Next Article in Journal
Visible-Light-Driven CO2 Photoreduction Using Ruthenium (II) Complexes: Mechanisms, Hybrid Systems and Recent Advances
Previous Article in Journal
Magnetic Cross-Linked Enzyme Aggregates of Glycoside Hydrolase: An Efficient and Stable Biocatalyst for Icaritin
Previous Article in Special Issue
Recent Advances in Energy-Related Materials—Special Issue Preface
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pd-Modified CoP and CoFeP Catalysts as Efficient Bifunctional Catalysts for Water Splitting

by
Huma Amber
*,
Aldona Balčiūnaitė
,
Virginija Kepenienė
,
Giedrius Stalnionis
,
Zenius Mockus
,
Loreta Tamašauskaitė-Tamašiūnaitė
* and
Eugenijus Norkus
Department of Catalysis, Center for Physical Sciences and Technology (FTMC), LT-10257 Vilnius, Lithuania
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1035; https://doi.org/10.3390/catal15111035 (registering DOI)
Submission received: 6 October 2025 / Revised: 28 October 2025 / Accepted: 28 October 2025 / Published: 2 November 2025
(This article belongs to the Special Issue Recent Advances in Energy-Related Materials in Catalysts, 3rd Edition)

Abstract

Developing highly efficient and stable electrocatalysts from inexpensive and earth-abundant elements represents a significant advancement in overall water splitting (OWS). This study focuses on the synthesis and evaluation of palladium-modified cobalt–phosphorus (PdCoP) and cobalt–iron–phosphorus (PdCoFeP) coatings for use as electrocatalysts in hydrogen evolution (HER), oxygen evolution (OER) and overall water splitting (OWS) in alkaline media. A facile electroless plating method is adopted to deposit the CoP and CoFeP coatings onto a copper surface (Cu sheet), with sodium hypophosphite (NaH2PO2) acting as the reducing agent. Pd crystallites were incorporated on CoP and CoFeP coatings using the galvanic displacement method. This study details morphological characterization (using SEM, EDX, and XRD), as well as electrochemical activity testing, for both HER and OER using linear sweep voltammetry (LSV) at different temperatures. The stability of the catalysts for HER was evaluated using chronoamperometry (CA) and chronopotentiometry (CP). The results show that the Pd-modified CoFeP and CoP catalysts exhibited lower overpotentials of 207 and 227 mV, respectively, for HER and 396 mV for OER at a current density of 10 mA cm−2 compared to the unmodified CoFeP and CoP catalysts. The innovation achieved in this study lies in combining a facile, low-cost deposition method (electroless plating followed by galvanic displacement) with a novel, highly effective ternary composition (PdCoFeP) that exploits synergistic electronic and morphological effects to achieve superior bifunctional performance for alkaline OWS, achieving a low cell voltage of 1.69 V at a current density of 10 mA cm−2. Overall, this research demonstrates that these synthesized materials are promising candidates for sustainable and economical hydrogen production.

1. Introduction

The increasing demand for clean and sustainable energy sources has been driven by two key factors: the overuse of fossil fuels and the rapid growth of environmental problems [1,2,3]. Hydrogen has been regarded as a potentially viable alternative fuel, primarily due to its clean and sustainable characteristics. Currently, the predominant method for producing hydrogen is fossil fuel reforming, which results in carbon dioxide emissions. On the other hand, water electrolysis is regarded as the most environmentally friendly technique for hydrogen production, and it can mitigate the intermittent nature of renewable electricity [4,5,6]. Electrochemical water splitting, comprising a hydrogen evolution reaction (HER) at the cathode and an oxygen evolution reaction (OER) at the anode, has been identified as a promising technology for large-scale hydrogen production. This is due to low energy consumption and high hydrogen purity, among other advantages [7,8,9]. As Van Troost Wijk and Deiman first reported in 1789, water electrolysis requires a large overpotential, and this is still true more than 200 years later due to the relatively sluggish kinetics of both cathodic HER and anodic OER. It is well established that the thermodynamic potential of water electrolysis is 1.23 V [10,11]. However, to achieve a practical rate of water dissociation, it is necessary to apply an overpotential to the water electrolyzer. It has been demonstrated that both HER and OER demand significant overpotentials to surpass the kinetic barriers that arise from the high activation energies necessary for forming reaction intermediates on the electrode surface [12,13,14]. Furthermore, these overpotentials need to drive the reaction at a specific current density, which is typically a geometric current density of 10 mA cm−2. Using an active catalyst is greatly important in reducing the overpotential requirements of these two half-reactions, thereby increasing the reaction rate. Ideally, a catalyst should exhibit high activity (large current density at low overpotentials) and durable performance over an extended period [15]. An efficient water splitting process necessitates using electrocatalysts characterized by high efficiency, stability, and corrosion resistance. These catalysts must facilitate both HER at the cathode surface and OER at the anode surface, with both reactions occurring simultaneously and exhibiting high rates of reaction [16,17]. This will enhance efficiency, lower the cost of developing an electrolysis system, and simplify the system [10,18,19,20,21]. Various catalysts have been synthesized to enhance the reaction rate and reduce the overpotential for HER and OER. Currently, noble metals and noble metal oxides, including Pt [22,23], RuO2 [24], and IrO2 [25], are regarded as the most effective catalysts for both HER and OER. However, the high cost and scarcity of noble metal catalysts have raised concerns regarding economic viability. Consequently, a strong research focus has emerged on developing low-cost, high-efficiency, non-precious electrocatalysts, which serve as an alternative to precious-based materials and exhibit high activity towards OER and HER [26,27,28,29,30].
Today, earth-abundant transition metal compounds, principally Co-, Fe-, Mn-, and Ni-based materials formed either as oxides or hydroxides, are utilized for OER. These metals are often combined with other non-metallic elements, including C, N, S, P, Se, and B, to form binary or ternary compositions. These compositions are employed in both HER and OER [31,32,33,34,35,36,37,38]. Significant research has been dedicated to investigating the potential of these compounds as electrocatalysts, given to their remarkable durability, particularly cobalt-based compounds such as sulfides [39,40,41], borides [42], (oxy)hydroxides [43,44,45] and phosphides [46,47], which have been evidenced to be remarkable electrolyzed water catalysts. Researchers have highlighted transition-metal phosphides (TMPs) as a promising class of catalysts for electrolyzing water. TMPs such as Co-P [48,49], Fe-P [50,51], Ni-P [52,53], and Mo-P [54,55] have been identified as highly attractive catalysts due to the superior conversion of their oxidation states, their cost-effectiveness, and their high activity. A significant number of studies have been conducted on the catalytic role of transition metal phosphide in HER. These studies have resulted in the development of varied of catalysts that exhibit high activity in hydrogen evolution. This enhanced activity has been achieved through adjusted catalyst morphology and surface modification [56]. Given the calculation results for density functional theory (DFT), it is hypothesized that the P atom in the phosphide catalyst has electronegativity, facilitating the adsorption and capture of protons, and enabling part of the charge of Co to be transferred to the P atom. Modifying the electronic structure has been demonstrated to be an effective strategy for reducing the Gibbs hydrogen adsorption free energy, resulting in the enhanced catalytic activity of hydrogen evolution [57,58]. A considerable number of monometallic phosphides are employed in electrolytic water and have demonstrated considerable success, exhibiting high catalytic activity during the reaction. Among these, the cobalt phosphide catalyst has been shown to exhibit low overpotential, long-term stability, and other advantageous properties, and it is regarded as the most promising catalyst candidate for HER [59,60,61].
Despite the high catalytic activity of the cobalt phosphide catalyst in hydrogen evolution reactions, a comparative analysis reveals a performance deficit for comprehensive performance when benchmarked against precious metal Pt-based catalytic materials. To enhance the activity of the catalyst, researchers have pursued a multifaceted approach involving the doping of one or more metal atoms into the single-metal phosphating process. This strategy aims to regulate the electronic structure and optimize surface properties, achieving enhanced catalytic functionality. For instance, Yang et al. synthesized Fe-doped three-dimensional Co-based phosphate FexCo3−x(PO4)2 on a copper wire using the electrodeposition method as an electrochemical catalyst. In an alkaline solution, a current density of 10 mA cm−2 required only 48.9 mV of overpotential for HER, and the catalytic performance was maintained after 100 h of operation. This study, thus, demonstrates the crucial role played by Fe doping in enhancing the catalytic performance of the FexCo3−x(PO4)2 catalyst for HER [62]. As research has progressed, it has become evident that transition metal phosphide exhibits not only exceptional performance for HER but also noteworthy catalytic performance for OER. The catalytic effect of cobalt-based phosphide materials on OER is attributable to the adsorption of OOH species by positively charged cobalt ion centers, while negatively charged P centers facilitate surface oxygen removal. It is therefore anticipated that transition metal phosphating will function effectively as a bifunctional catalyst for both HER and OER. The development of bifunctional electrocatalysts within the same electrolyte has attracted significant attention due to their numerous advantages, including the simplified fabrication process, the reduced cost, and enhanced hydrogen productivity. The majority of catalysts composed of Co- or Ni-based materials can catalyze a single reaction (OER or HER). Consequently, it is essential to identify an efficient bifunctional catalyst that exhibits robust stability for the overall water splitting process in an alkaline environment [63]. Incorporating transition metal Fe into cobalt phosphide has been demonstrated to enhance its oxygen evolution activity, while simultaneously having an optimizing effect on hydrogen evolution electrocatalysis [64,65]. The enhanced HER/OER performance of the FeCo-P catalyst is primarily attributable to the synergistic effect of Fe, Co, and P, as well as the robust interaction between ions [66].
The majority of bimetallic phosphides have been synthesized through utilizing hydrothermal, solid-phase, and gas-phase phosphating methodologies [67]. Even though these methodologies have been employed to prepare transition bimetallic phosphides of various morphological structures, which have been shown to have superior properties, it is still necessary to further simplify the preparation of bimetallic phosphides to meet the requirements of their practical applications [68]. As indicated in the relevant literature, the conventional phosphating reaction still has many disadvantages, including high reaction temperatures, the requirement for inert gas protection, and the usage of hazardous phosphorus resources [69]. These methods demand considerable energy input and rely on complicated apparatus. Using straightforward, economical, and effective techniques for catalyst fabrication has the potential to decrease the overall cost of hydrogen production through water electrolysis [70]. Consequently, electroless deposition has emerged as a highly efficient, cost-effective, eco-friendly, and straightforward catalyst preparation method. Electroless plating, also referred to as chemical or autocatalytic plating, is a non-galvanic plating method involving multiple reactions occurring simultaneously in an aquatic solution [71], and it does not require using complex equipment or hazardous power sources [72].
This study investigates the fabrication and evaluation of palladium-modified cobalt–phosphorus (PdCoP) and cobalt–iron–phosphorus (PdCoFeP) coatings as electrocatalysts for HER and OER in alkaline media. The CoP and CoFeP coatings were deposited on the copper surface using a straightforward and highly efficient electroless metal plating technique, using sodium hypophosphite (NaH2PO2) as the reducing agent. Incorporating of Pd crystallites on the CoP and CoFeP coatings using the galvanic displacement method has been shown to result in significant improvements in catalytic performance. Evaluations and characterizations of PdCoP and PdCoFeP were conducted to assess their chemical state, catalytic activity, and stability during HER and OER. Moreover, the catalytic mechanisms of PdCoP and PdCoFeP catalysts for overall water splitting were investigated. The findings of this research suggest that the synthesized catalysts have the potential to be promising candidates for utilization in alkaline media for performing water splitting, which could result in a more environmentally sustainable and economical method of hydrogen production.

2. Results and Discussion

2.1. Coatings, Microstructure and Morphology Studies

The primary objective of this study was to develop an effective catalyst for HER and OER in an alkaline medium (1.0 M KOH), employing a straightforward methodology. Consequently, Co-based coatings containing Co, Fe and P were deposited on Cu substrates via a facile electroless deposition process, where sodium hypophosphite was utilized as the reducing agent. The galvanic displacement method was used to incorporate palladium crystallites into the CoP and CoFeP coatings, and this enhanced the catalytic activity of the resulting coatings. Scanning electron microscopy (SEM) and energy-dispersive X-ray analysis (EDX) were employed to investigate the surface morphology and composition of the coatings. Figure 1 illustrates the SEM images of the fabricated Co-based coatings with different compositions such as CoP (a), CoFeP (b), PdCoP (c), and PdCoFeP (d). The presence of Pd, Co, Fe, and P in the coatings is confirmed by the corresponding EDX spectra of all the samples. The results of the EDX are outlined in Table 1, which shows that the fabricated CoP, CoFeP, PdCoP, and PdCoFeP electrocatalysts demonstrated Pd contents ranging from 3.54 to 8.36 at%, Co contents ranging from 64.08 to 90.22 at%, Fe contents ranging from 9.96 to 12.49 at%, and P contents ranging from 6.26 to 15.07 at%.
SEM analysis demonstrates distinct morphological variations induced by compositional modifications in the electrolessly deposited coatings. The CoP coating exhibited a compact and relatively homogeneous surface composition, characterized by densely packed grains [72,73], indicative of stable nucleation and uniform growth (Figure 1a). It is noteworthy that incorporating Fe into the CoFeP structure results in a significant structural transformation, creating a porous and heterogeneous morphology [74]. This substantial modification underscores the considerable impact of Fe on nucleation kinetics, surface energy, and phosphorus incorporation during electroless deposition (Figure 1b). Adding Pd to CoP leads to the surface becoming rougher and more fragmented. This is characterized by the presence of finer clusters and increased grain boundaries, indicative of an enhanced nucleation density.
The PdCoFeP coating exhibits the most irregular and nanostructured surface, characterized by textured characteristics and fine crystallites (Figure 1c,d). This observation indicates a synergistic effect of Pd and Fe in inducing high nucleation rates and complex growth pathways. The obtained results demonstrate the extreme sensitivity of electroless deposition processes to even minor compositional variations, which can dramatically alter the resulting microstructure.
Figure 2a shows the XRD pattern of the CoP catalyst. The diffraction patterns of the CoP revealed crystalline phases associated with orthorhombic CoP and metallic Co. The CoP peaks were detected at approximately 2θ of 45.1° and 48.1°. These peaks corresponded to the (021) and (121) planes, respectively, and matched with COD file No. 9,008,928 (a = 5.59900 Å, b = 5.07600 Å, c = 3.28100 Å). The 2θ peaks at approximately 41.7°, 44.9°, 47.7°, 62.9°, 76.2°, 84.9° and 93.1° corresponded to the metallic Co phase (COD No. 1512501, a—2.49680 Å, b = 2.49680 Å, c = 4.03081 Å), which were indexed to the (100), (002), (101), (102), (110), (103) and (112) crystal planes, respectively.
Following the incorporation of Fe into the CoP structure, the resulting CoFeP coating exhibits diffraction peaks at 2θ values of 44.97°, 65.48° and 82.96°, corresponding to the crystal planes of (110), (200), (211) and (220), respectively, of a CoFe-type solid solution phase (COD 1524319) (Figure 2b). Compared with pure CoP, a slight shift in peak positions is observed, confirming the successful incorporation of Fe atoms into the CoP lattice and the formation of a CoFeP alloy-like structure. Reflections corresponding to metallic Co (COD 1512501) and metallic Fe (COD 1512555) are also detected (Figure 2b). Introducing Fe induces lattice distortion and modifies the crystal structure, which could enhance electronic conductivity and catalytic active site density.
For PdCoP (Figure 2c), the additional reflections observed at 2θ = 40.24°, 46.80°, 68.34°, 82.39° and 86.93° correspond to the (111), (200), (220), (311) and (222) crystallographic planes of face-centered cubic Pd (COD 1011105), respectively, which confirms the presence of palladium on the surface. Reflections assigned to metallic Co (COD 1512501) are also detected. Incorporating Pd does not significantly alter the CoP crystal structure, as the main CoP peaks remain unchanged after Pd deposition. However, the slight broadening and reduction in intensity of the CoP peaks compared to pure CoP suggest the occurrence of surface modification and possible lattice strain due to Pd deposition. The presence of both CoP and Pd peaks indicates the formation of a bimetallic heterostructure (Pd–CoP) rather than a new intermetallic compound.
Figure 2d shows the XRD pattern of PdCoFeP, which exhibits diffraction peaks corresponding to the CoFe phase (COD 1524319), metallic Co (COD 1512501), metallic Fe (COD 1512555), and Pd (COD 1011105). These results confirm the successful deposition of Pd crystallites onto the CoFeP surface via galvanic displacement, with no alteration in the original CoFeP crystal structure.
The structural evolution from CoP to CoFeP and subsequently to PdCoP and PdCoFeP demonstrates a clear progression in crystal structure and surface composition. The CoP coating exhibits well-defined peaks corresponding to orthorhombic CoP, indicating high crystallinity with a small amount of residual metallic Co. After introducing Fe, the diffraction peaks undergo a slight shift and broadening, confirming the successful incorporation of Fe into the CoP lattice to form a CoFeP alloy-like structure. This Fe substitution introduces lattice distortion, enhancing electronic conductivity and potentially creating more catalytically active sites. Following Pd deposition via galvanic displacement, the characteristic CoP or CoFe peaks remain largely unchanged, indicating that Pd incorporation does not disturb the bulk CoP or CoFeP structures. However, new reflections assigned to fcc Pd appear, verifying the formation of Pd nanoparticles on the CoP or CoFeP surfaces. The sequential modification in the order CoP → CoFeP → PdCoP → PdCoFeP leads to a progressive structural and compositional evolution. Fe alloying alters the bulk crystal lattice, while Pd addition tunes the surface composition and introduces additional active sites. This synergistic structural tuning is expected to improve the electronic structures, surface reactivities, and catalytic performances of the PdCoP and PdCoFeP composite catalysts.

2.2. Electrocatalytic Activity Towards HER

The electrocatalytic performances of the prepared catalysts for the hydrogen evolution reaction (HER) were investigated by recording linear sweep voltammograms (LSVs) in 1.0 M KOH solution at a potential scan rate of 5 mV·s−1 at room temperature (Figure 3a). Summarized parameters for HER on the investigated catalysts in alkaline media at 25 °C are given in Table 2. The corresponding overpotential values at 10, 20, 50, and 100 mA cm−2 are given in Figure 3b. Among the catalysts investigated, the CoP catalyst exhibited the lowest overpotential value of 239 mV for the HER compared to CoFeP (245 mV), PdCoP (250 mV), and PdCoFeP (259 mV), attaining a current density of 10 mA cm−2 at 25 °C (Figure 3a and Table 2). However, a higher current density is delivered on PdCoFeP, indicating that more efficient hydrogen generation occurs under higher load conditions. Moreover, the overpotential of PdCoFeP at 100 mV cm−2 is 385 mV, which is smaller than those of PdCoP (420 mV), CoFeP (428 mV), and CoP (458 mV) (Figure 3b). These improvements are attributed to the synergistic interactions between Co, Fe, P, and Pd, which enhance electron transfer and facilitate HER kinetics.
The reaction kinetics and mechanism of the catalysts can be evaluated based on the Tafel slopes. The Tafel slopes were determined using the following equation [75]:
η = b · log j/j0
where η is the overpotential, b is the Tafel slope, j is the experimental current density, and j0 is the exchange current density. The plot of η versus log j represents the Tafel slope. In alkaline media, the HER process involves three main steps, as shown in Equations (2)–(4) [75]:
* + H2O + e ↔ *Hads + OH  (Volmer step, 120 mV dec−1)
*Hads + e + H2O ↔ H2 + OH + * (Heyrovsky step, 40 mV dec−1)
2 *Hads ↔ H2 + *  (Tafel step, 30 mV dec−1)
Hads denotes the H2 adsorbed to the metal sites, where * represents the metal sites. Figure 3c displays the plots of η versus log j for each catalyst. The Tafel slopes calculated were within a range of approximately 76 to 83 mV dec−1 (Table 2), whereas the CoFeP and PdCoP catalysts had the lowest Tafel slope (ca. 76 mV dec−1) relative to CoP (78.1 mV dec−1) and PdCoFeP (82.8 mV dec−1) (Figure 3c). This finding indicates that the HER may occur through the Volmer–Heyrovsky mechanism, in which water molecules or H2 adsorbs onto an electrode to generate MHads species.
The activation energy (Ea) is a critical factor in evaluating the performance of the catalyst. This energy can be estimated by employing the LSV curves at varying temperatures. The obtained HER LSVs for each catalyst at temperatures from 25 to 55 °C are shown in Figure S1 and Table S1 (Supplementary Material). The results obtained outline that the HER activity of the CoP, CoFeP, PdCoP, and PdCoFeP catalysts is enhanced as the electrolyte temperature increases, whereas the overpotentials required to achieve the current density of 10 mA cm−2 decrease with the increase in temperature for each catalyst (Table S2, Supplementary Material). Furthermore, the obtained Tafel slope values range from approximately 72 to 94 mV dec−1. These results indicate that the HER kinetic process is closely related to temperature. For determining Ea, the current values at different temperatures recorded on the catalysts at −0.40 V (vs. RHE) were extracted from the LSV profiles (Figure S1, Supplementary Material). The data were plotted as log(i) against 1/T for the CoP, CoFeP, PdCoP, and PdCoFeP catalysts and are shown in Figure 3d. The Ea was calculated using the slope of the obtained straight lines for each catalyst from the Arrhenius equation [76]. It was determined that the PdCoP, CoFeP, CoP, and PdCoFeP catalysts had Ea values of 4.7, 6.0, 11.3, and 11.4 kJ mol−1, respectively (Figure 3d). The obtained Ea values were significantly lower against Pt/C (20 wt%) benchmark, which were reported as follows: benchmark—16 kJ mol−1 [77]; Pd100—20.86 kJ mol−1 [77]; Pd5Co95—21.75 kJ mol−1 [77]; Pd/C—31 kJ mol−1 [77]; CoOx—29.79 kJ mol−1 [78]. The low activation energies implied that a small reaction kinetic barrier is required for the reaction to proceed efficiently. The reduction in Ea can be explained through several factors related to the catalyst’s composition, electronic structure, and morphological properties. Incorporating of Fe and subsequent modification with Pd crystallites reduces the Ea of CoP due to the synergistic interactions between Fe, Co, and Pd. These interactions are crucial because they enhance electron transfer and facilitate the kinetics of the HER.
EIS spectra were obtained on the PdCoFeP and CoFeP samples to investigate the influence of modification with Pd on HER activity. The Nyquist plots shown in Figure 4 are characterized by a broad semicircle, while at low frequencies (1.5–0.5 Hz), a small but pronounced inductive response emerges as Zim values become positive.
This phenomenon typically signals the slow relaxation of adsorbates in multistep reactions such as corrosion or electrocatalysis [79], suggesting that the electrode surface may undergo some restructuring or phase transformation during operation. Therefore, the impedance spectra were fitted to an equivalent circuit model that included an inductive element to account for the low-frequency inductive response [80]. The constant phase element (CPEdl) values were recalculated into effective capacitance Ceff vias Brug et al.’s formula [81], and the obtained values for CoFeP and PdCoFeP were 3.5 mF cm−2 and 7.4 mF cm−2, respectively. Additionally, the associated charge transfer resistance (Rct) of CoFeP (5.5 Ω cm2) was two times larger than that of the more electrochemically active PdCoFeP (3.7 Ω cm2). These observations correspond well to the HER activity results presented in Figure 3. The palladium-decorated film exhibits better HER activity because it has a significantly larger electrochemically active surface area and faster HER kinetics, as demonstrated by the larger Ceff and lower Rct values.

2.3. Electrocatalytic Activity Towards OER

The activity of the Co-based catalysts in the OER was also evaluated. The obtained data are given in Figure 5 and Table 3. In addition, the LSV was recorded on the Cu substrate for comparison. Higher OER current densities were obtained on the CoP, CoFeP, and Pd-modified CoP and CoFeP catalysts compared to the Cu substrate (Figure 5a), indicating that they demonstrated high activity during the OER. As shown by the LSVs recorded for the aforementioned catalysts in a 1 M KOH solution, the lowest OER overpotential of 396 mV was observed for the Pd-modified CoP and CoFeP, achieving a current density of 10 mA cm−2, compared to the unmodified CoP (431 mV) and CoFeP (435 mV), and Cu (645 mV) (Table 3). Lower Tafel slope values (in the range of 70.5 to 77.8 mV dec−1) were obtained for PdCoFeP, CoFeP, and PdCoP catalysts compared to the CoP catalyst (120.3 mV dec−1). These results indicated superior OER kinetics (Table 3). Higher current density values for OER were obtained on the Pd-modified CoP and CoFeP catalysts compared to the unmodified CoP and CoFeP catalysts (Figure 5b). The overpotential of PdCoFeP at 100 mA cm−2 is 549 mV, which is smaller than those of PdCoP (595 mV), CoFeP (747 mV), and CoP (778 mV). The obtained OER LSVs for each catalyst at temperatures ranging from 25 to 55 °C are shown in Figure S2 and Table S2 (Supplementary Material). The OER activity of the CoP, CoFeP, PdCoP, and PdCoFeP catalysts also increases as the temperature of electrolyte increases. An increase in temperature for each catalyst also results in a decrease in the overpotential values, which need to attain 10 mA cm−2 (Table S2, Supplementary Material). The obtained Tafel slope values, which also decrease with the increase in temperature for each catalyst, range from approximately 120 to 95 mV dec−1 for CoP, from 74 to 60 mV dec−1 for CoFeP, from 78 to 77 mV dec−1 for PdCoP, and from 70 to 65 mV dec−1 for PdCoFeP at a temperature from 25 and 55 °C, respectively (Table S2, Supplementary Material). To determine Ea for the OER, the current value and temperature value of the catalysts at 1.8 V (vs. RHE) were extracted from the LSV profiles (Figure S2, Supplementary Material). The PdCoFeP, PdCoP, CoP, and CoFeP catalysts have Ea values of 11.8, 12.2, 20.7 and 21.2 kJ mol−1, respectively (Figure 5d). In addition, the PdCoFeP has the lowest activation energy, i.e., 11.8 kJ mol−1, implying that it has the smallest reaction kinetic barrier.
A comparison of the OER performances of the investigated catalysts with those of previously reported Co-P-based and most advanced noble-metal catalysts is shown in Table 4. The obtained overpotential values of 396 mV for the PdCoP/Cu and PdCoFeP/Cu catalysts are lower compared to CoOx (423 mV) [82], Co2P2O7 (490 mV) [83], CoP hollow polyhedron (400 mV) [84], and reduced mesoporous Co3O4 nanowires (400 mV) [85] but higher compared to the overpotential values for CoP-MNA/Ni foam (390 mV) [86], CoP film (350 mV) [18], CoSi-P (309 mV) [87], CoP (300 mV) [88], Mn-CoP (288 mV) [89], RuO2/CF catalyst (360 mV) [90], RuO2 on NF (290 mV) [91], IrO2 commercial (339 mV) [92], Ir/C (254 mV) [93], Ni0.3Co2Fe-P (290 mV) [94], Co2Fe-P (367 mV) [94], CuO@FeCoP/CF (240 mV) [95], Mo-CoP@C (280 mV) [96], and CoFe@Co2C@Co-0.2/NF (246 mV) [97].

2.4. Determination of Electrochemically Active Surface Areas

The electrochemically active surface areas (ECSAs) of the unmodified and Pd-modified CoP and CoFeP catalysts were determined from the measurements of the electrochemical double-layer capacitance (Cdl). The CV curves were recorded at different scan rates in the non-Faradaic region (Figure 6a–d), which was followed by calculating the slope of the curve obtained by plotting the differences in the anodic and cathodic current versus the scan rate (Figure 6e). The Cdl was 2458.5, 11,652.2, 10,858.9 and 56,291.2 µF for CoP, CoFeP, PdCoP and PdCoFeP (Figure 6e), whereas the calculated ECSA values were 61.5, 291.3, 271.5 and 1407.3 cm2 for CoP, CoFeP, PdCoP and PdCoFeP, respectively.
The high ECSA of PdCoFeP was due to the higher number of active sites, which contributed to the higher OER electrocatalytic activity. The incorporation of Fe into the CoP coating, followed by its modification with Pd crystallites, allowed it achieve the lowest OER overpotential, Tafel slope, and Ea, which suggests that factors beyond active site availability influence OER performance. Other factors, such as electronic structure and conductivity, as well as a high ECSA, also significantly influence catalyst performance. The catalyst with 8.36 at% Pd had the best activity during the OER.

2.5. Investigation of Stability of Catalysts

Additionally, the stabilities of the most promising Pd-modified CoP and CoFeP catalysts were investigated using chronoamperometry by recording the chronoamperometric curves at a constant potential of −0.26 V for 10 h on the PdCoP and PdCoFeP catalysts (Figure 7). The PdCoFeP/Cu catalyst possessed excellent stability, with current retention of 96.02% over 10 h compared to the PdCoP/Cu catalyst.
Figure 8 shows the chronopotentiometric curves recorded on the PdCoP (a) and PdCoFeP (b) catalysts in a 1 M KOH solution at a constant current density of −10 mA cm−2 and 25 °C. It is evident that the potential required to attain −10 mA cm−2 slightly increased from −0.1945 V to −0.2131 V for PdCoP (Figure 8a) and from −0.1551 V to −0.1740 V for PdFeCoP (Figure 8b) over 10 h, demonstrating the good stability of both catalysts.

2.6. Investigation of Overall Water Splitting

To confirm the bifunctional activity of the unmodified and Pd-modified CoP and CoFeP catalysts, a two-electrode water electrolysis cell was constructed using two of the same investigated catalysts as the anode and the cathode. Figure 9 shows the overall catalytic performance of a two-electrode alkaline electrolyzer cell configuration using the same catalysts for both the anode and the cathode: CoP‖CoP, CoFeP‖CoFeP, PdCoP‖PdCoP, and PdCoFeP‖PdFeCoP.
The developed CoP, CoFeP, PdCoP and PdFeCoP electrodes exhibited cell potentials of 1.86 V, 1.81 V, 1.78 V and 1.69 V, respectively, at 10 mA cm−2 (Figure 9), comparable to the cell potential values recorded for Pt/C‖IrO2 (1.71 V), Pt/C‖Pt/C (1.83 V) [98], CoP/rGO-400‖CoP/rGO-400 (1.70 V) [98], Co(OH)2@NCNTs@NF‖Co(OH)2@NCNTs@NF (1.72 V) [99], Co-P/NC/CC‖Co-P/NC/CC (1.77 V) [100], and Co-P/NC-CC‖Co-P/NC-CC (1.95 V) [100] as well as for other studies reported in the literature (Table 5).
The energy efficiencies of the cells were calculated. The energy efficiencies of the CoP, CoFeP, PdCoP and PdCoFeP cells were 66.13%, 67.96%, 69.10%, and 72.78%, respectively. These catalysts appear to be promising candidates for practical overall water splitting applications.

3. Materials and Methods

3.1. Chemical Reagents

Cobalt(II) sulfate CoSO4·7H2O (99.5%, Chempur, Piekary Śląskie, Poland), sodium hypophosphite (NaH2PO2, 97%, Alfa Aesar, Kandel, Germany), glycine (NH2CH2COOH, 99%, Chempur, Piekary Śląskie, Poland), palladium(II) chloride PdCl2 (59.5%, Alfa Aesar, Ward Hill, MA, USA), hydrochloric acid (HCl, 35–38%, Chempur, Piekary Śląskie, Poland), potassium hydroxide KOH (98.8%, Chempur, Piekary Śląskie, Poland), Cu sheets (99%, Goodfellow Cambridge Limited, Huntingdon, UK), and calcium magnesium oxide, known as “Vienna Lime” (50–100%, Kremer Pigments GmbH & Co. KG, Aichstetten, Germany) were used. All the chemicals were analytical grade and used without any further purification. The electrolytes were prepared using deionized water from a Millipore Milli-Q Ultra system with a resistivity of 18.2 MW-cm or higher.

3.2. Preparation of Catalysts

CoP and CoFeP catalysts were deposited onto Cu substrates (1 × 1 cm) via an electroless deposition method. The process consisted of the following steps:
  • Pretreatment: The Cu substrate was pretreated with calcium magnesium oxide and subsequently rinsed with deionized water (DI).
  • Decapitation: The substrate was immersed in a HCl:H2O (1:1, v/v) solution for 1 min at 25 °C, followed by rinsing with DI.
  • Activation: The cleaned Cu substrate was activated via immersion in a 0.5 g L−1 PdCl2 solution for 1 min at 25 °C, and then rinsed with DI.
  • Electroless deposition:
    (i)
    For CoP deposition, the Pd-activated Cu substrate was immersed in a plating bath containing 0.1 M CoSO4, 0.6 M glycine, and 0.75 M NaH2PO2 (pH = 11) for 30 min at 60 °C.
    (ii)
    For CoFeP deposition, the same plating bath was used with the addition of 0.01 M FeSO4.
Finally, Pd crystallites were deposited on the CoP and CoFeP coatings via galvanic displacement, achieved by immersing the samples in a 1 mmol L−1 PdCl2 solution for 1 min at 25 °C, followed by rinsing with DI.

3.3. Characterization of Catalysts

An FEI Helios Nanolab 650 (Hillsboro, OR, USA) dual beam system with an energy dispersive X-ray (EDX) spectrometer INCA Energy 350 (Oxford Instruments, Abingdon, Oxfordshire, UK) and an X-Max 20 mm2 (Oxford Instruments, Abingdon, Oxfordshire, UK) detector was used to observe the surface morphology via scanning electron microscopy.
X-ray diffraction patterns were recorded using an X-ray diffractometer D2 Phaser using CuKα radiation (Bruker, Karlsruhe, Germany). A step-scan mode was used in a 2θ range from 10° to 100° with a step length of 0.04° and a counting time of 1 s per step. The samples were crushed prior to measurement. For XRD measurements, the CoP, CoFeP, PdCoP and PdCoFeP catalysts were deposited on the glass surface.

3.4. Evaluation of Catalysts Activity for HER and OER

The electrocatalytic activities of the prepared catalysts towards the HER and OER were evaluated via linear sweep voltammetry (LSV). This was conducted using a PGSTAT100 potentiostat/galvanostat (Metrohm Autolab B. V., Utrecht, The Netherlands) with a standard three-electrode electrochemical cell setup. The fabricated catalysts with geometric areas of 2 cm2 were employed as working electrodes. A Pt sheet was used as a counter electrode, and an Ag/AgCl (3 M KCl) electrode was used as a reference. All the reported potential values reported in this study were converted to the reversible hydrogen electrode (RHE) scale using the following equation:
ERHE = Emeasured + 0.059·pH + EAg/AgCl (3 M KCl)
where EAg/AgCl (3 M KCl) = 0.210 V.
Linear sweep voltammograms (LSVs) were recorded in an N2-saturated 1 M KOH solution at different temperatures from the open circuit potential (OCP) to ca. −0.5 V (vs. RHE) for the HER and from the OCP to ca. 2.1 V (vs. RHE) for the OER at a potential scan rate of 5 mV s−1. Additionally, to evaluate the long-term stability of the fabricated catalysts, chronoamperometric curves were recorded at a constant potential of −0.26 V in a 1.0 M KOH solution at 25 °C for 10 h, while chronopotentiometric curves were recorded at a constant current density of −10 mA cm−2 at 25 °C for 10 h.
To determine Ea, the current values from the LSVs at different temperatures at potentials of −0.40 V (vs. RHE) for the HER and 1.8 V (vs. RHE) for the OER were extracted from the LSV profiles presented in Supplementary Material (Figures S1 and S2). The data were plotted as log(i) against 1/T, resulting in a straight line with a gradient of −Ea/2.3R. Ea was calculated using the slope of this line according to Equation (6):
∂logi/∂(1/T) = −Ea/2.3R
where i is the current (mA), T is the temperature (K), and R is the gas constant (8.314 J mol−1 K−1) [76].
The HER and OER current densities presented in this study have been scaled to the geometric area of the catalysts.
To evaluate the ECSA of the catalysts, the double-layer capacitance (Cdl) was determined by recording the CVs at various scan rates under the non-faradaic region followed by the calculation of the slope of the curve obtained by plotting the difference in the anodic and cathodic current against the scan rate [101,102,103]. From the CVs, the charging current, Ic, of the electrodes at each scan rate was determined via Equation (7):
Ic [A] = (IanodicIcathodic)OCP
The Cdl values were evaluated by plotting a graph of the charging current vs. the scan rate and calculating the slope, as shown by Equation (8):
Slope = Cdl [F] = ΔIC [A]/Δν [V s−1]
Then, the ECSA values were calculated using the specific capacitance (Cs) of 40 μF cm−2 [101,102,103] and Equation (9):
ECSA [cm2] = Cdl [μF]/Cs [μF cm−2]
A two-electrode water electrolysis cell was constructed using two of the same Co-based catalysts as the anode and the cathode. The energy efficiency of the cell was calculated using the following Equation (10):
ηelectrolyzer = Eth/Ve at j
where Eth = 1.23 V; Ve at j is the input voltage required to drive the electrolysis at the current density of interest. The energy efficiency calculated in this study was obtained at j = 10 mA cm−2.
Electrochemical impedance spectroscopy (EIS) spectra were obtained on a Zahner Zennium potentiostat in the 10 kHz–50 mHz frequency range, with a 10 mV perturbation amplitude. Measurements were carried out potentiostatically at −0.2 V (vs. RHE) in 1 M KOH using the same three-electrode configuration. Equivalent circuit fitting was used to calculate double-layer capacitance and charge transfer resistance values.

4. Conclusions

In summary, Pd-decorated CoP and CoFeP coatings were successfully fabricated on Cu substrate via a two-step process involving electroless deposition followed by galvanic displacement. Structural and morphological characterization, performed using techniques such as scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS), confirmed the successful formation of uniform surface coverage. The synthesized catalysts showed significant activity for both the HER and the OER, which are essential half-reactions in OWS. Specifically, for the HER, the PdCoFeP catalyst exhibited the lowest overpotential value of 207 mV compared to PdCoP, which required 227 mV to attain a current density of 10 mA cm−2 at 25 °C. However, it delivered a higher current density, indicating more efficient hydrogen generation under higher load conditions. In addition, the OER performance of the PdCoFeP and PdCoP catalysts demonstrated a marginally increased overpotential value of 396 mV at a current density of 10 mA cm−2 and 25 °C compared to the unmodified CoP and CoFeP catalysts, which exhibited overpotentials of 431 and 435 mV, respectively. These improvements are attributed to the synergistic interactions between Co, Fe, P, and Pd, which enhance electron transfer and facilitate HER/OER kinetics. Meanwhile, the assembled electrolyzer that used PdCoFeP as the cathode and anode required a cell voltage of 1.69 V to achieve 10 mA cm−2. This study demonstrates that all the synthesized variants of CoP, CoFeP, PdCoP, and PdCoFeP are effective and stable electrocatalysts for overall alkaline water splitting. Among these catalysts, PdCoFeP exhibits particular promise for applications requiring high current densities. EIS measurements showed that the PdCoFeP films exhibited a larger double-layer capacitance and lower charge transfer resistance, confirming the beneficial effects of modification with Pd for HER activity. Overall, these materials have significant potential for use in large-scale hydrogen production, renewable energy storage, and clean fuel cell technologies to support the transition to a sustainable energy future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15111035/s1, Figure S1: HER Polarization curves recorded on CoP (a), CoFeP (b), PdCoP (c), and PdCoFeP (d) catalysts in an N2-saturated 1 M KOH solution at a potential scan rate of 5 mV s−1 and a temperature range from (25–55 °C); Figure S2: OER Polarization curves recorded on CoP (a), CoFeP (b), PdCoP (c), and PdCoFeP (d) catalysts in an N2-saturated 1 M KOH solution at a potential scan rate of 5 mV s−1 and a temperature range from 25 to 55 °C; Table S1: Electrochemical performance of catalysts for the HER in alkaline media; Table S2: Electrochemical performance of catalysts for OER in alkaline media.

Author Contributions

Conceptualization, L.T.-T. and E.N.; methodology, H.A., G.S. and Z.M.; validation, V.K. and G.S.; formal analysis, A.B.; investigation, H.A., V.K. and Z.M.; data curation, A.B.; writing—original draft preparation, E.N. and H.A.; writing—review and editing, L.T.-T.; visualization, H.A. and A.B.; supervision, L.T.-T.; project administration, L.T.-T.; funding acquisition, L.T.-T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant (No. P-MIP-23-467) from the Research Council of Lithuania.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors are grateful to Ramūnas Levinas from the Department of Catalysis, Center for Physical Sciences and Technology for materials characterization by EIS.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Xiao, Z.; Wang, Y.; Huang, Y.C.; Wei, Z.; Dong, C.L.; Ma, J.; Shen, S.; Li, Y.; Wang, S. Filling the oxygen vacancies in Co3O4 with phosphorus: An ultra-efficient electrocatalyst for overall water splitting. Energy Environ. Sci. 2017, 10, 2563–2569. [Google Scholar] [CrossRef]
  2. Wang, J.; Cui, W.; Liu, Q.; Xing, Z.; Asiri, A.M.; Sun, X. Recent progress in cobalt—Based heterogeneous catalysts for electrochemical water splitting. Adv. Mater. 2016, 28, 215–230. [Google Scholar] [CrossRef]
  3. Liu, H.; He, P.; Wang, S.; Gao, J.; Zhou, L.; Li, C.; Zhang, Y.; Yang, D.; He, M.; Jia, L.; et al. Facile one-step fabrication of bimetallic Co–Ni–P hollow nanospheres anchored on reduced graphene oxide as highly efficient electrocatalyst for hydrogen evolution reaction. Int. J. Hydrog. Energy 2019, 44, 24140–24150. [Google Scholar] [CrossRef]
  4. Pilavachi, P.A.; Chatzipanagi, A.I.; Spyropoulou, A.I. Evaluation of hydrogen production methods using the analytic hierarchy process. Int. J. Hydrog. Energy 2009, 34, 5294–5303. [Google Scholar] [CrossRef]
  5. Kim, H.; Oh, S.; Cho, E.; Kwon, H. 3D porous cobalt–iron–phosphorus bifunctional electrocatalyst for the oxygen and hydrogen evolution reactions. ACS Sustain. Chem. Engineer. 2018, 6, 6305–6311. [Google Scholar] [CrossRef]
  6. Balat, M. Potential importance of hydrogen as a future solution to environmental and transportation problems. Int. J. Hydrog. Energy 2008, 33, 4013–4029. [Google Scholar] [CrossRef]
  7. Wang, S.; He, P.; Jia, L.; He, M.; Zhang, T.; Dong, F.; Liu, M.; Liu, H.; Zhang, Y.; Li, C.; et al. Nanocoral-like composite of nickel selenide nanoparticles anchored on two-dimensional multi-layered graphitic carbon nitride: A highly efficient electrocatalyst for oxygen evolution reaction. Appl. Catal. B Environ. 2019, 243, 463–469. [Google Scholar] [CrossRef]
  8. Wang, Y.; Xie, C.; Liu, D.; Huang, X.; Huo, J.; Wang, S. Nanoparticle-stacked porous nickel–iron nitride nanosheet: A highly efficient bifunctional electrocatalyst for overall water splitting. ACS Appl. Mater. Interf. 2016, 8, 18652–18657. [Google Scholar] [CrossRef]
  9. Zhang, J.; Wang, Y.; Zhang, C.; Gao, H.; Lv, L.; Han, L.; Zhang, Z. Self-supported porous NiSe2 nanowrinkles as efficient bifunctional electrocatalysts for overall water splitting. ACS Sustain. Chem. Engineer. 2018, 6, 2231–2239. [Google Scholar] [CrossRef]
  10. Yu, F.; Yu, L.; Mishra, I.K.; Yu, Y.; Ren, Z.F.; Zhou, H.Q. Recent developments in earth-abundant and non-noble electrocatalysts for water electrolysis. Mater. Today Phys. 2018, 7, 121–138. [Google Scholar] [CrossRef]
  11. Sun, H.; Xu, X.; Kim, H.; Jung, W.; Zhou, W.; Shao, Z. Electrochemical water splitting: Bridging the gaps between fundamental research and industrial applications. Energy Environ. Mater. 2023, 6, e12441. [Google Scholar] [CrossRef]
  12. Zhang, K.; Zou, R. Advanced transition metal-based OER electrocatalysts: Current status, opportunities, and challenges. Small 2021, 17, 2100129. [Google Scholar] [CrossRef]
  13. Mahmood, N.; Yao, Y.; Zhang, J.W.; Pan, L.; Zhang, X.; Zou, J.J. Electrocatalysts for hydrogen evolution in alkaline electrolytes: Mechanisms, challenges, and prospective solutions. Adv. Sci. 2018, 5, 1700464. [Google Scholar] [CrossRef]
  14. Huang, G.; Xiao, Z.; Chen, R.; Wang, S. Defect engineering of cobalt-based materials for electrocatalytic water splitting. ACS Sustain. Chem. Eng. 2018, 6, 15954–15969. [Google Scholar] [CrossRef]
  15. You, B.; Tang, M.T.; Tsai, C.; Abild-Pedersen, F.; Zheng, X.; Li, H. Enhancing electrocatalytic water splitting by strain engineering. Adv. Mater. 2019, 31, 1807001. [Google Scholar] [CrossRef]
  16. Wu, D.; Wei, Y.; Ren, X.; Ji, X.; Liu, Y.; Guo, X.; Liu, Z.; Asiri, A.M.; Wei, Q.; Sun, X. Co (OH)2 nanoparticle—Encapsulating conductive nanowires array: Room—Temperature electrochemical preparation for high—Performance water oxidation electrocatalysis. Adv. Mater. 2018, 30, 1705366. [Google Scholar] [CrossRef]
  17. Liu, Q.; Xie, L.; Liu, Z.; Du, G.; Asiri, A.M.; Sun, X. A Zn-doped Ni3 S2 nanosheet array as a high-performance electrochemical water oxidation catalyst in alkaline solution. Chem. Commun. 2017, 53, 12446–12449. [Google Scholar] [CrossRef]
  18. Jiang, N.; You, B.; Sheng, M.; Sun, Y. Electrodeposited cobalt-phosphorous-derived films as competent bifunctional catalysts for overall water splitting. Angew. Chem. 2015, 127, 6349–6352. [Google Scholar] [CrossRef]
  19. Gao, X.; Zhang, H.; Li, Q.; Yu, X.; Hong, Z.; Zhang, X.; Liang, C.; Lin, Z. Hierarchical NiCo2O4 hollow microcuboids as bifunctional electrocatalysts for overall water—Splitting. Angew. Chem. Int. Ed. 2016, 55, 6290–6294. [Google Scholar] [CrossRef]
  20. Ouyang, T.; Ye, Y.Q.; Wu, C.Y.; Xiao, K.; Liu, Z.Q. Heterostructures composed of N-doped carbon nanotubes encapsulating cobalt and β-Mo2C nanoparticles as bifunctional electrodes for water splitting. Angew. Chem. Int. Ed. 2019, 58, 4923–4928. [Google Scholar] [CrossRef]
  21. Zhang, J.; Shang, X.; Ren, H.; Chi, J.; Fu, H.; Dong, B.; Liu, C.; Chai, Y. Modulation of inverse spinel Fe3O4 by phosphorus doping as an industrially promising electrocatalyst for hydrogen evolution. Adv. Mater. 2019, 31, 1905107. [Google Scholar] [CrossRef]
  22. Esposito, D.V.; Hunt, S.T.; Stottlemyer, A.L.; Dobson, K.D.; McCandless, B.E.; Birkmire, R.W.; Chen, J.G. Low-cost hydrogen-evolution catalysts based on monolayer platinum on tungsten monocarbide substrates. Angew. Chemie Int. Ed. 2010, 51, 9859–9862. [Google Scholar] [CrossRef]
  23. Huang, X.; Zeng, Z.; Bao, S.; Wang, M.; Qi, X.; Fan, Z.; Zhang, H. Solution-phase epitaxial growth of noble metal nanostructures on dispersible single-layer molybdenum disulfide nanosheets. Nat. Commun. 2013, 4, 1444. [Google Scholar] [CrossRef]
  24. Li, X.; Chen, G.; Liu, Y.; Lu, R.; Ma, C.; Wang, Z.; Han, Y.; Wang, D. Optimal selection of RuO2 for durable oxygen evolution reactions in acidic media by continuous regulation of Ru–O covalency. Energy Environ. Sci. 2025, 18, 4200–4209. [Google Scholar] [CrossRef]
  25. Li, L.; Wang, B.; Zhang, G.; Yang, G.; Yang, T.; Yang, S.; Yang, S. Electrochemically modifying the electronic structure of IrO2 nanoparticles for overall electrochemical water splitting with extensive adaptability. Adv. Energy Mater. 2020, 10, 2001600. [Google Scholar] [CrossRef]
  26. Fan, J.; Qi, K.; Zhang, L.; Zhang, H.; Yu, S.; Cui, X. Engineering Pt/Pd interfacial electronic structures for highly efficient hydrogen evolution and alcohol oxidation. ACS Appl. Mater. Interf. 2017, 9, 18008–18014. [Google Scholar] [CrossRef]
  27. Nsanzimana, J.M.V.; Peng, Y.; Xu, Y.Y.; Thia, L.; Wang, C.; Xia, B.Y.; Wang, X. An efficient and earth-abundant oxygen-evolving electrocatalyst based on amorphous metal borides. Adv. Energy Mater. 2018, 8, 1701475. [Google Scholar] [CrossRef]
  28. Shang, X.; Zhang, X.Y.; Xie, J.Y.; Dong, B.; Chi, J.Q.; Guo, B.Y.; Yang, M.; Chai, Y.M.; Liu, C.G. Double-catalytic-site engineering of nickel-based electrocatalysts by group VB metals doping coupling with in-situ cathodic activation for hydrogen evolution. Appl. Catal. B Environ. 2019, 258, 117984. [Google Scholar] [CrossRef]
  29. Du, C.F.; Liang, Q.; Dangol, R.; Zhao, J.; Ren, H.; Madhavi, S.; Yan, Q. Layered trichalcogenidophosphate: A new catalyst family for water splitting. Nano-Micro Lett. 2018, 10, 67. [Google Scholar] [CrossRef]
  30. Anantharaj, S.; Ede, S.R.; Sakthikumar, K.; Karthick, K.; Mishra, S.; Kundu, S. Recent trends and perspectives in electrochemical water splitting with an emphasis on sulfide, selenide, and phosphide catalysts of Fe, Co, and Ni: A review. ACS Catal. 2016, 6, 8069–8097. [Google Scholar] [CrossRef]
  31. Fareed, I.; Khan, M.D.; Murtaza, S.; Farooq Mu, H.; Rehman, Z.U.; Farooq, M.U.; Butt, F.K.; Tahir, M. Investigating metal (M = Mn, Fe, and Ni) doped Co(OH)2 nanofibers for electrocatalytic oxygen evolution and electrochemical biosensing performance. RSC Adv. 2024, 14, 26556–26567. [Google Scholar] [CrossRef]
  32. Linke, J.; Rohrbach, T.; Clark, A.H.; Borca, C.; Huthwelker, T.; Buchauer, F.L.; Kraglund, M.R.; Chatzichristodoulou, C.; Meade, E.; Guehl, J.; et al. The role of Fe incorporation into Ni-MOF-74 derived oxygen evolution electrocatalysts for anion exchange membrane water electrolysis. EES Catal. 2025, 3, 505–514. [Google Scholar] [CrossRef]
  33. Wang, J.; Li, S.; Guo, Y.; Ding, J.; Lu, Z. Preparation of S-doped Ni-Mn-Fe layered hydroxide for high performance of oxygen evolution reaction. Coatings 2025, 15, 825. [Google Scholar] [CrossRef]
  34. Lee, J.; Lee, J. Recent developments in manganese-based materials for acidic oxygen evolution catalysis. Korean J. Chem. Eng. 2025, 1–15. [Google Scholar] [CrossRef]
  35. Ahmed, K.W.; Fowler, M. Performance evaluation and durability analysis of NiFeCoOₓ catalysts for alkaline water electrolysis in anion exchange membrane electrolyzers. Catalysts 2024, 14, 322. [Google Scholar] [CrossRef]
  36. Masa, J.; Weide, P.; Peeters, D.; Sinev, I.; Xia, W.; Sun, Z.; Somsen, C.; Muhler, M.; Schuhmann, W. Amorphous cobalt boride (Co2B) as a highly efficient nonprecious catalyst for electrochemical water splitting: Oxygen and hydrogen evolution. Adv. Energy Mater. 2016, 6, 1502313. [Google Scholar] [CrossRef]
  37. Guan, M.; Fu, S.; Qin, F.; Zuo, P.; Shen, W. Urchin-like bifunctional CoP/FeP electrocatalyst for overall water splitting. Int. J. Hydrog. Energy 2024, 93, 147–157. [Google Scholar] [CrossRef]
  38. Ashraf, M.A.; Li, C.; Pham, B.T.; Zhang, D. Electrodeposition of Ni–Fe–Mn ternary nanosheets as affordable and efficient electrocatalyst for both hydrogen and oxygen evolution reactions. Int. J. Hydrog. Energy 2020, 45, 24670–24683. [Google Scholar] [CrossRef]
  39. Guo, Y.; Guo, D.; Ye, F.; Wang, K.; Shi, Z. Synthesis of lawn-like NiS2 nanowires on carbon fiber paper as bifunctional electrodes for water splitting. Int. J. Hydrog. Energy 2017, 42, 17038–17048. [Google Scholar] [CrossRef]
  40. Liu, Y.; Xie, X.; Zhu, G.; Mao, Y.; Yu, Y.; Ju, S.; Shen, X.; Pang, H. Small sized Fe–Co sulfide nanoclusters anchored on carbon for oxygen evolution. J. Mater. Chem. A 2019, 7, 15851–15861. [Google Scholar] [CrossRef]
  41. Li, F.; Zhang, L.; Li, J.; Lin, X.; Li, X.; Fang, Y.; Huang, J.; Li, W.; Tian, M.; Jin, J.; et al. Synthesis of Cu–MoS2/rGO hybrid as non-noble metal electrocatalysts for the hydrogen evolution reaction. J. Power Sources 2015, 292, 15–22. [Google Scholar] [CrossRef]
  42. Gupta, S.; Patel, N.; Fernandes, R.; Kadrekar, R.; Dashora, A.; Yadav, A.K.; Bhattacharyya, D.; Jha, S.N.; Miotello, A.; Kothari, D.C. Co–Ni–B nanocatalyst for efficient hydrogen evolution reaction in wide pH range. Appl. Catal. B Environ. 2016, 192, 126–133. [Google Scholar] [CrossRef]
  43. Chen, M.; Xie, Y.; Wu, J.X.; Huang, H.; Teng, J.; Wang, D.; Fan, Y.; Jiang, J.J.; Wang, H.P.; Su, C.Y. Cobalt (oxy) hydroxide nanosheet arrays with exceptional porosity and rich defects as a highly efficient oxygen evolution electrocatalyst under neutral conditions. J. Mater. Chem. A 2019, 7, 10217–10224. [Google Scholar] [CrossRef]
  44. Babar, P.; Lokhande, A.; Shin, H.H.; Pawar, B.; Gang, M.G.; Pawar, S.; Kim, J.H. Cobalt iron hydroxide as a precious metal—Free bifunctional electrocatalyst for efficient overall water splitting. Small 2018, 14, 1702568. [Google Scholar] [CrossRef]
  45. Liu, W.; Liu, H.; Dang, L.; Zhang, H.; Wu, X.; Yang, B.; Li, Z.; Zhang, X.; Lei, L.; Jin, S. Amorphous cobalt–iron hydroxide nanosheet electrocatalyst for efficient electrochemical and photo—Electrochemical oxygen evolution. Adv. Funct. Mater. 2017, 27, 1603904. [Google Scholar] [CrossRef]
  46. Wang, K.; She, X.; Chen, S.; Liu, H.; Li, D.; Wang, Y.; Zhang, H.; Yang, D.; Yao, X. Boosting hydrogen evolution via optimized hydrogen adsorption at the interface of CoP3 and Ni2P. J. Mater. Chem. A 2018, 6, 5560–5565. [Google Scholar] [CrossRef]
  47. Beltrán-Suito, R.; Menezes, P.W.; Driess, M. Amorphous outperforms crystalline nanomaterials: Surface modifications of molecularly derived CoP electro (pre) catalysts for efficient water-splitting. J. Mater. Chem. A 2019, 7, 15749–15756. [Google Scholar] [CrossRef]
  48. Amber, H.; Balčiūnaitė, A.; Sukackienė, Z.; Tamašauskaitė-Tamašiūnaitė, L.; Norkus, E. Electrolessly deposited cobalt–phosphorus coatings for efficient hydrogen and oxygen evolution reactions. Catalysts 2024, 15, 8. [Google Scholar] [CrossRef]
  49. Chen, M.; Zhu, C.; Zhu, X.; Yuan, K. Developing a cobalt phosphide catalyst with combined cobalt defects and phosphorus vacancies to boost oxygen evolution reaction. Materials 2024, 17, 4647. [Google Scholar] [CrossRef]
  50. Du, Y.; Zhong, Z.; Zhou, L.; Chen, W.; Shi, Z.; Song, P.; Liu, Y.; Yao, Y.; Liu, Y.; Dou Sh Xiao, Y. Coordination polymer derived transition metal phosphide/carbon composites for bifunctional oxygen electrocatalyst. J. Mater. Sci. Technol. 2025, 222, 142–151. [Google Scholar] [CrossRef]
  51. Zhang, T.; Zhong, J.; Gao, W.; Wang, Y. Interface engineering induced N, P doped carbon shell encapsulated FeP/NiP2/Ni5P4/NiP nanoparticles for highly efficient hydrogen evolution reaction. Coatings 2024, 14, 817. [Google Scholar] [CrossRef]
  52. Yadav, A.; Das, M.R.; Deka, S. Fine tuning of torus shaped Mo doped Ni2P nanorings for enhanced seawater electrolysis. Small 2025, 21, 2408036. [Google Scholar] [CrossRef]
  53. Parvanova-Mancheva, T.; Chakarova, V.; Gabrovska, M.; Atanasova-Vladimirova, S.; Baeva, A.; Andreeva, R.; Hodzhaoglu, F.; Nikolova, D. Evaluating the bifunctional properties towards HER and OER of NiCo electrodeposited coatings: Combined influence of support, Ni/Co ratio, and phosphorus doping. Catal. Today 2025, 460, 115495. [Google Scholar] [CrossRef]
  54. Thang, S.; Li, D.; Wu, X. Co0.4Fe1.6P electrocatalysts through Mo ion doping strategies for stable alkaline water splitting. Cryst. Eng. Comm. 2025, 27, 5319–5326. [Google Scholar] [CrossRef]
  55. Ren, P.; Wang, R.; Yang, Y.; Wang, T.; Hong, Y.; Zheng, Y.; Ren, X.; Jia, Z. Facile synthesis of MoP and its composite structure with Ru as an efficient electrocatalyst for hydrogen evolution reaction in both acidic and alkaline conditions. Materials 2025, 18, 1184. [Google Scholar] [CrossRef]
  56. Xiao, L.; Bao, W.; Zhang, J.; Yang, C.; Ai, T.; Li, Y.; Wei, X.; Jiang, P.; Kou, L. Interfacial interaction between NiMoP and NiFe-LDH to regulate the electronic structure toward high-efficiency electrocatalytic oxygen evolution reaction. Int. J. Hydrog. Energy 2022, 47, 9230–9238. [Google Scholar] [CrossRef]
  57. Liu, P.; Rodriguez, J.A. Catalysts for hydrogen evolution from the [NiFe] hydrogenase to the Ni2P (001) surface: The importance of ensemble effect. J. Am. Chem. Soc. 2005, 127, 14871–14878. [Google Scholar] [CrossRef]
  58. Sirisomboonchai, S.; Kitiphatpiboon, N.; Chen, M.; Li, S.; Li, X.; Kongparakul, S.; Samart, C.; Zhang, L.; Abudula, A.; Guan, G. Multi-hierarchical porous Mn-doped CoP catalyst on nickel phosphide foam for hydrogen evolution reaction. ACS Appl. Energy Mater. 2021, 5, 149–158. [Google Scholar] [CrossRef]
  59. Huang, X.; Xu, X.; Li, C.; Wu, D.; Cheng, D.; Cao, D. Vertical CoP nanoarray wrapped by N, P-doped carbon for hydrogen evolution reaction in both acidic and alkaline conditions. Adv. Energy Mater. 2019, 9, 1803970. [Google Scholar] [CrossRef]
  60. Liu, Q.; Tian, J.; Cui, W.; Jiang, P.; Cheng, N.; Asiri, A.M.; Sun, X. Carbon nanotubes decorated with CoP nanocrystals: A highly active non-noble—Metal nanohybrid electrocatalyst for hydrogen evolution. Angew. Chem. Int. Ed. 2014, 53, 6710–6714. [Google Scholar] [CrossRef]
  61. Wu, L.; Yu, L.; McElhenny, B.; Xing, X.; Luo, D.; Zhang, F.; Bao, J.; Chen, S.; Ren, Z. Rational design of core-shell-structured CoPx@ FeOOH for efficient seawater electrolysis. Appl. Catal. B Environ. 2021, 294, 120256. [Google Scholar] [CrossRef]
  62. Yang, C.; He, T.; Zhou, W.; Deng, R.; Hang, Q. Iron-tuned 3D cobalt–phosphate catalysts for efficient hydrogen and oxygen evolution reactions over a wide pH range. ACS Sustain. Chem. Engineer. 2020, 8, 13793–13804. [Google Scholar] [CrossRef]
  63. Lu, S.S.; Zhang, L.M.; Dong, Y.W.; Zhang, J.Q.; Yan, X.T.; Sun, D.F.; Shang, X.; Chi, J.Q.; Chai, Y.M.; Dong, B. Tungsten-doped Ni–Co phosphides with multiple catalytic sites as efficient electrocatalysts for overall water splitting. J. Mater. Chem. A 2019, 7, 16859–16866. [Google Scholar] [CrossRef]
  64. Muthurasu, A.; Ojha, G.P.; Lee, M.; Kim, H.Y. Integration of cobalt metal–organic frameworks into an interpenetrated prussian blue analogue to derive dual metal–organic framework-assisted cobalt iron derivatives for enhancing electrochemical total water splitting. J. Phys. Chem. C 2020, 124, 14465–14476. [Google Scholar] [CrossRef]
  65. Zhang, Q.; Yan, D.; Nie, Z.; Qiu, X.; Wang, S.; Yuan, J.; Su, D.; Wang, G.; Wu, Z. Iron-doped NiCoP porous nanosheet arrays as a highly efficient electrocatalyst for oxygen evolution reaction. ACS Appl. Energy Mater. 2018, 1, 571–579. [Google Scholar] [CrossRef]
  66. Tang, C.; Gan, L.; Zhang, R.; Lu, W.; Jiang, X.; Asiri, A.M.; Sun, X.; Wang, J.; Chen, L. Ternary FexCo1–x P nanowire array as a robust hydrogen evolution reaction electrocatalyst with Pt-like activity: Experimental and theoretical insight. Nano Lett. 2016, 16, 6617–6621. [Google Scholar] [CrossRef]
  67. Li, C.; Zhu, D.; Cheng, S.; Zuo, Y.; Wang, Y.; Ma, C.; Dong, H. Recent research progress of bimetallic phosphides-based nanomaterials as cocatalyst for photocatalytic hydrogen evolution. Chin. Chem. Lett. 2022, 33, 1141–1153. [Google Scholar] [CrossRef]
  68. Agyapoma, A.L.; Jin, C.; Zhang, Q.; Zeng, X. Construction of self-supporting heterobimetallic phosphides for oxygen evolution reaction. Mater. Today Catal. 2025, 10, 100114. [Google Scholar] [CrossRef]
  69. Shi, Y.; Zhang, B. Recent advances in transition metal phosphide nanomaterials: Synthesis and applications in hydrogen evolution reaction. Chem. Soc. Rev. 2016, 45, 1529–1541. [Google Scholar] [CrossRef]
  70. Chen, Z.; Wei, W.; Ni, B.J. Cost-effective catalysts for renewable hydrogen production via electrochemical water splitting: Recent advances. Curr. Opin. Green Sustain. Chem. 2021, 27, 100398. [Google Scholar] [CrossRef]
  71. Cho, J.; Moon, J.; Jeong, K.; Cho, G. Application of PU-sealing into Cu/Ni electroless plated polyester fabrics for e-textiles. Fiber. Polym. 2007, 8, 330–334. [Google Scholar] [CrossRef]
  72. Zhang, X.; Wang, F.; Zhou, Y.; Liang, A.; Zhang, J. Aluminum-induced direct electroless deposition of Co and Co-P coatings on copper and their catalytic performance for electrochemical water splitting. Surf. Coat. Technol. 2018, 352, 42–48. [Google Scholar] [CrossRef]
  73. Lin, C.C.; Chuang, C.C.; Li, X.H.; Chin, T.S.; Chang, J.Y.; Sung, C.K.; Wang, S.C. Ultrasound-assisted electroless deposition of Co-P hard magnetic films. Surf. Coat. Technol. 2020, 388, 125577. [Google Scholar] [CrossRef]
  74. Du, Y.; Qu, H.; Liu, Y.; Han, Y.; Wang, L.; Dong, B. Bimetallic CoFeP hollow microspheres as highly efficient bifunctional electrocatalysts for overall water splitting in alkaline media. Appl. Surf. Sci. 2019, 465, 816–823. [Google Scholar] [CrossRef]
  75. Zahra, R.; Pervaiz, E.; Baig, M.M.; Rabi, O. Three-dimensional hierarchical flowers-like cobalt-nickel sulfide constructed on graphitic carbon nitride: Bifunctional non-noble electrocatalyst for overall water splitting. Electrochim. Acta 2022, 418, 140346. [Google Scholar] [CrossRef]
  76. Kong, F.; Zhang, W.; Sun, L.; Huo, L.; Zhao, H. Interface electronic coupling in hierarchical FeLDH(FeCo)/Co(OH)2 arrays for efficient electrocatalytic oxygen evolution. ChemSusChem 2019, 12, 3592–3601. [Google Scholar] [CrossRef]
  77. Kaya, D.; Isik, H.H.; Isik, I.B.; Sigircik, G.; Tuken, T.; Karadag, F.; Ekicibil, A. Electrocatalytic hydrogen evolution on metallic and bimetallic Pd-Co alloy nanoparticles. Int. J. Hydrog. Energy 2023, 48, 14633–14641. [Google Scholar] [CrossRef]
  78. Goryachev, A.; Gao, L.; Zhang, Y.; Rohling, R.Y.; Vervuurt, R.H.J.; Bol, A.A.; Hofmann, J.P.; Hensen, E.J.M. Stability of CoPx electrocatalysts in continuous and interrupted acidic electrolysis of water. ChemElectroChem 2018, 5, 1230–1239. [Google Scholar] [CrossRef]
  79. Li, C.; Peng, Z.; Huang, J. Impedance response of electrochemical interfaces: Part IV─Low-frequency inductive loop for a single-electron reaction. J. Phys. Chem. C 2023, 127, 16367–16373. [Google Scholar] [CrossRef]
  80. Levinas, R.; Tsyntsaru, N.; Lelis, M.; Cesiulis, H. Synthesis, electrochemical impedance spectroscopy study and photoelectrochemical behaviour of as-deposited and annealed WO3 films. Electrochim. Acta 2017, 225, 29–38. [Google Scholar] [CrossRef]
  81. Brug, G.J.; van den Eeden, A.L.G.; Sluyters-Rehbach, M.; Sluyters, J.H. The analysis of electrode impedances complicated by the presence of a constant phase element. J. Electroanal. Chem. Interf. Electrochem. 1984, 176, 275–295. [Google Scholar] [CrossRef]
  82. Trotochaud, L.; Ranney, J.K.; Williams, K.N.; Boettcher, S.W. Solution-cast metal oxide thin film electrocatalysts for oxygen evolution. J. Am. Chem. Soc. 2012, 134, 17253–17261. [Google Scholar] [CrossRef]
  83. Chang, Y.; Shi, N.E.; Zhao, S.; Xu, D.; Liu, C.; Tang, Y.J.; Dai, Z.; Lan, Y.Q.; Han, M.; Bao, J. Coralloid Co2P2O7 nanocrystals encapsulated by thin carbon shells for enhanced electrochemical water oxidation. ACS Appl. Mater. Interf. 2016, 8, 22534–22544. [Google Scholar] [CrossRef]
  84. Liu, M.; Li, J. Cobalt phosphide hollow polyhedron as efficient bifunctional electrocatalysts for the evolution reaction of hydrogen and oxygen. ACS Appl. Mater. Interfaces 2016, 8, 2158–2165. [Google Scholar] [CrossRef]
  85. Wang, Y.; Zhou, T.; Jiang, K.; Da, P.; Peng, Z.; Tang, J.; Kong, B.; Cai, W.B.; Yang, Z.; Zheng, G. Reduced mesoporous Co3O4 nanowires as efficient water oxidation electrocatalysts and supercapacitor electrodes. Adv. Energy Mater. 2014, 4, 1400696. [Google Scholar] [CrossRef]
  86. Zhu, Y.P.; Liu, Y.P.; Ren, T.Z.; Yuan, Z.Y. Self-supported cobalt phosphide mesoporous nanorod arrays: A flexible and bifunctional electrode for highly active electrocatalytic water reduction and oxidation. Adv. Funct. Mater. 2015, 25, 7337–7347. [Google Scholar] [CrossRef]
  87. Ding, C.; Yu, Y.; Wang, Y.; Mu, Y.; Dong, X.; Meng, C.; Huang, C.; Zhang, Y. Phosphate-modified cobalt silicate hydroxide with improved oxygen evolution reaction. J. Colloid. Interface Sci. 2023, 648, 251–258. [Google Scholar] [CrossRef]
  88. Yang, Y.; Fei, H.; Ruan, G.; Tour, J.M. Porous cobalt-based thin film as a bifunctional catalyst for hydrogen generation and oxygen generation. Adv. Mater. 2015, 27, 3175–3180. [Google Scholar] [CrossRef]
  89. Liu, Y.; Ran, N.; Ge, R.; Liu, J.; Li, W.; Chen, Y.; Feng, L.; Che, R. Porous Mn-doped cobalt phosphide nanosheets as highly active electrocatalysts for oxygen evolution reaction. Chem. Eng. J. 2021, 425, 131642. [Google Scholar] [CrossRef]
  90. Huang, G.; Liang, W.; Wu, Y.; Li, J.; Jin, Y.Q.; Zeng, H.; Zhang, H.; Xie, F.; Chen, J.; Wang, N.; et al. Co2P/CoP hybrid as a reversible electrocatalyst for hydrogen oxidation/evolution reactions in alkaline medium. J. Catal. 2020, 390, 23–29. [Google Scholar] [CrossRef]
  91. Tang, C.; Asiri, A.M.; Luo, Y.; Sun, X. Electrodeposited Ni-P alloy nanoparticle films for efficiently catalyzing hydrogen- and oxygen-evolution reactions. ChemNanoMat. 2015, 1, 558–561. [Google Scholar] [CrossRef]
  92. Wang, X.; Xiao, X.; Yan, H.; Chen, C.; Zhang, D.; Sun, D.; Xu, X. The defect-rich porous CoP4/Co4S3 microcubes as robust electrocatalyst for clean H2 energy production via alkaline overall water splitting. Appl. Catal. A-Gen. 2023, 668, 119461. [Google Scholar] [CrossRef]
  93. Zhao, S.; Li, C.; Huang, H.; Liu, Y.; Kang, Z. Carbon n+anodots modified cobalt phosphate as efficient electrocatalyst for water oxidation. J. Mater. 2015, 1, 236–244. [Google Scholar]
  94. Yang, L.; Li, L. Synthesis of uniformly dispersed NiCoFe trimetallic phosphide for efficient overall water splitting. J. Phys. Chem. Solids 2025, 207, 112926. [Google Scholar] [CrossRef]
  95. Jiang, S.; Li, Q.; Zhao, Q.; Cheng, L.; Jiang, T. A hierarchical core shell CuO@FeCoP/CF heterostructure for efficient overall water splitting in alkaline media. J. Alloys Compd. 2024, 1034, 177291. [Google Scholar] [CrossRef]
  96. Luo, J.; Tuo, Y.; Zhang, J.; Li, W.; Wang, S.; Li, Y.; Pan, Y.; Zhou, Y.; Zhang, J. Carbon confined Mo doped CoP for alkali overall water splitting with enhanced catalytic activity and durability. Chem. Eng. J. 2025, 510, 161519. [Google Scholar] [CrossRef]
  97. Yang, Z.; Li, K.; Zhang, D.; Yang, H. The synergistic effect of hard magnetic Co2C and soft magnetic CoFe improves electrocatalytic overall water splitting performance. J. Alloys Compds. 2025, 1036, 181593. [Google Scholar] [CrossRef]
  98. Jiao, L.; Zhou, Y.X.; Jiang, H.L. Metal-organic framework-based CoP/reduced graphene oxide: High-performance bifunctional electrocatalyst for overall water splitting. Chem. Sci. 2016, 7, 1690–1695. [Google Scholar] [CrossRef]
  99. Guo, P.; Wu, J.; Li, X.-B.; Luo, J.; Lau, W.-M.; Liu, H.; Sun, X.-L.; Liu, L.-M. A highly stable bifunctional catalyst based on 3D Co(OH)2@NCNTs@NF towards overall water-splitting. Nano Energy 2018, 47, 96–104. [Google Scholar] [CrossRef]
  100. Liu, X.; Dong, J.; You, B.; Sun, Y. Competent overall water-splitting electrocatalysts derived from ZIF-67 grown on carbon cloth. RSC Adv. 2016, 6, 73336–73342. [Google Scholar] [CrossRef]
  101. Niyitanga, T.; Kim, H. Time-dependent oxidation of graphite and cobalt oxide nanoparticles as electrocatalysts for the oxygen evolution reaction. J. Electroanal. Chem. 2022, 914, 116297. [Google Scholar] [CrossRef]
  102. Luo, W.; Wang, Y.; Cheng, C. Ru-based electrocatalysts for hydrogen evolution reaction: Recent research advances and perspectives. Mater. Today Phys. 2020, 15, 100274. [Google Scholar] [CrossRef]
  103. Cossar, E.; Houache, M.S.E.; Zhang, Z.; Baranova, E.A. Comparison of electrochemical active surface area methods for various nickel nanostructures. J. Electroanal. Chem. 2020, 870, 114246. [Google Scholar] [CrossRef]
Figure 1. SEM images of CoP (a), CoFeP (b), PdCoP (c), and PdCoFeP (d) deposited on the Cu surface. (a’d’) EDX spectra corresponding to the aforementioned SEM images.
Figure 1. SEM images of CoP (a), CoFeP (b), PdCoP (c), and PdCoFeP (d) deposited on the Cu surface. (a’d’) EDX spectra corresponding to the aforementioned SEM images.
Catalysts 15 01035 g001
Figure 2. XRD patterns of CoP (a), CoFeP (b), PdCoP (c), and PdCoFeP (d) deposited on the glass surface. Peak positions are indicated according to Crystallography Open Database (COD) data cards: Co—1512501; CoP—9008928; Fe—1512319; CoFe—1524319; Pd—1011105.
Figure 2. XRD patterns of CoP (a), CoFeP (b), PdCoP (c), and PdCoFeP (d) deposited on the glass surface. Peak positions are indicated according to Crystallography Open Database (COD) data cards: Co—1512501; CoP—9008928; Fe—1512319; CoFe—1524319; Pd—1011105.
Catalysts 15 01035 g002
Figure 3. (a) HER polarization curves in an N2-saturated 1 M KOH solution at a potential scan rate of 5 mV s−1 and a temperature of 25 °C. (b) Bar columns of the corresponding overpotentials at current densities of 10, 20, 50 and 100 mA cm−2. (c,d) The corresponding Tafel slopes and Arrhenius plots of the current at −0.40 V versus RHE for each catalyst.
Figure 3. (a) HER polarization curves in an N2-saturated 1 M KOH solution at a potential scan rate of 5 mV s−1 and a temperature of 25 °C. (b) Bar columns of the corresponding overpotentials at current densities of 10, 20, 50 and 100 mA cm−2. (c,d) The corresponding Tafel slopes and Arrhenius plots of the current at −0.40 V versus RHE for each catalyst.
Catalysts 15 01035 g003
Figure 4. Nyquist plots of the CoFeP and PdCoFeP catalysts, obtained at −0.2 V vs. RHE. Solid lines represent fits to the equivalent circuits shown in the inset.
Figure 4. Nyquist plots of the CoFeP and PdCoFeP catalysts, obtained at −0.2 V vs. RHE. Solid lines represent fits to the equivalent circuits shown in the inset.
Catalysts 15 01035 g004
Figure 5. (a) OER polarization curves in an N2-saturated 1 M KOH solution at a potential scan rate of 5 mV s−1 and a temperature of 25 °C. (b) Bar columns of the corresponding overpotentials at 10, 20, 50 and 100 mA cm−2. (c,d) The corresponding Tafel slopes and Arrhenius plots of the current at 1.8 V versus RHE for each catalyst.
Figure 5. (a) OER polarization curves in an N2-saturated 1 M KOH solution at a potential scan rate of 5 mV s−1 and a temperature of 25 °C. (b) Bar columns of the corresponding overpotentials at 10, 20, 50 and 100 mA cm−2. (c,d) The corresponding Tafel slopes and Arrhenius plots of the current at 1.8 V versus RHE for each catalyst.
Catalysts 15 01035 g005
Figure 6. CVs of (a) CoP, (b) CoFeP, (c) PdCoP and (d) PdCoFeP in an N2-saturated 1 M KOH solution in the non-Faradaic potential region at different scan rates (10–50 mV s−1). (e) The capacitive current as a function of the scan rate.
Figure 6. CVs of (a) CoP, (b) CoFeP, (c) PdCoP and (d) PdCoFeP in an N2-saturated 1 M KOH solution in the non-Faradaic potential region at different scan rates (10–50 mV s−1). (e) The capacitive current as a function of the scan rate.
Catalysts 15 01035 g006
Figure 7. The CA curves of the investigated PdCoP (a) and PdCoFeP (b) catalysts in a 1 M KOH solution at 25 °C and a potential value of −0.26 V (vs. RHE).
Figure 7. The CA curves of the investigated PdCoP (a) and PdCoFeP (b) catalysts in a 1 M KOH solution at 25 °C and a potential value of −0.26 V (vs. RHE).
Catalysts 15 01035 g007
Figure 8. The CP curves of the investigated PdCoP (a) and PdCoFeP (b) catalysts in a 1 M KOH solution at a constant current density of −10 mA cm−2 and 25 °C.
Figure 8. The CP curves of the investigated PdCoP (a) and PdCoFeP (b) catalysts in a 1 M KOH solution at a constant current density of −10 mA cm−2 and 25 °C.
Catalysts 15 01035 g008
Figure 9. The polarization curves of the CoP, CoFeP, PdCoP, and PdCoFeP catalysts used as both an anode and a cathode for overall water splitting performance in the two-electrode setup at a scan rate of 5 mV s−1.
Figure 9. The polarization curves of the CoP, CoFeP, PdCoP, and PdCoFeP catalysts used as both an anode and a cathode for overall water splitting performance in the two-electrode setup at a scan rate of 5 mV s−1.
Catalysts 15 01035 g009
Table 1. Composition of coatings deposited on Cu surface determined by EDX analysis.
Table 1. Composition of coatings deposited on Cu surface determined by EDX analysis.
SampleElement, at%
PdCoFeP
CoP-90.22-9.78
CoFeP-83.789.966.26
PdCoP3.5483.72-12.74
PdCoFeP8.3664.0812.4915.07
Table 2. Summarized parameters for HER on the investigated catalysts in alkaline media.
Table 2. Summarized parameters for HER on the investigated catalysts in alkaline media.
SampleEonset *, Vη10 **, mVTafel Slope,
mV dec−1
Ea,
kJ mol−1
CoP0.16623978.111.3
CoFeP0.18824576.46.0
PdCoP0.17125076.34.7
PdCoFeP0.21325982.811.4
* Potential at −1 mA cm−2, ** Overpotential at 10 mA cm−2.
Table 3. Electrochemical performance of the investigated catalysts for OER in alkaline media at 25 °C.
Table 3. Electrochemical performance of the investigated catalysts for OER in alkaline media at 25 °C.
SampleEonset *, Vηonset, mVE **, Vη10 **, mVTafel Slope,
mV dec−1
Ea,
kJ mol−1
CoP1.52052901.6607431120.320.7
CoFeP1.58463551.665043574.521.2
PdCoP1.54263131.625539677.812.2
PdCoFeP1.55773281.625639670.511.8
Cu1.76385341.8750645--
* Values at 1 mA cm−2, ** Values at 10 mA cm−2.
Table 4. Comparison of OER performance of the investigated catalysts with some previously reported Co-P-based and most advanced noble–metal catalysts.
Table 4. Comparison of OER performance of the investigated catalysts with some previously reported Co-P-based and most advanced noble–metal catalysts.
Catalystη10 *, mV Tafel Slope,
mV dec−1
ElectrolyteRef.
PdCoFeP39670.51 M KOHThis study
PdCoP39677.81 M KOHThis study
CoOx423421 M KOH[82]
Co2P2O7490861 M KOH[83]
Co(PO3)2 nanosheets5741061 M KOH[83]
CoP hollow polyhedron400571 M KOH[84]
Reduced mesoporous Co3O4 nanowires400721 M KOH[85]
CoP-MNA/Ni foam390651 M KOH[86]
CoP350471 M KOH[18]
CoSi-P3091211 M KOH[87]
CoP300651 M KOH[88]
Mn-CoP28877.21 M KOH[89]
RuO2/CF3601641 M KOH[90]
RuO2 on NF290811 M KOH[91]
IrO2 commercial33994.51 M KOH[92]
Ir/C25471.91 M KOH[93]
Ni0.3Co2Fe-P290-1 M KOH[94]
Co2Fe-P367-1 M KOH[94]
CuO@FeCoP/CF240-1 M KOH[95]
Mo-CoP@C28053.11 M KOH[96]
CoFe@Co2C@Co-0.2/NF246-1 M KOH[97]
* Values at 10 mA cm−2.
Table 5. Comparison with various electrocatalysts for overall water splitting.
Table 5. Comparison with various electrocatalysts for overall water splitting.
Anode II CathodeCell Voltage, VElectrolyteRef.
PdCoFeP‖PdCoFeP1.69 1 M KOHThis study
PdCoP‖PdCoP1.78 1 M KOHThis study
Pt/C‖IrO21.71 1 M KOH[98]
Pt/C‖Pt/C1.83 1 M KOH[98]
CoP/rGO-400‖CoP/rGO-4001.70 1 M KOH[98]
Co(OH)2@NCNTs@NF‖
Co(OH)2@NCNTs@NF
1.72 1 M KOH[99]
Co-P/NC/CC‖Co-P/NC/CC1.77 1 M KOH[100]
Co-P/NC-CC‖Co-P/NC-CC1.95 1 M KOH[100]
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

Amber, H.; Balčiūnaitė, A.; Kepenienė, V.; Stalnionis, G.; Mockus, Z.; Tamašauskaitė-Tamašiūnaitė, L.; Norkus, E. Pd-Modified CoP and CoFeP Catalysts as Efficient Bifunctional Catalysts for Water Splitting. Catalysts 2025, 15, 1035. https://doi.org/10.3390/catal15111035

AMA Style

Amber H, Balčiūnaitė A, Kepenienė V, Stalnionis G, Mockus Z, Tamašauskaitė-Tamašiūnaitė L, Norkus E. Pd-Modified CoP and CoFeP Catalysts as Efficient Bifunctional Catalysts for Water Splitting. Catalysts. 2025; 15(11):1035. https://doi.org/10.3390/catal15111035

Chicago/Turabian Style

Amber, Huma, Aldona Balčiūnaitė, Virginija Kepenienė, Giedrius Stalnionis, Zenius Mockus, Loreta Tamašauskaitė-Tamašiūnaitė, and Eugenijus Norkus. 2025. "Pd-Modified CoP and CoFeP Catalysts as Efficient Bifunctional Catalysts for Water Splitting" Catalysts 15, no. 11: 1035. https://doi.org/10.3390/catal15111035

APA Style

Amber, H., Balčiūnaitė, A., Kepenienė, V., Stalnionis, G., Mockus, Z., Tamašauskaitė-Tamašiūnaitė, L., & Norkus, E. (2025). Pd-Modified CoP and CoFeP Catalysts as Efficient Bifunctional Catalysts for Water Splitting. Catalysts, 15(11), 1035. https://doi.org/10.3390/catal15111035

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