reduction in the Rct indicates a significantly improved rate of electron transfer during the HER, highlighting the effectiveness of the NiFe catalyst in promoting the hydrogen evolution process, which is also correlated with an increase in the current density observed in the HER region for the CV measurements. Also, for the NiFe-modified electrodes under increasing cathodic overpotentials (from −50 to −450 mV), a substantial reduction in the Rct and Cdl parameter values was observed (see Table 2). Hence, the Rct parameter decreased from 123.1 Ω cm2 to 6.1 Ω cm2 within the tested potential range, indicating a significant enhancement of the HER kinetics at the electrode surface, while a decrease in the Cdl parameter from 13,277 to 3474 µF cm−2 was most likely caused by the evolution of hydrogen gas at the electrode surface, which could considerably obstruct its electroactive areas, thus leading to a reduced effective surface area.
In our exploration of the NiFe catalyst loadings on expanded graphite electrodes for enhanced electrocatalytic applications, our investigation revealed a subtle relationship between the amount of catalyst and the observed catalytic performance. Initial loading of the NiFe (0.5 mg cm−2) on the expanded graphite electrodes resulted in a marked improvement in the electrochemical behavior, as evidenced by the reduction in the charge transfer resistance and an increase in the double-layer capacitance (see Figure 6a and Table 2). However, as the catalyst loading was increased to 2.5, 5.0, and 7.5 mg cm−2, the anticipated continuation of the performance enhancement did not materialize. Instead, a decline in the catalytic efficiency of the composites was observed; namely, the Rct parameter increased by 1.01, 1.45, and 2.24 times for 2.5, 5.0, and 7.5 mg cm−2 of the NiFe loading, respectively. The above finding is in contrast to expectations based on the composites’ increase in the real surface area, as derived based on the AFM measurements. This actual surface area increase suggests greater textural complexity, due to the catalyst’s deposition, yet the electrochemically active surface area obtained from the Cdl measurements was considerably reduced.
Higher catalyst loadings likely lead to NiFe particle agglomeration. Although this increases the electrode’s textural complexity and the real surface area, it may reduce access to active sites, thus diminishing its effective electrochemical surface area. With increased NiFe loading, the surface chemistry modifications could significantly influence the adsorption-desorption kinetics of hydrogen. The interaction between pre-existing active sites on the graphite substrate and those introduced by the lowest NiFe catalyst loadings could eventually lead to an enhancement in the composite’s electrocatalytic activity.
In contrast to the behavior of the expanded graphite, the EIS analysis of the unmodified compressed graphite electrodes (see Figure 7 and Table 2 and Table S3) exhibited two distinct semicircles across the potential range of −50 to −700 mV vs. RHE. The semicircle observed at high frequencies corresponded to the porosity of the electrode, reflecting the resistance and capacitance associated with the electrode’s porous structure. The values of the Rp and Cp parameters were relatively stable and were not influenced by the applied potential, with Rp showing fluctuations between 2.9 and 7.6 Ω cm2, and Cp values varying between 4066 and 9783 µF cm−2. The semicircle visible in the medium to low-frequency range was associated with HER reactivity. The Rct parameter decreased (from 2162.0 to 140.0 Ω cm2) with increasing overpotential, suggesting an enhancement in the kinetics of electrochemical reactions at higher overpotentials. Meanwhile, the Cdl parameter exhibited a significant increase (2766 and 40,964 µF cm−2), indicating activation of the electrochemical surface area related to the rising overpotential.
It is important to highlight that while the expanded graphite did not show significant reactivity toward the HER until reaching the potential of −450 mV, the Rct at this potential point was slightly lower than that observed for the compressed graphite, showing 1.25× enhanced catalytic activity. This disparity in the Rct values further widened with the rising overpotential and eventually reached an approximately 2.11× reduction in favor of the expanded graphite at a potential of −700 mV. Despite its delayed reactivity, this lower Rct value recorded on the expanded graphite could be due to its intrinsic structural and electrochemical properties, such as the efficient arrangement or exposure of catalytically active sites to the working solution. This nuanced behavior highlights the complex interplay between surface area expansion, active site distribution, and electrochemical activity, underscoring the finding that a larger surface area of compressed graphite does not directly translate into enhanced catalytic efficiency.
Then, upon modification with the NiFe alloy (0.5 mg cm−2), the compressed graphite electrodes demonstrated a significant improvement in their electrochemical properties. Notably, the Rct values were approximately 86.67× lower at the potential of −450 mV for the NiFe-modified compressed graphite than for its unmodified counterpart. This substantial decrease in the Rct underscores the effectiveness of the NiFe catalyst in facilitating the electron transfer process, thereby enhancing the overall electrocatalytic activity of the electrodes. Interestingly, after NiFe modification, no porosity response was detected by the AC impedance tool.
Similarly to the behavior observed in the unmodified compressed graphite, the NiFe-activated electrodes exhibited analogous trends in the Rct and Cdl parameters upon rising overpotential. Moreover, analogously to the observations of the expanded graphite-based electrodes, the compressed graphite-based specimens also exhibited a decline in their catalytic efficiency with increasing NiFe loading levels. Specifically, a reduction in their efficiency by factors of 2.06, 6.70, and 8.98 for NiFe loadings of 2.5, 5.0, and 7.5 mg cm−2 was observed, respectively, at a potential of −50 mV, corresponding to 0.5 mg cm−2. This trend underscores the complex interplay between catalyst loading and electrode performance, where beyond a certain threshold, any additional catalyst does not translate to improved efficiency and may, in fact, hinder the catalytic process, due to such important factors as catalyst agglomeration, the blocking of active sites, or the detrimental effects of increased hydrogen bubble accumulation.
The relationship between the logarithmic inverse of the charge transfer resistance (−log Rct) and overpotential (η) was analyzed for kinetically controlled reactions across a potential range from −50 to −700 mV vs. RHE, with intervals of 50 mV. This analysis enabled the calculation of exchange current densities (j0) for the HER, leveraging the Butler–Volmer equation to correlate j0 with the Rct parameter as the overpotential approached zero [14].
For expanded graphite and compressed graphite electrodes, including those modified with varying loadings of NiFe alloy (such as 0.5, 2.5, 5.0, and 7.5 mg cm−2), the calculated j0 values are presented in Table 3.
For an unmodified expanded graphite electrode, a baseline j0 value of 2.78 × 10−7 A cm−2 at low overpotential ranges signifies a small amount of inherent activity for the HER. The introduction of the NiFe catalyst markedly amplifies this activity, exhibiting the effectiveness of NiFe in promoting the hydrogen evolution process. However, an intriguing pattern emerges with increasing NiFe loadings; the j0 values at a low overpotential range (50 to 200 mV) decrease from 5.7 × 10−5 to 1.5 × 10−5 A cm−2 as the loading progresses from the lowest to the highest level. This trend suggests that additional NiFe does not translate to improved catalytic efficiency beyond a certain threshold, possibly due to factors such as catalyst agglomeration or the surface saturation of active sites.
The unmodified compressed graphite electrodes start with a higher baseline j0 of 3.12 × 10−6 A cm−2 at low overpotential ranges, compared to the expanded graphite, highlighting the differences in material properties and their effect on HER activity. Then, modification with NiFe (0.5 mg cm−2) particles significantly increases the j0 values to 1.7 × 10−4 and 4.3 × 10−4 A cm−2 for low and high overpotential ranges (50–200 and 200–500 mV), respectively. This enhancement underscores the potential of NiFe as an effective catalyst across different substrate materials. Similarly to the expanded graphite electrodes, the compressed graphite materials also exhibit a descending trend in the j0 values with increased loadings of NiFe, culminating in a decrease to about 1.2 × 10−5 and 2.6 × 10−4 A cm−2 for low and high overpotential ranges at the highest NiFe loading. This decline across both low and high overpotential ranges for compressed graphite electrodes further emphasizes the subtle balance between optimizing catalyst loading and achieving maximum catalytic efficiency.

3.2.3. Polarization Technique—HER

The Tafel polarization plots for expanded graphite and compressed graphite electrodes, both unmodified and activated with various NiFe loadings (0.5, 2.5, 5.0, and 7.5 mg cm−2), are illustrated in Figure 8. The cathodic Tafel slopes (bc) and the exchange current densities (j0) for the HER are detailed in Table 3. For the NiFe-modified expanded graphite and compressed graphite samples, measurements were taken within the potential range of −50 to −200 mV vs. RHE. However, due to the delayed onset of hydrogen evolution on the unmodified graphite samples, a more negative potential range of −700 to −900 mV vs. RHE was selected for these electrodes. This shift in potential range for the onset of hydrogen evolution on unmodified graphite electrodes aligns with previous observations made with the EIS results.
The j0 values derived from the Tafel plots are in line with those calculated by means of the Butler–Volmer equation, indicating a significant enhancement in catalytic properties following NiFe modification. Additionally, these electrodes demonstrated a more positive onset potential for H2 evolution, compared to their unmodified counterparts (see Figure 9). The exchange current density values derived from this analysis are in good agreement with the literature values, reinforcing the validity of the observed trends [24,25].
The exchange current density (j0) is a fundamental parameter that provides deep insights into the intrinsic electrochemical activity of an electrode material, highlighting how readily the HER can be initiated without significant energy barriers. This parameter essentially quantifies the electrode’s efficiency in catalyzing the HER at a point where the overpotential is zero, making it a valuable indicator of the material’s catalytic capability under equilibrium conditions.
Despite the significant role of j0 in providing insights into the intrinsic electrochemical activity of electrode materials for the HER, it is frequently observed that many studies in the field prioritize parameters, such as the Tafel slope (bc), and operational metrics, such as the overpotential needed to reach a current density of 10 mA cm−2. This focus stems from the relative ease of measuring these operational standards, compared to calculating the j0 parameter, which demands in-depth electrochemical analysis incorporating the Butler–Volmer equation. By providing direct, application-relevant insights into electrode performance, operational metrics become invaluable for comparative and benchmarking purposes.
In order to align our study with common literature practices, especially regarding research emphasizing operational benchmarks, we also conducted these specific measurements in 1 M of NaOH solution (see Figures S3 and S4). The concentrations used in this work are widely recognized for their ability to offer a consistent and relevant benchmark for comparison, thus allowing for the direct analogy of the operational performance of our electrode materials against a wide array of existing research findings. By selecting both concentrations for these measurements, our study not only adheres to established experimental standards but also ensures a clearer and more impactful comparison of our results with those in a broader field. These operational metrics are presented in Table 3, thus ensuring a comprehensive overview of our electrodes’ performance in a context that maximizes both scientific rigor and relevance to ongoing research efforts.
Table 3. HER kinetic parameters for the selected catalytic materials.
Table 3. HER kinetic parameters for the selected catalytic materials.
Base MaterialCatalystLoading [mg cm−2]ElectrolyteEISTafelRefs.
j0 [A cm−2]j0 [A cm−2]bc [mV dec−1]η(j=10 mA cm−2) [mV]
Low ηHigh η
Unmodified expanded graphite None00.1 M NaOH2.78 × 10−72.78 × 10−71.97 × 10−6203-This article
1.0 M NaOH--2.18 × 10−6163596
Expanded graphiteNiFe0.50.1 M NaOH5.67 × 10−52.37 × 10−42.00 × 10−4137290This article
1.0 M NaOH--1.40 × 10−4112200
Expanded graphiteNiFe2.50.1 M NaOH5.41 × 10−52.31 × 10−41.80 × 10−4133245This article
1.0 M NaOH--1.35 × 10−4110203
Expanded graphiteNiFe5.00.1 M NaOH3.10 × 10−52.32 × 10−41.03 × 10−4116252This article
1.0 M NaOH--2.63 × 10−4123198
Expanded graphiteNiFe7.50.1 M NaOH1.53 × 10−52.34 × 10−43.14 × 10−577219This article
1.0 M NaOH--1.40 × 10−4119222
Unmodified compressed graphite None00.1 M NaOH3.12 × 10−63.06 × 10−65.00 × 10−584-This article
1.0 M NaOH--1.42 × 10−8116596
Compressed graphiteNiFe0.50.1 M NaOH1.70 × 10−44.31 × 10−41.20 × 10−4110194This article
1.0 M NaOH--1.13 × 10−493199
Compressed graphiteNiFe2.50.1 M NaOH7.77 × 10−52.24 × 10−41.30 × 10−4113-This article
1.0 M NaOH--1.30 × 10−496201
Compressed graphiteNiFe5.00.1 M NaOH1.65 × 10−53.48 × 10−43.47 × 10−594-This article
1.0 M NaOH--4.50 × 10−5102198
Compressed graphiteNiFe7.50.1 M NaOH1.19 × 10−52.61 × 10−43.49 × 10−5104-This article
1.0 M NaOH--2.50 × 10−5101204
Ni foamNiFeRu LDH1.21.0 M KOH---3129[24]
Ni foamNiFe LDH1.21.0 M KOH---153269[24]
Glassy carbonNi3FeN NPs0.351.0 M KOH---42158[26]
Carbon clothNi-FeP/TiN-1.0 M KOH---7375[27]
Ni foamNiFe LDH0.451.0 M KOH---120195[25]
Ni foamNiFeIr LDH0.451.0 M KOH---5651[25]
Ni foamNiFeRh LDH0.451.0 M KOH---2724[25]
Carbon clothPtNi-Ninanoarray-0.1 M KOH---4238[28]
The findings presented in this work are consistent with a significant portion of the relevant literature on electrocatalysis, demonstrating that the NiFe-modified expanded graphite and compressed graphite electrodes exhibit competitive catalytic performance for the HER without necessitating complex preparation methods. The above method contrasts with some documented sophisticated approaches, such as the in situ growth of ultrafine PtNi nanoparticle-decorated Ni nanosheet arrays on carbon cloth [28]. Such methods, while yielding high-performance catalysts, often involve complex preparation steps that can limit their scalability and practical application.
However, it is important to acknowledge that there are instances within the literature where certain catalysts exhibit superior performance metrics. These standout examples often share a common factor: the surface area used for calculating performance results is sometimes questionable or it is not directly comparable with standard measurements. This discrepancy could lead to challenges in directly comparing the efficiency of these advanced materials with those prepared using uncomplicated procedures.
In essence, while the latter approach may not match the peak performance of the most sophisticated catalysts in every instance, it offers a valuable balance between ease of preparation, cost-effectiveness, and competitive catalytic activity. This balance is crucial for the advancement of practical electrocatalytic applications, where scalability and accessibility of preparation techniques are as important as the ultimate efficiency of the catalyst.

3.2.4. AC Impedance—OER

The EIS results for the expanded graphite and compressed graphite electrodes, both in their unmodified form and activated with various loadings of NiFe (0.5, 2.5, 5.0, and 7.5 mg cm−2), are gathered in Table 4 (the full potential range is presented in Table S4). With respect to the OER, Figure 10 elucidates that the incorporation of NiFe modifications onto the base graphite electrodes significantly elevates their electrocatalytic reactivity.
Across the range of potentials tested for the unmodified expanded graphite electrodes, the Rp and Cp parameters demonstrated variability, lacking a definitive pattern, yet generally showing a tendency to decrease with an increase in electrode potential. Further insights into the behavior of these parameters and their implications for electrochemical performance were elaborated upon in Section 3.2.2, providing a deeper understanding of how the potential variations influence electrode dynamics. The catalytic OER enhancement due to NiFe deposition was pronounced for the expanded graphite electrode, with the disappearance of a semicircle related to the Rp and Cp parameters. Notably, the introduction of the NiFe particles resulted in an Rct reduction by approximately 60 times and a Cdl increase by about 2 times at the potential of 1600 mV. Similarly to the HER findings, the catalytic properties for the OER on the expanded graphite were significantly augmented by NiFe deposition rather than a mere increase in the electrochemically active surface area.
In contrast, the compressed graphite electrode did not display the semicircle typically indicative of a porosity response, and subsequent NiFe modifications left this feature unchanged. Modification with the NiFe alloy led to a notable improvement in the OER activity, as evidenced by a reduction in the Rct parameter by approximately 100 times at the potential of 1600 mV, with significant alterations in the Cdl parameter by ca. 39 times. Such radical changes in both the Rct and Cdl parameters underscore the effectiveness of the NiFe alloy in optimizing the electrode’s surface catalytic properties for the OER, effectively enhancing its catalytic activity and implying improved alignment of the electrochemical properties with the desired reaction dynamics.
For all NiFe-modified electrodes (using the same base, with different loadings), an observable reduction trend in the reaction resistance with rising overpotential values could be noted, particularly for the lowest NiFe loading on the expanded graphite, where the resistance diminished from 41.1 to 3.4 Ω cm2, while for analogous NiFe loading on the compressed graphite samples, the Rct reduced from 32.2 to 4.3 Ω cm2 for the potential range of 1500–1800 mV, respectively. Unlike in the HER scenario, the Cdl parameter for the OER showed an unspecific fluctuation with rising electrode potential for the NiFe-modified expanded graphite and compressed graphite-based electrodes, most likely indicating the significant impact of O2 bubble formation on the electrode surface (Table 4 and Table S4).
It was observed that while an initial NiFe loading of 0.5 mg cm−2 significantly enhanced the electrochemical behavior, as demonstrated by decreased Rct and increased Cdl parameter values, further increases in the NiFe loading did not continue to enhance their OER performance. Despite the anticipated benefits of having increased the catalyst presence, the performance reached a plateau, suggesting that beyond a certain threshold, additional NiFe mass does not contribute to further catalytic efficiency. This outcome suggests that, similarly to the HER behavior, there is an optimal NiFe loading level for maximizing the OER activity on expanded graphite electrodes beyond which the benefits diminish, possibly due to such factors as catalyst agglomeration or the obstruction of active sites, which counteract the potential advantages of having a larger catalytic surface area.

3.2.5. Polarization Technique—OER

The Tafel polarization curves for both the unmodified and NiFe-modified expanded graphite and compressed graphite electrodes are illustrated in Figure 11a,b, demonstrating the electrocatalytic OER performance enhancements brought about by NiFe surface modification. The analysis of these curves within the 1500–1600 mV potential range provides a detailed view of the electrodes’ performance, with anodic slope (ba) and current density at an overpotential of 0.3 V (j(ŋ = 0.3 V)), as presented in Table 5 for comprehensive analysis.
Further highlighting the catalytic improvements, Figure 12a,b demonstrates that the NiFe-modified samples achieve significantly lower OER onset potential compared to their base material counterparts, signifying an earlier initiation of the oxygen evolution process. This performance enhancement positions the NiFe-modified electrodes on a par with or superior to those utilizing noble metal catalysts such as platinum, ruthenium, and iridium, as evidenced by the data recorded via the comparative analysis provided in Table 5. Table 5 highlights the effectiveness of NiFe as a catalyst and showcases a variety of transition metal combinations with comparable catalytic properties. This revelation significantly expands the pool of potential catalysts available for such applications and provides an opportunity to uncover more efficient and cost-effective catalytic materials that can be tailored to meet the demands of this important electrochemical process.
The recorded current density values (Table 5) at an anodic overpotential of 300 mV closely align with those of bulk NiFe-layered double hydroxide (LDH) materials, emphasizing the competitive nature of the NiFe modifications. By achieving even higher current densities than on such materials as IrO2 or CoP (see Table 5), the NiFe-modified expanded graphite and compressed graphite electrodes stand out for their superior electrocatalytic performance and their simplicity and scalability potential. This performance underscores the value of further research aimed at optimizing these materials, especially for enhancing water electrolysis technology.
Moreover, the overpotentials recorded at a current density of 10 mA cm−2 for the NiFe-modified electrodes with various NiFe alloy loadings underscore their efficacy and potential superiority when compared with other catalysts, as detailed in Table 5. These findings collectively emphasize the significant role of the NiFe modifications in advancing the electrocatalytic performance of graphite-based electrodes for OER applications, balancing between innovative catalytic efficiency and the practicality of their broader industrial use.
All NiFe-modified electrodes exhibit similar polarization behavior across the different electrolyte concentrations. However, there are noticeable differences among specific NiFe loadings; a clear trend shows that higher loadings initiate passivation at lower potentials. For instance, passivation starts at approximately 1.35 V for NiFe loadings of 0.5 mg cm−2, whereas for loadings of 7.5 mg cm−2, it begins around 1.15 V. This passivation process involves the formation of a protective Ni(OH)2 and Fe(OH)3 layer on the nickel-iron surfaces through a series of electrochemical reactions, leading to the development of this insoluble film [29,30]. As the potential increases beyond about 1.4 V, the formation of the β-NiOOH film commences and lasts until it reaches about 1.5 V. In the so-called “transpassive” region, beyond 1.5 V, the anodic current density dramatically increases. This phenomenon indicates a breakdown of the passivation layer, which may similarly affect the NiFe-modified electrodes, particularly given the similarity in chemical behavior that is imparted by the nickel component [31]. Thus, the long-term stability and corrosion resistance of these NiFe-modified electrodes under high potential operation in the transpassive region remain subjects of concern.
Table 5. OER kinetic parameters for the selected catalyst materials.
Table 5. OER kinetic parameters for the selected catalyst materials.
Base MaterialCatalystLoading
[mg cm−2]
ElectrolyteTafelRefs.
j0.3V [A cm−2]ba [mV dec−1]η(j = 10 mA cm−2) [mV]
Pure expanded graphiteNone00.1 M NaOH6.9 × 10−5111-This article
1.0 M NaOH1.3 × 10−479-
Expanded graphiteNiFe0.50.1 M NaOH4.2 × 10−371-This article
1.0 M NaOH1.7 × 10−255286
Expanded graphiteNiFe2.50.1 M NaOH2.8 × 10−368-This article
1.0 M NaOH1.4 × 10−283270
Expanded graphiteNiFe5.00.1 M NaOH2.3 × 10−377-This article
1.0 M NaOH2.2 × 10−271278
Expanded graphiteNiFe7.50.1 M NaOH1.7 × 10−391-This article
1.0 M NaOH2.2 × 10−266282
Pure compressed graphite None00.1 M NaOH2.4 × 10−4126-This article
1.0 M NaOH1.0 × 10−4124-
Compressed graphite NiFe0.50.1 M NaOH7.3 × 10−355-This article
1.0 M NaOH2.7 × 10−236260
Compressed graphite NiFe2.50.1 M NaOH2.4 × 10−359-This article
1.0 M NaOH2.6 × 10−256277
Compressed graphite NiFe5.00.1 M NaOH6.7 × 140−340-This article
1.0 M NaOH6.4 × 10−352310
Compressed graphite NiFe7.50.1 M NaOH1.3 × 10−365-This article
1.0 M NaOH1.0 × 10−254300
Fluorine-doped tin oxide glassNi3Co3Fe3 LDH-1.0 M NaOH-65290[32]
Ni foamNiCo LDH1.760.1 M KOH-113290[33]
Glassy carbonO-NiCoFe LDH0.120.1 M KOH--420[34]
Ni foamNiFe LDH/N-rOG0.360.1 M KOH-63258[35]
Carbon fiber clothNiCoFe LDH0.41.0 M KOH-32239[36]
Ni foamNi3FeAl0.91 LDH0.51.0 M KOH-57304[37]
Ni foamNiFeRu LDH1.21.0 M KOH-32225[24]
Ni foamNiFe LDH1.21.0 M KOH--230[24]
Carbon paperNiFe LDH-UF0.351.0 M KOH-32254[38]
Glassy carbonIrO20.711.0 M KOH 76340[39]
Glassy carbonCoP NP5.01.0 M KOH 99340[39]
Glassy carbonRuO2/CeO2-0.1 M NaOH1.0 × 10−344-[40]
-RuCu nanosheets-1.0 M KOH--234[41]

4. Conclusions

In conclusion, this investigation into the NiFe alloy modifications of expanded graphite and compressed graphite electrodes unveils a significant advancement in the realm of electrocatalysis, specifically for applications within alkaline water electrolyzers. The deposition of NiFe at various loadings markedly enhances the electrochemical performance of the HER and OER. This is evidenced by the notable reduction of the charge transfer resistance parameter, which indicates a more efficient electrochemical process under operational conditions.
Crucially, this study highlights an optimal threshold for NiFe loading (namely, 0.5 mg cm−2), beyond which any additional catalyst material does not yield proportional gains in electrode performance, potentially due to such factors as catalyst agglomeration or active site saturation. The exploration of transition metal combinations further broadens the horizons for developing more efficient, cost-effective catalysts that are tailored to specific electrochemical processes, thus opening a rich field for future research and development.
While these NiFe-modified graphite electrodes demonstrate promising results, surpassing even some noble metal counterparts in specific metrics, a clear avenue still remains for their catalytic optimization. Enhancing the HER activity and fine-tuning the material’s composition and electrochemical properties will be essential to realizing a highly efficient, commercially viable alkaline water electrolysis system. The pursuit of green hydrogen production as a clean energy carrier underscores the importance of such advancements, driving toward sustainable and economically feasible solutions in the energy sector.
Thus, the next stage should involve constructing an electrolyzer featuring a stack using such advanced electrodes, a critical step that is outlined for forthcoming research endeavors. This initiative is integral to transitioning from laboratory discoveries of tangible, scalable solutions in the realm of green hydrogen production, aligning with broader objectives to foster sustainable and economically viable energy alternatives. These efforts, planned for future works, will be instrumental in evaluating the practical efficacy, durability, and overall system efficiency, along with a comprehensive understanding of the potential uses for these electrodes in real-world applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29194755/s1, Figure S1: Elemental Fe and Ni content and ratio depending on the deposition time for expanded graphite (a) and compressed graphite (b) composites; Figure S2: Average current measured at the potential of 0.7 V vs. RHE plotted as a function of scan rate. The slope of the linear fit gives the double-layer capacitance; Figure S3: Quasi-potentiostatic cathodic polarisation curves for the HER obtained at expanded graphite (a) and compressed graphite (b) electrodes both unmodified and modified with various NiFe loadings in 1.0 M NaOH electrolyte. The curves were recorded at a scan rate of 0.5 mVs−1 (iR-corrected); Figure S4: Linear Sweep Voltammetry (LSV) curves of expanded graphite (a) and compressed graphite (b) electrodes both unmodified and modified with various NiFe loadings in 1.0 M NaOH solution carried out with a scan rate of 0.5 mV s−1 for the HER (iR-corrected); Figure S5: Quasi-potentiostatic cathodic polarisation curves for the OER obtained at expanded graphite (a) and compressed graphite (b) electrodes both unmodified and modified with various NiFe loadings in 1.0 M NaOH electrolyte. The curves were recorded at a scan rate of 0.5 mVs−1 (iR-corrected); Figure S6: Linear Sweep Voltammetry (LSV) curves of expanded graphite (a) and compressed graphite (b) electrodes both unmodified and modified with various NiFe loadings in 1.0 M NaOH solution carried out with a scan rate of 0.5 mV s−1 for the OER (iR-corrected); Table S1: Composition of electrodeposition baths and process parameters to prepare NiFe alloy deposits; Table S2: AFM derived the average increase of the surface area for NiFe-modified graphite electrodes regarding all scan sizes; Table S3: Electrochemical impedance parameters for the HER obtained on expanded graphite and compressed graphite electrodes both unmodified and modified with various NiFe loadings in 0.1 M NaOH supporting solution; Table S4: Electrochemical parameters for the OER obtained on expanded graphite and compressed graphite electrodes both unmodified and modified with various NiFe loading levels in 0.1 M NaOH supporting solution..

Author Contributions

Conceptualization, T.M. and M.K.; methodology, T.M. and M.K.; validation, T.M., M.K. and J.K.; investigation, M.K. and J.K.; data curation, M.K. and J.K.; writing—original draft preparation, T.M. and M.K.; writing—review and editing, B.P. and T.M.; supervision, B.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been primarily financed by the internal research grant no. 30.610.001-110, provided by the University of Warmia and Mazury in Olsztyn.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Dataset available on request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, W.; Cui, L.; Liu, J. Recent Advances in Cobalt-Based Electrocatalysts for Hydrogen and Oxygen Evolution Reactions. J. Alloys Compd. 2020, 821, 153542. [Google Scholar] [CrossRef]
  2. Cai, Z.; Bu, X.; Wang, P.; Ho, J.C.; Yang, J.; Wang, X. Recent Advances in Layered Double Hydroxide Electrocatalysts for the Oxygen Evolution Reaction. J. Mater. Chem. A 2019, 7, 5069–5089. [Google Scholar] [CrossRef]
  3. Liu, J.; Zhu, D.; Ling, T.; Vasileff, A.; Qiao, S.-Z. S-NiFe2O4 Ultra-Small Nanoparticle Built Nanosheets for Efficient Water Splitting in Alkaline and Neutral pH. Nano Energy 2017, 40, 264–273. [Google Scholar] [CrossRef]
  4. Kumar, A.; Bhattacharyya, S. Porous NiFe-Oxide Nanocubes as Bifunctional Electrocatalysts for Efficient Water-Splitting. ACS Appl. Mater. Interfaces 2017, 9, 41906–41915. [Google Scholar] [CrossRef] [PubMed]
  5. Xu, X.; Zhong, Y.; Wajrak, M.; Bhatelia, T.; Jiang, S.P.; Shao, Z. Grain Boundary Engineering: An Emerging Pathway toward Efficient Electrocatalysis. InfoMat 2024, 6, e12608. [Google Scholar] [CrossRef]
  6. Abdelghafar, F.; Xu, X.; Jiang, S.P.; Shao, Z. Designing Single-Atom Catalysts toward Improved Alkaline Hydrogen Evolution Reaction. Mater. Rep. Energy 2022, 2, 100144. [Google Scholar] [CrossRef]
  7. Guan, D.; Xu, H.; Zhang, Q.; Huang, Y.-C.; Shi, C.; Chang, Y.-C.; Xu, X.; Tang, J.; Gu, Y.; Pao, C.-W.; et al. Identifying a Universal Activity Descriptor and a Unifying Mechanism Concept on Perovskite Oxides for Green Hydrogen Production. Adv. Mater. 2023, 35, 2305074. [Google Scholar] [CrossRef]
  8. Zhang, H.; Guan, D.; Gu, Y.; Xu, H.; Wang, C.; Shao, Z.; Guo, Y. Tuning Synergy between Nickel and Iron in Ruddlesden–Popper Perovskites through Controllable Crystal Dimensionalities towards Enhanced Oxygen-Evolving Activity and Stability. Carbon Energy 2024, 6, e465. [Google Scholar] [CrossRef]
  9. Zang, Y.; Lu, D.-Q.; Wang, K.; Li, B.; Peng, P.; Lan, Y.-Q.; Zang, S.-Q. A Pyrolysis-Free Ni/Fe Bimetallic Electrocatalyst for Overall Water Splitting. Nat. Commun. 2023, 14, 1792. [Google Scholar] [CrossRef]
  10. Gong, L.; Yang, H.; Douka, A.I.; Yan, Y.; Xia, B.Y. Recent Progress on NiFe-Based Electrocatalysts for Alkaline Oxygen Evolution. Adv. Sustain. Syst. 2021, 5, 2000136. [Google Scholar] [CrossRef]
  11. Dionigi, F.; Zeng, Z.; Sinev, I.; Merzdorf, T.; Deshpande, S.; Lopez, M.B.; Kunze, S.; Zegkinoglou, I.; Sarodnik, H.; Fan, D.; et al. In-Situ Structure and Catalytic Mechanism of NiFe and CoFe Layered Double Hydroxides during Oxygen Evolution. Nat. Commun. 2020, 11, 2522. [Google Scholar] [CrossRef] [PubMed]
  12. Suryanto, B.H.R.; Wang, Y.; Hocking, R.K.; Adamson, W.; Zhao, C. Overall Electrochemical Splitting of Water at the Heterogeneous Interface of Nickel and Iron Oxide. Nat. Commun. 2019, 10, 5599. [Google Scholar] [CrossRef] [PubMed]
  13. Kuczyński, M.; Mikołajczyk, T.; Pierożyński, B. Alkaline Water Splitting by Ni-Fe Nanoparticles Deposited on Carbon Fibre and Nickel-Coated Carbon Fibre Substrates. Catalysts 2023, 13, 1468. [Google Scholar] [CrossRef]
  14. Pierozynski, B.; Smoczynski, L. Kinetics of Hydrogen Evolution Reaction at Nickel-Coated Carbon Fiber Materials in 0.5 M H2SO4 and 0.1 M NaOH Solutions. J. Electrochem. Soc. 2009, 156, B1045. [Google Scholar] [CrossRef]
  15. Pierozynski, B. On the Hydrogen Evolution Reaction at Nickel-Coated Carbon Fibre in 30 Wt.% KOH Solution. Int. J. Electrochem. Sci. 2011, 6, 63–77. [Google Scholar] [CrossRef]
  16. Pierozynski, B. Hydrogen Evolution Reaction at Pd-Modified Carbon Fibre and Nickel-Coated Carbon Fibre Materials. Int. J. Hydrogen Energy 2013, 38, 7733–7740. [Google Scholar] [CrossRef]
  17. Macdonald, J.R.; Kenan, W.R. Impedance Spectroscopy: Emphasizing Solid Materials and Systems, 1st ed.; Wiley-Interscience: New York, NY, USA, 1987; ISBN 978-0-471-83122-8. [Google Scholar]
  18. Boudet, A.; Henrotte, O.; Limani, N.; El Orf, F.; Oswald, F.; Jousselme, B.; Cornut, R. Unraveling the Link between Catalytic Activity and Agglomeration State with Scanning Electrochemical Microscopy and Atomic Force Microscopy. Anal. Chem. 2022, 94, 1697–1704. [Google Scholar] [CrossRef]
  19. Li, J.; Lv, Y.; Wu, X.; Zhao, K.; Zhang, H.; Guo, J.; Jia, D. Effectively Enhanced Activity of Hydrogen Evolution through Strong Interfacial Coupling on SnS2/MoS2/Ni3S2 Heterostructured Porous Nanosheets. Colloids Surf. Physicochem. Eng. Asp. 2023, 670, 131634. [Google Scholar] [CrossRef]
  20. Connor, P.; Schuch, J.; Kaiser, B.; Jaegermann, W. The Determination of Electrochemical Active Surface Area and Specific Capacity Revisited for the System MnOx as an Oxygen Evolution Catalyst. Z. Für Phys. Chem. 2020, 234, 979–994. [Google Scholar] [CrossRef]
  21. Schalenbach, M. Impedance Spectroscopy and Cyclic Voltammetry to Determine Double Layer Capacitances and Electrochemically Active Surface Areas. 2018. [Google Scholar] [CrossRef]
  22. Feng, B.; Jiang, W.; Deng, R.; Lu, J.; Tsiakaras, P.; Yin, S. Agglomeration Inhibition Engineering of Nickel–Cobalt Alloys by a Sacrificial Template for Efficient Urea Electrolysis. J. Colloid Interface Sci. 2024, 663, 1019–1027. [Google Scholar] [CrossRef]
  23. Dong, Y.; Zhang, D.; Li, D.; Jia, H.; Qin, W. Control of Ostwald Ripening. Sci. China Mater. 2023, 66, 1249–1255. [Google Scholar] [CrossRef]
  24. Chen, G.; Wang, T.; Zhang, J.; Liu, P.; Sun, H.; Zhuang, X.; Chen, M.; Feng, X. Accelerated Hydrogen Evolution Kinetics on NiFe-Layered Double Hydroxide Electrocatalysts by Tailoring Water Dissociation Active Sites. Adv. Mater. 2018, 30, 1706279. [Google Scholar] [CrossRef] [PubMed]
  25. Sun, H.; Zhang, W.; Li, J.-G.; Li, Z.; Ao, X.; Xue, K.-H.; Ostrikov, K.K.; Tang, J.; Wang, C. Rh-Engineered Ultrathin NiFe-LDH Nanosheets Enable Highly-Efficient Overall Water Splitting and Urea Electrolysis. Appl. Catal. B Environ. 2021, 284, 119740. [Google Scholar] [CrossRef]
  26. Jia, X.; Zhao, Y.; Chen, G.; Shang, L.; Shi, R.; Kang, X.; Waterhouse, G.I.N.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. Ni3FeN Nanoparticles Derived from Ultrathin NiFe-Layered Double Hydroxide Nanosheets: An Efficient Overall Water Splitting Electrocatalyst. Adv. Energy Mater. 2016, 6, 1502585. [Google Scholar] [CrossRef]
  27. Kibsgaard, J.; Tsai, C.; Chan, K.; Benck, J.D.; Nørskov, J.K.; Abild-Pedersen, F.; Jaramillo, T.F. Designing an Improved Transition Metal Phosphide Catalyst for Hydrogen Evolution Using Experimental and Theoretical Trends. Energy Environ. Sci. 2015, 8, 3022–3029. [Google Scholar] [CrossRef]
  28. Xie, L.; Liu, Q.; Shi, X.; Asiri, A.M.; Luo, Y.; Sun, X. Superior Alkaline Hydrogen Evolution Electrocatalysis Enabled by an Ultrafine PtNi Nanoparticle-Decorated Ni Nanoarray with Ultralow Pt Loading. Inorg. Chem. Front. 2018, 5, 1365–1369. [Google Scholar] [CrossRef]
  29. Díez-Pérez, I.; Gorostiza, P.; Sanz, F.; Müller, C. First Stages of Electrochemical Growth of the Passive Film on Iron. J. Electrochem. Soc. 2001, 148, B307. [Google Scholar] [CrossRef]
  30. Alsabet, M.; Grden, M.; Jerkiewicz, G. Electrochemical Growth of Surface Oxides on Nickel. Part 2: Formation of β-Ni(OH)2 and NiO in Relation to the Polarization Potential, Polarization Time, and Temperature. Electrocatalysis 2014, 5, 136–147. [Google Scholar] [CrossRef]
  31. Pierozynski, B.; Smoczynski, L. Electrochemical Corrosion Behavior of Nickel-Coated Carbon Fiber Materials in Various Electrolytic Media. J. Electrochem. Soc. 2008, 155, C427. [Google Scholar] [CrossRef]
  32. Guzmán-Vargas, A.; Vazquez-Samperio, J.; Oliver-Tolentino, M.A.; Nava, N.; Castillo, N.; Macías-Hernández, M.J.; Reguera, E. Influence of Cobalt on Electrocatalytic Water Splitting in NiCoFe Layered Double Hydroxides. J. Mater. Sci. 2018, 53, 4515–4526. [Google Scholar] [CrossRef]
  33. Jiang, J.; Zhang, A.; Li, L.; Ai, L. Nickel–Cobalt Layered Double Hydroxide Nanosheets as High-Performance Electrocatalyst for Oxygen Evolution Reaction. J. Power Sources 2015, 278, 445–451. [Google Scholar] [CrossRef]
  34. Qian, L.; Lu, Z.; Xu, T.; Wu, X.; Tian, Y.; Li, Y.; Huo, Z.; Sun, X.; Duan, X. Trinary Layered Double Hydroxides as High-Performance Bifunctional Materials for Oxygen Electrocatalysis. Adv. Energy Mater. 2015, 5, 1500245. [Google Scholar] [CrossRef]
  35. Zhan, T.; Liu, X.; Lu, S.; Hou, W. Nitrogen Doped NiFe Layered Double Hydroxide/Reduced Graphene Oxide Mesoporous Nanosphere as an Effective Bifunctional Electrocatalyst for Oxygen Reduction and Evolution Reactions. Appl. Catal. B Environ. 2017, 205, 551–558. [Google Scholar] [CrossRef]
  36. Wang, A.-L.; Xu, H.; Li, G.-R. NiCoFe Layered Triple Hydroxides with Porous Structures as High-Performance Electrocatalysts for Overall Water Splitting. ACS Energy Lett. 2016, 1, 445–453. [Google Scholar] [CrossRef]
  37. Liu, H.; Wang, Y.; Lu, X.; Hu, Y.; Zhu, G.; Chen, R.; Ma, L.; Zhu, H.; Tie, Z.; Liu, J.; et al. The Effects of Al Substitution and Partial Dissolution on Ultrathin NiFeAl Trinary Layered Double Hydroxide Nanosheets for Oxygen Evolution Reaction in Alkaline Solution. Nano Energy 2017, 35, 350–357. [Google Scholar] [CrossRef]
  38. Zhang, X.; Zhao, Y.; Zhao, Y.; Shi, R.; Waterhouse, G.I.N.; Zhang, T. A Simple Synthetic Strategy toward Defect-Rich Porous Monolayer NiFe-Layered Double Hydroxide Nanosheets for Efficient Electrocatalytic Water Oxidation. Adv. Energy Mater. 2019, 9, 1900881. [Google Scholar] [CrossRef]
  39. Chang, J.; Xiao, Y.; Xiao, M.; Ge, J.; Liu, C.; Xing, W. Surface Oxidized Cobalt-Phosphide Nanorods As an Advanced Oxygen Evolution Catalyst in Alkaline Solution. ACS Catal. 2015, 5, 6874–6878. [Google Scholar] [CrossRef]
  40. Galani, S.M.; Mondal, A.; Srivastava, D.N.; Panda, A.B. Development of RuO2/CeO2 Heterostructure as an Efficient OER Electrocatalyst for Alkaline Water Splitting. Int. J. Hydrogen Energy 2020, 45, 18635–18644. [Google Scholar] [CrossRef]
  41. Yang, J.; Ji, Y.; Shao, Q.; Zhang, N.; Li, Y.; Huang, X. A Universal Strategy to Metal Wavy Nanowires for Efficient Electrochemical Water Splitting at pH-Universal Conditions. Adv. Funct. Mater. 2018, 28, 1803722. [Google Scholar] [CrossRef]
Figure 1. SEM micrograph pictures of expanded graphite, taken at 10,000× magnification (top left); EDX elemental mappings (top right) with the maps color-coded as follows: pink for carbon, yellow for oxygen, blue for nitrogen, purple for nickel, and green for iron, while areas where green and purple overlap indicate the presence of both iron and nickel; EDX pattern (bottom) for the unmodified electrode (a) and for all NiFe loadings, namely: 0.5 (b), 2.5 (c), 5.0 (d), and 7.5 (e) mg of NiFe per cm2.
Figure 1. SEM micrograph pictures of expanded graphite, taken at 10,000× magnification (top left); EDX elemental mappings (top right) with the maps color-coded as follows: pink for carbon, yellow for oxygen, blue for nitrogen, purple for nickel, and green for iron, while areas where green and purple overlap indicate the presence of both iron and nickel; EDX pattern (bottom) for the unmodified electrode (a) and for all NiFe loadings, namely: 0.5 (b), 2.5 (c), 5.0 (d), and 7.5 (e) mg of NiFe per cm2.
Molecules 29 04755 g001aMolecules 29 04755 g001bMolecules 29 04755 g001c
Figure 2. SEM micrograph pictures of compressed graphite, taken at 10,000× magnification (top left); EDX elemental mappings (top right), with the maps color-coded as follows: pink for carbon, yellow for oxygen, blue for nitrogen, purple for nickel, and green for iron, while areas where green and purple overlap indicate the presence of both iron and nickel. EDX pattern (bottom) for the unmodified electrode (a) and for all NiFe loadings, namely: 0.5 (b), 2.5 (c), 5.0 (d), and 7.5 (e) mg of NiFe per cm2.
Figure 2. SEM micrograph pictures of compressed graphite, taken at 10,000× magnification (top left); EDX elemental mappings (top right), with the maps color-coded as follows: pink for carbon, yellow for oxygen, blue for nitrogen, purple for nickel, and green for iron, while areas where green and purple overlap indicate the presence of both iron and nickel. EDX pattern (bottom) for the unmodified electrode (a) and for all NiFe loadings, namely: 0.5 (b), 2.5 (c), 5.0 (d), and 7.5 (e) mg of NiFe per cm2.
Molecules 29 04755 g002aMolecules 29 04755 g002bMolecules 29 04755 g002c
Figure 3. AFM images comparing unmodified and NiFe-modified graphite electrodes for (a) expanded graphite and (b) compressed graphite.
Figure 3. AFM images comparing unmodified and NiFe-modified graphite electrodes for (a) expanded graphite and (b) compressed graphite.
Molecules 29 04755 g003
Figure 4. XRD patterns of unmodified and NiFe-modified graphite electrodes for (a) expanded graphite and (b) compressed graphite.
Figure 4. XRD patterns of unmodified and NiFe-modified graphite electrodes for (a) expanded graphite and (b) compressed graphite.
Molecules 29 04755 g004aMolecules 29 04755 g004b
Figure 5. Cyclic voltammogram curves of (a) expanded graphite and NiFe-modified expanded graphite; (b) compressed graphite and NiFe-modified compressed graphite electrodes in contact with 0.1 M NaOH medium, carried out at a scan rate of 50 mV s−1 over the potential span of −0.5 to 1.8 V vs. RHE for the indicated NiFe loading levels.
Figure 5. Cyclic voltammogram curves of (a) expanded graphite and NiFe-modified expanded graphite; (b) compressed graphite and NiFe-modified compressed graphite electrodes in contact with 0.1 M NaOH medium, carried out at a scan rate of 50 mV s−1 over the potential span of −0.5 to 1.8 V vs. RHE for the indicated NiFe loading levels.
Molecules 29 04755 g005
Figure 6. (a) Electrochemical Nyquist impedance plots for the HER on unmodified and NiFe-modified electrode surfaces in 0.1 M NaOH for the potential of −200 mV vs. RHE; (bd) equivalent circuits used to fit the above processes, where Cp is the Faradaic pseudocapacitance, Rp is the Faradaic resistance, and Cdl is the double-layer capacitance (both capacitance parameters are CPE or constant phase element−modified), jointly in series with an uncompensated solution resistance, Rs. The data derived from the equivalent circuits are represented by the solid lines.
Figure 6. (a) Electrochemical Nyquist impedance plots for the HER on unmodified and NiFe-modified electrode surfaces in 0.1 M NaOH for the potential of −200 mV vs. RHE; (bd) equivalent circuits used to fit the above processes, where Cp is the Faradaic pseudocapacitance, Rp is the Faradaic resistance, and Cdl is the double-layer capacitance (both capacitance parameters are CPE or constant phase element−modified), jointly in series with an uncompensated solution resistance, Rs. The data derived from the equivalent circuits are represented by the solid lines.
Molecules 29 04755 g006aMolecules 29 04755 g006b
Figure 7. Electrochemical Nyquist impedance plots for the HER on unmodified and NiFe-modified electrode surfaces in 0.1 M NaOH for the potential of −200 mV vs. RHE.
Figure 7. Electrochemical Nyquist impedance plots for the HER on unmodified and NiFe-modified electrode surfaces in 0.1 M NaOH for the potential of −200 mV vs. RHE.
Molecules 29 04755 g007
Figure 8. Quasi-potentiostatic cathodic Tafel polarization curves for the HER, obtained from the expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified, with various NiFe loadings in 0.1 M NaOH electrolyte. The curves were recorded at a scan rate of 0.5 mVs−1 (iR-corrected).
Figure 8. Quasi-potentiostatic cathodic Tafel polarization curves for the HER, obtained from the expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified, with various NiFe loadings in 0.1 M NaOH electrolyte. The curves were recorded at a scan rate of 0.5 mVs−1 (iR-corrected).
Molecules 29 04755 g008aMolecules 29 04755 g008b
Figure 9. Linear sweep voltammetry (LSV) curves of expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified, with various NiFe loadings in 0.1 M NaOH solution, carried out with a scan rate of 0.5 mV s−1 for the HER (iR-corrected).
Figure 9. Linear sweep voltammetry (LSV) curves of expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified, with various NiFe loadings in 0.1 M NaOH solution, carried out with a scan rate of 0.5 mV s−1 for the HER (iR-corrected).
Molecules 29 04755 g009
Figure 10. Electrochemical Nyquist impedance plots for the OER on expanded graphite (a) and compressed graphite (b) and their NiFe-modified electrode surfaces in contact with 0.1 M NaOH solution for a potential of 1500 mV vs. RHE.
Figure 10. Electrochemical Nyquist impedance plots for the OER on expanded graphite (a) and compressed graphite (b) and their NiFe-modified electrode surfaces in contact with 0.1 M NaOH solution for a potential of 1500 mV vs. RHE.
Molecules 29 04755 g010
Figure 11. Quasi-potentiostatic cathodic Tafel polarization curves for the OER, obtained for the expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified with various NiFe loadings in 0.1 M of NaOH electrolyte. The curves were recorded at a scan rate of 0.5 mVs−1 (iR-corrected).
Figure 11. Quasi-potentiostatic cathodic Tafel polarization curves for the OER, obtained for the expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified with various NiFe loadings in 0.1 M of NaOH electrolyte. The curves were recorded at a scan rate of 0.5 mVs−1 (iR-corrected).
Molecules 29 04755 g011
Figure 12. Linear sweep voltammetry (LSV) curves of expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified with various NiFe loadings in 0.1 M of NaOH solution, carried out with a scan rate of 0.5 mV s−1 for the OER (iR-corrected).
Figure 12. Linear sweep voltammetry (LSV) curves of expanded graphite (a) and compressed graphite (b) electrodes, both unmodified and modified with various NiFe loadings in 0.1 M of NaOH solution, carried out with a scan rate of 0.5 mV s−1 for the OER (iR-corrected).
Molecules 29 04755 g012
Table 1. Double−layer capacitance (Cdl) values for expanded graphite and compressed graphite electrodes with varying NiFe loadings (the data are relevant to Figure S1).
Table 1. Double−layer capacitance (Cdl) values for expanded graphite and compressed graphite electrodes with varying NiFe loadings (the data are relevant to Figure S1).
Expanded GraphiteCompressed Graphite
Loading
[mg cm−2]
00.52.55.07.500.52.55.07.5
Capacitance [uF]90007700450036002500290022400750073003500
Table 2. Electrochemical impedance parameters for the HER obtained for the expanded graphite and compressed graphite electrodes, both unmodified and modified with various NiFe loadings in 0.1 M NaOH supporting solution. The results obtained here were recorded by fitting the CPE-modified Randles equivalent circuits (see Figure 6b–d) to the experimentally obtained impedance data (reproducibility is usually below 10%, χ2 = 1.56 × 10−6 to 1.74 × 10−5).
Table 2. Electrochemical impedance parameters for the HER obtained for the expanded graphite and compressed graphite electrodes, both unmodified and modified with various NiFe loadings in 0.1 M NaOH supporting solution. The results obtained here were recorded by fitting the CPE-modified Randles equivalent circuits (see Figure 6b–d) to the experimentally obtained impedance data (reproducibility is usually below 10%, χ2 = 1.56 × 10−6 to 1.74 × 10−5).
E/mVRp/Ω cm2Cp/µF cm−2Rct/Ω cm2Cdl/µF cm−2
Unmodifiedexpanded graphite
−5017.3 ± 1.13495 ± 346-7999 ± 380
−40028.3 ± 1.17863 ± 143-4693 ± 152
−45010.5 ± 0.34447 ± 320396.2 ± 7.412,351 ± 35
−7007.1 ± 0.51960 ± 29824.5 ± 0.812,815 ± 290
NiFe 0.5 mg cm−2
−5051.2 ± 3.311,879 ± 237123.1 ± 0.713,277 ± 185
−450--6.1 ± 0.13474 ± 241
NiFe 2.5 mg cm−2
−5022.2 ± 3.29257 ± 436123.8 ± 1.410,673 ± 173
−450--4.7 ± 0.01948 ± 43
NiFe 5.0 mg cm−2
−50170.6 ± 0.83970 ± 155178.0 ± 24.63022 ± 168
−450--4.4 ± 0.1848 ± 56
NiFe 7.5 mg cm−2
−50238.0 ± 0.92205 ± 37276.5 ± 4.61360 ± 16
−450--5.7 ± 0.1464 ± 27
Unmodifiedcompressed graphite
−504.3 ± 0.29783 ± 6882162.9 ± 19.12766 ± 4
−4501.4 ± 0.24066 ± 772494.3 ± 5.93763 ± 16
−65038.2 ± 3.67097 ± 200140.8 ± 6.920,992 ± 2072
−70022.4 ± 6.85242 ± 46451.7 ± 5.440,964 ± 3549
NiFe 0.5 mg cm−2
−503.6 ± 0.83923 ± 115039.1 ± 1.99741 ± 110
−450--5.7 ± 0.15501 ± 172
NiFe 2.5 mg cm−2
−50--97.5 ± 0.93154 ± 45
−450--6.7 ± 0.11457 ± 99
NiFe 5.0 mg cm−2
−50--318.7 ± 1.9976 ± 7
−450--8.2 ± 0.9676 ± 60
NiFe 7.5 mg cm−2
−50--362.8 ± 3.8841 ± 9
−450--5.7 ± 0.1417 ± 25
Table 4. Electrochemical parameters for the OER obtained for the expanded graphite and compressed graphite electrodes, both unmodified and modified with various NiFe loading levels in 0.1 M of NaOH supporting solution. The results obtained here were recorded by fitting the CPE-modified Randles equivalent circuits (see Figure 6b–d) to the experimentally obtained impedance data (reproducibility is usually below 10%, χ2 = 1.56 × 10−6 to 1.74 × 10−5).
Table 4. Electrochemical parameters for the OER obtained for the expanded graphite and compressed graphite electrodes, both unmodified and modified with various NiFe loading levels in 0.1 M of NaOH supporting solution. The results obtained here were recorded by fitting the CPE-modified Randles equivalent circuits (see Figure 6b–d) to the experimentally obtained impedance data (reproducibility is usually below 10%, χ2 = 1.56 × 10−6 to 1.74 × 10−5).
E/mVRp/Ω cm2Cp/µF cm−2Rct/Ω cm2Cdl/µF cm−2
Unmodified expanded graphite
130072.5 ± 4.21427 ± 129-4762 ± 153
140050.2 ± 2.22302 ± 144-5379 ± 173
160044.9 ± 3.53103 ± 247523.2 ± 80.78816 ± 108
180036.4 ± 1.7906 ± 2753.0 ± 3.511,069 ± 847
NiFe 0.5 mg cm−2
130035.3 ± 3.33606 ± 301-14,140 ± 195
1400--351.9 ± 26.331,733 ± 188
1500 41.1 ± 0.120,773 ± 51
1600--8.6 ± 0.116,106 ± 238
1800 3.4 ± 0.219,283 ± 2338
NiFe 2.5 mg cm−2
130033.0 ± 2.14241 ± 230-9763 ± 166
1400--2427.2 ± 469.427,367 ± 81
1500 33.6 ± 0.145,891 ± 160
1600--6.8 ± 0.144,530 ± 873
1800 2.2 ± 0.134,172 ± 3782
NiFe 5.0 mg cm−2
1300489.1 ± 15.42594 ± 18-4710 ± 148
1400--874.4 ± 23.39843 ± 39
1500 47.1 ± 0.122,100 ± 32
1600--8.5 ± 0.020,093 ± 238
1800 2.8 ± 0.14710 ± 148
NiFe 7.5 mg cm−2
1300305.5 ± 7.82362 ± 15-4038 ± 54
1400--506.6 ± 15.518,150 ± 89
1500 51.3 ± 0.121,468 ± 42
1600--7.7 ± 0.021,551 ± 143
1800 2.5 ± 0.218,765 ± 589
Unmodified compressed graphite
1300---1607 ± 6
1400---1657 ± 7
1500--6020.1 ± 227.81703 ± 6
1600--942.9 ± 9.71651 ± 9
1800--101.0 ± 0.51498 ± 19
NiFe 0.5 mg cm−2
130046.5 ± 6.211,402 ± 303-3527 ± 314
14002.1 ± 0.231,534 ± 2106-30,076 ± 2104
15000.5 ± 0.062,836 ± 883832.2 ± 0.266,025 ± 188
1600--9.5 ± 0.163,908 ± 453
1800--4.3 ± 0.167,769 ± 2389
NiFe 2.5 mg cm−2
1300248.2 ± 53.67532 ± 279--
14008.9 ± 0.96037 ± 788--
1500--34.1 ± 0.110,430 ± 335
1600--8.3 ± 0.050,340 ± 178
1800--3.4 ± 0.152,233 ± 462
NiFe 5.0 mg cm−2
1300397.5 ± 173719 ± 28-3395 ± 110
1400132.9 ± 9.830,423 ± 303-15,350 ± 45
1500--32.8 ± 0.126,630 ± 93
1600--7.1 ± 0.029,546 ± 359
1800--2.5 ± 0.130,010 ± 991
NiFe 7.5 mg cm−2
1300---2826 ± 11
1400---26,412 ± 133
1500--65.3 ± 0.319,398 ± 64
1600--8.3 ± 0.121,559 ± 456
1800--2.4 ± 0.116,191 ± 2277
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

Kuczyński, M.; Mikołajczyk, T.; Pierożyński, B.; Karczewski, J. Optimizing NiFe-Modified Graphite for Enhanced Catalytic Performance in Alkaline Water Electrolysis: Influence of Substrate Geometry and Catalyst Loading. Molecules 2024, 29, 4755. https://doi.org/10.3390/molecules29194755

AMA Style

Kuczyński M, Mikołajczyk T, Pierożyński B, Karczewski J. Optimizing NiFe-Modified Graphite for Enhanced Catalytic Performance in Alkaline Water Electrolysis: Influence of Substrate Geometry and Catalyst Loading. Molecules. 2024; 29(19):4755. https://doi.org/10.3390/molecules29194755

Chicago/Turabian Style

Kuczyński, Mateusz, Tomasz Mikołajczyk, Bogusław Pierożyński, and Jakub Karczewski. 2024. "Optimizing NiFe-Modified Graphite for Enhanced Catalytic Performance in Alkaline Water Electrolysis: Influence of Substrate Geometry and Catalyst Loading" Molecules 29, no. 19: 4755. https://doi.org/10.3390/molecules29194755

APA Style

Kuczyński, M., Mikołajczyk, T., Pierożyński, B., & Karczewski, J. (2024). Optimizing NiFe-Modified Graphite for Enhanced Catalytic Performance in Alkaline Water Electrolysis: Influence of Substrate Geometry and Catalyst Loading. Molecules, 29(19), 4755. https://doi.org/10.3390/molecules29194755

Article Metrics

Back to TopTop