Next Article in Journal
Genome-Wide Characterization and Expression Pattern Analysis Insights into Plant NBS-LRR Gene Family of Salvia miltiorrhiza
Previous Article in Journal
Fenofibrate as a PPARα Agonist Modulates Neuroinflammation and Glutamate Receptors in a Rat Model of Temporal Lobe Epilepsy: Region-Specific Effects and Behavioral Outcomes
Previous Article in Special Issue
Hydrogen-Selective Pd-Ag-Ru Membranes and the Secret of High Permeability: The Influence of the Morphology of the Nano-Structured Coating on the Rate of Surface Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Paste Electrode Based on Copper and Gallium Mixed Metal Oxides-Decorated CNT for Highly Electrocatalyzed Hydrogen Evolution Reaction

1
Centro de Investigación de Estudios Avanzados del Maule (CIEAM), Vicerrectoría de Investigación y Postgrado, Universidad Católica del Maule, Av. San Miguel 3605, Talca 34809112, Chile
2
Physical Chemistry Department, Faculty of Chemical Sciences, University of Concepción, Víctor Lamas 1290, Concepción 4070386, Chile
3
Departamento de Química Inorgánica, Facultad de Química, Pontificia Universidad Católica de Chile, Av. Vicuña Mackenna 4860, Santiago 8331150, Chile
4
Millennium Institute on Green Ammonia as Energy Vector (MIGA), Av. Vicuña Mackenna 4860, Macul, Santiago 7820436, Chile
5
Departamento de Medicina Traslacional, Facultad de Medicina, Universidad Católica del Maule, Av. San Miguel 3605, Talca 34809112, Chile
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(18), 9057; https://doi.org/10.3390/ijms26189057
Submission received: 1 August 2025 / Revised: 11 September 2025 / Accepted: 16 September 2025 / Published: 17 September 2025
(This article belongs to the Special Issue Ion and Molecule Transport in Membrane Systems, 6th Edition)

Abstract

H2 has become one of the most attractive alternatives to replace fossil fuels in clean energy production, but large-scale production remains a challenge. A key step toward this goal is to develop new efficient electrocatalysts for H2 production. This work presents a new mixed metal oxides-decorated CNT paste electrode (MMO@C), which is highly electrocatalytic, for use in the hydrogen evolution reaction (HER). MMO@C is synthesized by a solvothermal method and used as an easy-to-prepare paste electrode. XPS and X-ray analysis indicate that the electrocatalyst corresponds to a mixed surface of Ga2O3-CuO-Cu2O-Cu(OH)2@C. The MMO@C electrocatalyst shows a positive Eo of 0.12 V vs. RHE at −10 mA cm−2 towards the HER in a neutral medium. In neutral and alkaline media, the presence of Ga2O3 facilitates the reduction of CuO to Cu(I) species, which is followed by the formation of Cu(s) active sites. Therefore, the excellent electrocatalytic performance toward the HER in a neutral medium is attributed to the synergistic effect between gallium and copper oxides on the electrode surface. The prominent H2 production using MMO@C electrocatalyst is 1.31 × 10−2 mol cm−2, with a turnover number (TON) of 39,423, a turnover frequency (TOF) of 13,141 h−1, and a faradaic efficiency (FE) of 94.3%. Although the Tafel slope reveals slow reaction kinetics, the outstanding onset potential allows for the coupling of the electrocatalyst to renewable energy production systems, making it an attractive candidate for producing green H2 and for application in membrane water electrolyzers.

1. Introduction

To reduce global greenhouse gas emissions and dependence on fossil fuels, most countries around the world are moving towards developing renewable energy sources [1]. An alternative to fossil fuels is hydrogen, which can be produced by electrolyzing water using electric current to separate it into H2 and O2. Water electrolysis produces no greenhouse gases, and when the electricity used to power the process comes entirely from renewable sources the resulting source is “green” hydrogen [2]. H2 has broad application prospects in the advancement of clean energy due to its extremely high energy density, renewable capacity, and zero carbon content [3]. It is even believed that in the future H2 will serve as the main energy storage technology, central heating fuel, and primary transportation fuel for cars, trucks, airplanes, and more [2]. Despite numerous research efforts in the design and development of electrocatalysts, the performance of water electrolysis is still insufficient for commercialization [4]. Ongoing research on innovative materials and understanding of their electrocatalytic mechanisms are critical for designing more efficient and stable electrolyzers [4]. Although platinum, ruthenium, and iridium-based materials have long been considered the reference electrocatalysts for electrochemical water splitting, their application is hampered by their high economic cost and limited abundance [5,6]. Recent research has allowed the effective design of electrocatalytic materials for H2 production based on transition metals such as cobalt (Co), nickel (Ni), copper (Cu), and iron (Fe), studied as oxides [7,8,9,10] and hydroxides [11,12,13,14] under different pH conditions. Cu-based electrocatalysts are a promising option for large-scale H2 production due to the low cost, high abundance, high electrical and thermal conductivity, and electrocatalytic versatility of Cu [15]. Cu is easily recyclable, reducing the need to mine new metals [16] and displays strong corrosion resistance compared to Co, Ni, and Fe [17]. Some copper oxides are advantageous due to their low preparation costs and diverse oxidation states. In this context, monovalent and divalent copper oxide (Cu2O and CuO) are considered excellent choices due to their high stability, enhanced electrochemical activity, and superior redox capabilities compared to other transition metal oxides [9,18,19]. For this reason, they are highly attractive materials to produce H2 [20] and for other reactions of energetic and environmental interest. As an example, H. Son et al. proposed a facile oxidation method to fabricate flexible and scalable Cu/Cu2O/CuO nanoleaves with a self-initiated charge transport platform via induced oxidation to enhance electrochemical water splitting, with an H2 production of 4.76 μmol cm−2 [21]. In a study by Y.J. Seo et al., the formation of a Cu2O/CuO/Cu(OH)2 layer was reported for the photoelectrochemical hydrogen evolution reaction (HER), and it was proposed that the generation of sufficient cation vacancies was closely related to the electronic conductivity and charge transfer performance. The results confirmed the presence of a slightly thin Cu(OH)2 film as an efficient cocatalyst, which led to a significant enhancement in the current density [20]. Regarding the use of monovalent and divalent copper-based materials as electrocatalysts for other reactions of energetic and environmental interest, N. Mumtazah et al. reported the effects of copper (I) and (II) oxide electrocatalysts on the oxidation of 5-hydroxymethylfurfural (HMF), exhibiting higher electrocatalytic activity for the dominant Cu2O electrode [22]. In another recent research, N. Dong et al. proposed the use of hierarchical Cu/Cu2O/CuO nanosheets for the electrocatalytic conversion of nitrate to ammonia. The authors indicated that the electronic interaction and interface effect between Cu/Cu2O/CuO allow the regulation of the d-band center of Cu and the control of the adsorption energies of the intermediates [23]. On the other hand, gallium-based materials attract attention due to their interesting properties. With a low melting point of 30 °C and the widest liquid temperature range between 30 °C and 2400 °C, gallium combines easily with other metals or nonmetals to build composite materials [24]. However, research on the construction of gallium-based materials for H2 production is relatively scarce [24]. A. Kakoria et al. applied highly porous β-gallium oxide (β-Ga2O3) nanofibers with a specific surface area of ~100–300 m2 g−1 as an efficient bifunctional electrocatalyst for H2 and O2 production. The β-Ga2O3 nanofibers electrocatalyzed the ORR at an onset potential of 0.84 V (vs RHE), and for the HER, although the onset potential was −0.34 V (vs RHE), the current density was visibly better than the Pt/C catalyst. The attractive performance related to low onset potential and high current density for both reactions was solely attributed to the large surface area and unique morphology presented by the material [25]. Y. Zang et al. showed a synergistic effect between Ga and Co in the preparation of ultrathin Ga-doped CoP nanosheets (Ga-CoP NSs). Ga-CoP NSs deliver improved HER electrocatalytic activities. Density functional theory simulations indicated that the Ga dopant could systematically enhance the HER activity of CoP by improving the H2O adsorption, weakening OH adsorption, and optimizing the H adsorption/desorption [26]. Recently, L. Chen et al. developed an innovative and energy-efficient Cu5Ga1 bimetallic electrocatalyst for CO2 conversion, where the HER acts as a competing reaction. In this study, it was determined that Cu5Ga1 exhibits a much higher activation energy for H2 production (39.7 kJ mol−1 at −1 V vs. RHE) compared to the reference Cu electrocatalyst (27.3 kJ mol−1 at −1 V vs. RHE). The authors observed that the presence of Ga shows a remarkable effect in preventing Cu oxidation under ambient conditions. This effect is probably due to the high oxophilicity and low electronegativity of Ga, which enhances the electrocatalytic properties of the bimetallic system [27]. Therefore, the use of mixed metal oxides of gallium and copper as electrocatalysts for hydrogen evolution seems convenient. In general, metal oxides can further enhance their electrochemical properties when supported on conductive materials, such as multi-walled carbon nanotubes (MWCNTs) [28,29,30]. The highly resonant MWCNTs system and their cylindrical structures enable directed electron flow, improving both redox processes and charge transfer [31]. For these reasons, it is proposed that Ga2O3 can enhance the electrocatalytic performance of copper species when located in proximity over a MWCNTs surface. This work proposes a new paste electrode composed of MWCNTs decorated with mixed metal oxides of gallium and copper (MMO@C) to electrocatalyze the HER. The novelty of this work lies in the analysis of the overall mechanism by which an electrochemical process between both metals improves HER performance. The new electrocatalyst developed in this work holds promise for membrane water electrolyzers.

2. Results and Discussion

2.1. Structural, Surface, and Morphological Properties

The XRD powder patterns of Ga2O3@C and CuO@C pure composites are shown in Figure 1a,b, respectively. For better visualization, the oxidized MWCNTs, Ga2O3, and CuO single crystal diffractograms were included. For the oxidized MWNCTs, the characteristic diffraction peaks at 2θ 26.05° and 43.03° correspond to the planes (002) and (100), indicating multi-layered carbon nanotubes with concentric overlapping graphene sheets. In Figure 1a, the appearance of only three diffraction peaks is seen for Ga2O3@C, attributed to the (002) plane of the MWCNTs and cubic Ga2O3 (Fd-3m ICSD- 194506) [32], whereas in Figure 1b for CuO@C a large number of diffraction peaks are seen. Therefore, in the so-called CuO@C pure composite the presence of monoclinic CuO (C2/c ICSD- 1011148) [33] as well as some impurities of Cu(OH)2 and Cu2O are detected. Consequently, the XRD characterization indicates that Ga2O3@C is structurally more homogeneous than CuO@C, and in both materials disappearance of the plane (100) of MWNCTs at 2θ 43° is detected. This is an important finding due to the disappearance of the (100) plane in MWCNTs, with the composites indicating a modification of the interlayer or collapse of the structure [34]. Zhu et al. report that the strong interactions between cobalt oxide and the carbon framework in Co-bpdc/MWCNT composites disrupt the graphitic structure with the disappearance of the (100) plane after calcination [35]. Abdullah et al. report significant changes in the intensity and position in the XRD patterns of MWCNTs decorated with iron oxide layers [36], attributing the composite structure of metal oxide layers shielding the underlying carbon structure with the disappearance of the (100) diffraction peaks. Oke et al. highlight that an electronic or structural change in MWCNTs decorated with some metal oxide produces a reconfiguration of the carbon network which is responsible for the disappearance of certain peaks [37]. For that reason, the disappearance of the (100) plane in Ga2O3@C and CuO@C indicates that the MWCNTs were successfully decorated by the gallium and copper oxides, altering the interlayer spacing and disrupting the graphitic structure [34,35,36,37]. Regarding the XRD pattern of MMO@C (Figure 1c), it shows several diffraction peaks indicating planes of different oxide components present which are decorating the MWCNTs [38,39,40]. A decrease in the 2θ 25° peak indexed to the (002) plane of the hexagonal graphite structure of the MWCNTs is also observed along with well-defined diffraction peaks belonging to Ga2O3 and CuO, and Cu(OH)2 and Cu2O, as segregated phases [41,42]. The small number of sharp peaks in Figure 1c also indicates the presence of heterogeneous amorphous species, which could play significant roles in determining the final characteristics of these materials and are essential for their application in various fields [43]. The functional groups present in carbon nanotubes (such as carboxyl groups (COOH)) interact with the surface of the oxide nanoparticles, generating interactions that modify the electronic configuration and the arrangement of the atoms in the surface layers of the oxide, altering its crystalline network or deformation (strain) in the crystal lattice of the oxide. This deformation results in a change in the interplanar distances, shifting the peaks and making them wider and less intense.
Since the MMO@C electrocatalyst has several mixed oxides, XPS analyses of C 1s, O 1s, Ga 2p, and Cu 2p were carried out to investigate their surface chemical compositions. The binding energies for each element are summarized in Table 1. Surface C 1s (Figure S1) corresponds to the oxygenated functional groups and the O1s, Ga 2p, and Cu 2p spectra shown in Figure 2. The O 1s split into three peaks using the combination of Gaussian and Lorentzian deconvolution, indicating surface oxygen species at 532.0 eV (57%), 532.9 eV (6%), and 534.0 eV (37%) associated with oxygen bonds with different valence states. The Ga 2p3/2 energy level is almost a single peak that can be split into two surface species at 1118.7 eV (6%) and 1119.6 (94%). Similarly to O 1s analysis, no signal of Ga3+ associated with the CuGa2O4 spinel structure at 1117.3 eV is observed [44], indicating surface Ga2O3. The respective Cu 2p3/2 spectra indicate that Cu is present as a mixture of surface Cu2+ and Cu+ species. The three deconvolute peaks of Cu 2p3/2 spectra indicate surface Cu+ at 932.7 eV (33%) and Cu2+ as CuO at 934.2 eV (61%) and Cu(OH)2 at 936.0 eV (6%). The surface Ga/Cu atomic ratio calculated by comparing the corrected area under the curves (Table 2) was 6, which was larger than the nominal 2 for the formation of CuGa2O4 spinel. The Ga/Cu atomic ratio for MMO@C determined by EDX is 7.1 (Figure S2, Table S1), close to the surface Ga/Cu atomic ratio of 6 determined by XPS. The small difference obtained using both techniques is mainly because EDX is not a surface-sensitive technique, and its ratios are intermediate between the bulk and surface ratios [45]. Therefore, the surface enrichment of Ga and the presence of @C inhibits the formation of the spinel structure (CuGa2O4). Given the above, the proposed CuGa2O4@C corresponds to a mix of surface Ga2O3-CuO-Cu2O-Cu(OH)2@C and for the sake of clarity will be labeled as MMO@C.
The TEM images in Figure 3 provide morphological information about the surface of the synthesized materials. The elongated cylindrical structure characteristic of undecorated MWCNTs (@C) and morphological differences for the decorated systems can be noted. The Ga2O3@C electrocatalyst shows particle agglomerations covering a large part of the MWCNT surface, in contrast to the CuO@C electrocatalyst which presents isolated particles covering a considerably smaller portion of the MWCNT surface. Like the Ga2O3@C system, the MMO@C electrocatalyst exhibits similar particle agglomerations, indicating that the surface of the MMO@C is predominantly decorated with Ga2O3. This information is consistent with the surface atomic percentages measured by XPS, which reveal a surface enrichment of Ga with a Ga/Cu ratio equal to 6 (Table 2). The FESEM images in Figure 4 compare the morphology of the @C as blank and the MMO@C electrocatalyst. Similarly to the TEM analysis, the MMO@C electrocatalyst shows well-distributed agglomerations on the MWCNTs. The agglomerations have an average size of 8 nm, implying the formation of nanoparticles that decorate the surface of the MWCNTs. These nanoparticles correspond to the mixed metal oxides with no formation of spinel structure, according to X-ray (Figure 1) and XPS (Figure 2) analyses. Both the nanometer size of the particles and the surface distribution [46,47] added to their intrinsic electrochemical and physicochemical properties can influence their electrocatalytic performance towards the HER. Additionally, the BET surface area (SBET) and pore volume were measured for the @C, CuO@C, Ga2O3@C, and MMO@C electrocatalysts (Table 3). The N2 physisorption isotherms are shown in Figure S3, which shows a typical type II isotherm of non-porous materials. It is observed that the MMO@C electrocatalyst presents a larger SBET (62 m2 g−1) compared to the Ga2O3@C (41 m2 g−1) and CuO@C (22 m2 g−1) electrocatalysts. Furthermore, the three decorated systems increased their SBET compared to the @C undecorated system (13 m2 g−1). It is seen that the pore volume increases consistently with increasing SBET. The increase in the SBET for the MMO@C electrocatalyst may be related to the formation of different surface species (Ga2O3; CuO; Cu2O and Cu(OH)2). The largest surface area for the MMO@C electrocatalyst allows a greater exposure of its available electroactive sites to improve the performance towards electrocatalytic reactions [48].

2.2. Electrochemical Characterization

Figure 5a shows the cyclic voltammograms for the @C and MMO@C electrocatalysts towards a ferri/ferrocyanide redox couple, where it is observed that ∆Ep values are 252 mV for MMO@C and 465 mV for @C. Thus, both systems are electrochemically quasi-reversible, and due to the lower ∆Ep value MMO@C has a higher electron transfer rate than @C [49]. Likewise, MMO@C presents a peak current approximately 10 times higher compared to @C, which indicates that mixed metal decoration on CNTs considerably improves the electrochemical properties of the electrocatalyst. In this way, MMO@C has the capacity to convert a greater amount of electroactive species [50]. Figure 5b shows the scan rate study of the MMO@C electrocatalyst towards a ferri/ferrocyanide redox couple between 0.025 and 0.200 V s−1. With the values of peak current (Ip) and scan rate (υ) (Figure 5b), the Log (Ip) vs. Log (υ) graph in Figure 5c is obtained, and the line equation shows a slope of 0.516. Since the slope is very close to 0.5, the system is diffusion-controlled [51]. Then, the slope obtained in Figure 5d corresponding to the linear equation Ip vs. υ½ is 0.0072 is replaced in the Randles–Sevcik equation for reversible systems controlled by diffusion (Equation (1)), which is as follows:
I p υ 1 / 2 =   2.69 × 10 5   n 3 / 2   C o   D 1 / 2   A
where the number of electrons transferred (n) is 1; Co is the concentration of the electroactive species (1 × 10−4 mol cm−3); D is the diffusion coefficient, which is 6.5 × 10−6 cm2 s−1 for the ferrocyanide ion [52]; and A is the electroactive area to be determined. Replacing values, the MMO@C paste electrode has an electroactive area of 0.105 cm2, being around 3.4 times larger than its geometric area (0.031 cm2). The above implies a great availability of active sites to carry out electrochemical processes [53].

2.3. Electrocatalytic Study of the HER

Figure 6a shows the linear voltammograms towards the HER in PBS buffer at neutral pH for the MMO@C, Ga2O3@C, CuO@C, and @C paste electrodes. The electrocatalytic performance summarized in Table 3 indicates that a large improvement in terms of onset potential (Eo) was obtained for all three decorated systems compared to @C, and differences in their electrocatalytic behavior is clearly observed. Particularly, MMO@C (Eo = 0.12 V vs. RHE) presents an outstanding improvement compared to its respective simple metal oxides Ga2O3@C (Eo = −0.33 V vs. RHE) and CuO@C (Eo = −0.38 V vs. RHE). Regarding MMO@C, its Eo value is 450 mV and 500 mV higher than Ga2O4@C and CuO@C, respectively, proving to be energetically more favorable for initiating the HER. The large synergistic effect for the MMO@C electrocatalyst is not only expressed through its remarkable Eo value but also reaches a higher current when the potential magnitude is high (E ~ −1 V) (Figure 6a). It can be observed that the MMO@C electrocatalyst at the onset of the current drop associated with the HER exhibits a parallel cathodic process centered at ~ −0.25 V. To understand the cause of this cathodic process, the electrocatalytic performances of the MMO@C, Ga2O3@C, and CuO@C electrocatalysts were studied in 0.1 M KOH solution (Figure 6b). Analogously to what occurs in a neutral medium (Figure 6a), the synergistic effect between Cu and Ga in MMO@C is maintained under alkaline conditions for both current and potential. However, a great difference is evident for the individual metal oxides in alkaline medium, obtaining an electrocatalytic improvement of CuO@C over Ga2O3@C. Additionally, in Figure 6b the CuO@C and MMO@C paste electrodes show a cathodic process centered at ~ 0 V vs. RHE, with a more pronounced reduction process for MMO@C. This result indicates that the presence of Ga in MMO@C favors a potential shift towards positive values, facilitating the anodic process associated with the presence of Cu. This anodic process is also observed at neutral pH for the MMO@C electrocatalyst centered at ~ −0.25 V, but not for CuO@C (Figure 6a). According to the Pourbaix diagram for copper [54], when a potential sweep is applied towards negative values in neutral and alkaline conditions there is a tendency for Cu(I) species to form Cu(s). The XRD results (Figure 1) indicate that the MMO@C electrocatalyst corresponds to a mixture of metals oxides. The Ga/Cu surface atomic ratio of 6 for MMO@C obtained by XPS indicates a surface enrichment Ga and confirms the XRD results about the presence of segregated phases. The surface composition of the MMO@C electrode demonstrates that despite its lower abundance CuO can generate a notable electrocatalytic effect towards the HER (Figure 6a), helped by the presence of Ga2O3. The higher standard reduction potential value of the Cu2+/Cu+ pair (E° = +0.15 V) compared to the Ga3+/Ga2+ pair (E° = −0.65 V) indicates that when both metals are together gallium has a large tendency to remain in its oxidized form, donating electron density. Since the resonant system of the MWCNTs promotes the movement of electrons through their structure [28], the donation of electron density between both metals could also be favored by the conjugated π system of the MWCNTs, facilitating the electrochemical reduction of Cu2+ to Cu+ to produce the Cu(I) species on the electrode. According to the Pourbaix diagram [54], these potential and pH conditions produce Cu(s). The above is evident through the cathodic process in an alkaline medium centered at ~ 0 V, corresponding to Cu2+ + e → Cu+ (Figure 6b). In this way, the electrochemical reduction to produce Cu(I) facilitated by the electronic donation of Ga3+, in neutral and alkaline conditions, drives the generation of Cu(s) active sites on the MWCNTs, reducing the requirement for proton adsorption on the electrocatalyst [55].
A chronoamperometric study was performed for the MMO@C electrocatalyst as an electrolysis test to identify if the HER occurs during the cathodic process associated with CuO reduction. The voltametric profile in acidic conditions (0.5 M H2SO4) makes it easier to visualize the production of H2 bubbles on the electrode [56,57], as shown in Figure 7a. As expected, in acidic conditions higher magnitudes of currents are observed when applying a potential sweep towards negative values in an acid solution, indicating a greater H2 production than in neutral or alkaline conditions. The chronoamperometry performed at a fixed potential of −0.4 V and −0.6 V vs. RHE on the surface of MMO@C is presented in Figure 7b. At a fixed potential of −0.6 V, the characteristic noise associated with the formation of gaseous species identified as H2 bubbles is observed. This result verified that the HER occurs during the reduction process prior to ~−0.8 V (Figure 7a), where there is an inflection point and hydrogen production becomes more evident. Additionally, the open circuit potential (OCP) of the MMO@C electrocatalyst at pH = 7.0 is 0.72 V vs. RHE (Figure S4). When two metals are together in the same solution, the OCP corresponds to the highest possible potential difference without applying an external potential [58]. Therefore, the OCP defines the capacity of the electrode to oxidize or reduce [59]. In this sense, the electrochemical interaction of the different species in the MMO@C system tends to occur at a potential of 0.72 V, relatively far from the onset potential in a neutral medium (Figure 6a).
The Tafel slope was determined in an acidic medium to decrease the contribution of CuO reduction in the global reaction mechanism. In general terms, the Tafel slope represents the intrinsic activity of electrocatalysts toward the HER, and smaller Tafel slopes are related to faster HER rates [60]. Three possible reaction steps are usually accepted for the HER associated with the obtained Tafel slope values [60]. The Volmer steps in acid and alkaline medium (Equation (2)) correspond to the formation of hydrogen adsorbed on the electrocatalyst (MHads) through a process known as proton discharge. The Heyrovsky steps (Equation (3)) correspond to Hads binding to a proton or water in an alkaline medium, and through the contribution of electrons provided by the electrocatalyst gaseous H2 is generated. The Tafel steps in acid and alkaline medium (Equation (4)) are where two adsorbed hydrogens on nearby active sites (2MH) combine to give rise to H2. Considering that the Heyrovsky and Tafel correspond to electrochemical desorption reactions, it is proposed that H2 must be produced by a combination of Volmer with Heyrovsky or Tafel steps. If the Volmer step is rate limiting, the Tafel slope is close to 120 mV dec−1 and indicates a mechanism associated with slow kinetics. Whereas, if the Heyrovsky step is the rate-limiting reaction then the Tafel slopes must be closer to 40 mV dec−1, and if the process is limited by the Tafel step then the expected slope values are 30 mV dec−1 [61]. Even though the Tafel slope values guide the identification of HER mechanisms, it should be noted that these calculations are based on a set of assumptions that are not universally accepted as a clear identification of HER mechanisms [61,62]. According to the literature, it has been studied that some Ni/Zn and Ni/Al pressed powder electrodes have shown Tafel slopes over 120 mV dec−1, even reaching 307 mV dec−1 towards the HER, where the high Tafel slopes are related to a larger apparent surface area [63]. Additionally, it has been reported that some carbon paste electrodes have Tafel slopes between 186 and 508 mV dec−1 [64]. In accordance with the above, a Tafel slope of 240 mV dec−1 for a porous carbon electrode was found [65]. Within this context it has been studied that an excessive thickness of the catalyst layer influences the resistance to mass transport of H2 through the catalyst layer [62,66], resulting in high Tafel slopes, which is difficult to modulate when porous paste electrodes with a large surface area are used. However, it is known that a large surface area value, as in the case of CNT-based materials, can lead to a better electrocatalytic effect [29,67] and to obtaining higher faradaic currents [68] given the exposure of active sites available for the electrochemical process to occur.
In this work, the Tafel slope of the MMO@C electrocatalyst is 350 mV dec−1, which is higher than 120 mV dec−1. A high value of electroactive area was previously calculated for the MMO@C electrocatalyst (0.105 cm2, about 3.4 times larger than its geometric area), and this high value is associated with the high porosity and surface area of the decorated CNTs. On the one hand, as previously discussed, the large electrode surface may be related to a high value of the Tafel slope [62,63,66]. On the other hand, although the Tafel slope was obtained in an acidic medium, two reactions are occurring in parallel in the study area making it difficult to establish an HER mechanism associated only with the Tafel slope. Although the Tafel slope reveals slow global kinetics, the MMO@C electrocatalyst presents an excellent initiation potential in neutral medium (Figure 7a), being a good candidate to be coupled to renewable energy sources to produce green hydrogen.
Volmer   step   in   acid   medium :   H + + M + e     M H a d s . Volmer   step   in   alkaline   medium :   H 2 O + M + e     M H a d s + O H .
Heyrovsky   step   in   acid   medium :   M H + H + + e     M + H 2 . Heyrovsky   step   in   alkaline   medium :   M H + H 2 O + e     M + O H + H 2 .
Tafel   step   in   acid   and   alkaline   medium :   2 M H     2 M + H 2 .

2.4. Quantitative Study

To quantify the H2 production monitored by GC, the electrolysis was performed at a fixed potential of −0.8 V vs. RHE for 3 h using an MMO@C paste electrode in 0.5 M H2SO4 (pH = 0.3). Once again, an acidic solution was used, since having a higher concentration of H+ in the medium favors a better quantification of H2 gas. The plot of charge vs. time during the electrolysis (Figure S5) shows the progressive increase in charge magnitude on the electrode surface throughout the electrolysis, reaching a final charge of 83.9 C. The total volume of the sealed electrochemical cell is 30 mL; therefore 26 mL of solution was used, leaving a head space of 4 mL for the accumulation of the gaseous product. After 3 h, 50 μL of gas was extracted from the headspace to be injected into the GC to obtain a chromatographic area of 1,083,696 μV/s. Then, to calculate the amount of hydrogen produced during electrolysis, a calibration curve obtained in a previous work was used [69]. It was determined that the MMO@C electrode produces a large amount of 4.1 × 10−4 mol of H2, corresponding to 1.31 × 10−2 mol cm−2. The charge on the electrode surface at Eo is calculated from the ratio between faradaic current and time (Equation (5)), and time is obtained from Equation (6). These are as follows:
Q = I t
t = E o υ
where Q is the charge on the electrode surface at Eo (C), I is the Faradaic current at Eo (1.6 × 10−4 A), and t is the time at Eo. Given that Eo in an acidic medium is −0.13 V and the scan rate (υ) is 0.01 Vs−1, it is calculated that t = 13 s and therefore Q = 0.002 C. The number of active sites is calculated from Faraday’s Law of Electrolysis (Equation (7)), which is as follows:
N = Q n F
where N is the number of active sites, Q is the charge on the electrode surface at Eo previously calculated (0.002 C), F is the Faraday constant (96,485 C∙mol−1), and n is the number of electrons transferred (2 for HER). Thus, N = 1.04 × 10−8 mol. The turnover number (TON) measures the total activity of a catalyst and represents the total number of substrate molecules converted by a single catalyst molecule before it becomes inactive. The turnover frequency (TOF) measures the catalyst’s instantaneous efficiency and indicates the number of catalytic cycles, or molecules converted per active site per unit time. The TON and TOF for MMO@C electrocatalyst are obtained from the following Equations (8) and (9), respectively:
T O N = n H 2 N
T O F = T O N t
where n H 2 is the number of moles of hydrogen produced by the MMO@C electrocatalyst (4.1 × 10−4 moles), N = 1.04 × 10−8 mol, and t is electrolysis time (3 h). Replacing values, it is determined that TON = 39,423 and TOF = 13,141 h−1. Finally, faradaic efficiency FE (%) is obtained through Equation (10), which is as follows:
F E H 2 % =   n H 2 n   F q   100
Since the amount of hydrogen produced ( n H 2 ) is 4.1 × 10−4 mol, the number of electrons transferred for the HER (n) is 2, the Faraday constant (F) is 96,485 C∙mol−1, and the final charge (q) after electrolysis is 83.9 C (Figure S5), so FE (%) is 94.3%.

3. Materials and Methods

3.1. Chemicals

Multi-walled carbon nanotubes (MWCNTs) powder, sodium hydroxide (pellets for analysis), sodium chloride, and potassium chloride were purchased from Merck, Chile. Potassium ferricyanide (K3[Fe(CN)6]) 98.5%, potassium ferrocyanide (K4[Fe(CN)6]), potassium hydroxide, potassium dihydrogen phosphate 99.5%, disodium hydrogen phosphate 99%, sulfuric acid 98%, absolute ethanol, copper (II) nitrate, and gallium (III) acetylacetonate were obtained from Sigma-Aldrich, Chile. Ultra-pure water was provided by Millipore-Q system (18.2 MΩcm). All salts are of analytical grade.

3.2. Synthesis of Electrocatalysts

The chemical oxidation of MWCNTs was carried out using 1 g of MWCNTs sonicated in 150 mL of a 3:1 (v/v) H2SO4/HNO3 mixture for 6 h at 70 °C under a fume hood. The mixture was allowed to cool to room temperature, then filtered and washed with ultra-pure water until neutralization was complete (pH = 7). The obtained solid product was dried in an oven at 105 °C for 5 h. The synthesis of electrocatalysts was carried out by mixing 0.3 g of previously oxidized MWCNTs with 4.124 mmol of copper nitrate for CuO@C and 4.124 mmol of gallium acetylacetonate for Ga2O3@C, respectively. The mixtures were homogenized in an agate mortar and then sonicated for 20 min in ethanol. After evaporation of ethanol, the obtained solid was added to 35 mL of 0.05 M NaOH in a Teflon-lined stainless steel autoclave apparatus for hydrothermal treatment at 180 °C for 12 h. The resulting solid was filtered, washed with ultra-pure water (Milli Q), and finally dried in an oven at 105 °C for 5 h. For MMO@C synthesis, the same procedure described above was performed using 1.375 mmol of copper salt and 2.750 mmol of gallium salt. The characterization results indicate the presence of a mixture of metal oxides (MMOs) without a spinel structure. Therefore, the materials were labeled as CuO@C, Ga2O3@C, and MMO@C.

3.3. Paste Electrodes Preparation

The paste electrodes contain a mixture of decorated MWCNTs and mineral oil in a 7:3 (w/w) ratio. The paste electrodes were prepared with a 7:3 proportion (w/w) of the MWCNTs-decorated material (CuO@C, Ga2O3@C, and MMO@C) and mineral oil, respectively. The mixture was homogenized in an agar mortar, adding diethyl ether. Also, a blank electrode was prepared only with undecorated MWCNTs following the above methodology. When diethyl ether was completely evaporated, the pastes were packed into the Teflon hollow electrodes and then heated in oven at 90 °C for 5 h. Finally, electrodes were allowed to cool at room temperature, and the surfaces were polished to a smooth finish. All the measurements with the paste electrodes were reproducible after surface renewing, showing the same voltametric profile and a current change below 3% for each measurement. The geometric area of all the electrodes was 0.031 cm2.

3.4. Instrumentation

The electrocatalytic activity of electrodes was studied by using a three-compartment electrochemical cell. One compartment is available for the reference electrode Ag/AgCl (3 M KCl), a second compartment for the counter electrode (Pt spiral of great area), and a third compartment for the working electrode, which is separated from the counter electrode compartment by a porous medium. The transformation between Ag/AgCl and RHE reference electrodes is ERHE = EAg/AgCl + 0.059pH + E0Ag/AgCl; E0Ag/AgCl = 0.1976 V. Linear and cyclic voltametric analyses and electrolysis were performed with a CH Instruments 650E potentiostat. Electrolysis for H2 quantification was performed at −0.8 V vs. RHE fixed potential for the HER. Powder X-ray diffraction (PXRD) patterns were collected at room temperature on a Bruker D8 Advance diffractometer (Bruker, Billerica, MA, USA) equipped with a Cu Kα radiation source, in a range of 5° < 2θ < 80. For the morphological study, the electrocatalysts were analyzed by transmission electron microscopy (TEM) Talos F200X G2 (Thermo Fisher Scientific, Hillsboro, OR, USA) and field emission scanning electron microscopy (FE-SEM) with energy dispersive X-ray spectroscopy (EDX) QUANTA FEG250 (Thermo Fisher Scientific, Hillsboro, OR, USA). X-ray photoelectron spectroscopy (XPS) of the electrocatalysts was recorded on a VG Escalab 200R electron spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) using a Mg Kα (1253.6 eV) photon source. The binding energies (BEs) were referenced to the C 1s level of the carbon support at 284.8 eV. An estimated error of ± 0.1 eV can be assumed for all measurements. The intensities of the peaks were calculated from the respective peak areas after background subtraction and spectrum fitting by standard computer-based statistical analysis, which included fitting the experimental spectra. Nitrogen physisorption isotherms of the electrocatalysts were obtained at −196 °C using a Micromeritics TriStar II 3020 instrument (Micromeritics Instrument Corporation, Norcross, GA, USA) to evaluate the BET specific surface area (SBET). Prior to the measurements, the samples were degassed under vacuum at 200 °C for 3 h in a nitrogen flow. SBET was calculated from the adsorption branch in the range 0.05 ≤ P/Po ≤ 0.25. Hydrogen production was monitored using a completely sealed two compartment electrolysis cell for the most electrocatalytic system (MMO@C paste electrode). During the 3 h of electrolysis, the 0.5 M Ar-saturated sulfuric acid solution (pH = 0.3) was kept in agitation with a magnetic stirring bar, aiming to favor hydrogen bubbles spread to the head space. During the experiment, 50 μL of gas was extracted from the head space and analyzed by gas chromatography (GC-2030 Plus, Shimadzu, Japan) with a thermal conductivity detector (TCD) Subsequently, a correction was applied considering the total volume of the head space (4 mL). The GC operated under the following working conditions: injector temperature = 50 °C; P = 248 kPa; column temperature = 25 °C; flow = 5.5 mL/min; carrier gas = Ar; and detector temperature = 200 °C.

4. Conclusions

Novel paste electrodes based on MWCNTs decorated with pure single oxides (Ga and Cu) and mixed metal oxides (Ga with Cu) were synthesized to be used in the electrocatalytic reaction of the HER. Significant differences in the structural and surface properties of the MMO@C paste electrode relative to the pure electrodes were found. The excellent electrocatalytic performance of the MMO@C paste electrode is associated with the synergistic effect between gallium and copper oxides on the electrode surface and the formation of surface mixed species of Ga2O3-CuO-Cu2O-Cu(OH)2@C. It is worth noting that although the Tafel slope reveals slow reaction kinetics, the H2 production of 1.31∙10−2 mol∙cm−2 with a TON of 39423, a TOF of 13,141 h−1, and a faradaic efficiency of 94.3% indicates the potential of the MMO@C electrocatalyst to produce green H2. Furthermore, its outstanding onset potential of +0.12 V vs. RHE in a neutral medium allows it to be coupled to renewable energy production systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms26189057/s1.

Author Contributions

Conceptualization, C.B., S.M. and G.P.; methodology, D.A., J.I. and L.G.; formal analysis, C.B., G.P., G.R. and L.G.; resources, G.P., G.R. and L.G.; data curation, S.M., D.A. and J.I.; writing—original draft preparation, L.G.; writing—review and editing, C.B., S.M., G.P. and G.R.; supervision, C.B.; project administration, L.G. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support provided by the FONDECYT Project No.: 11220164; Scholarship ANID Project No.: 21210433; FONDECYT Project No.: 1220107; Millennium Institute on Green Ammonia as Energy Vector, MIGA, ICN 2021–023; Proyecto Anillo Tecnológico ANID/ACT240020; and ANID Postdoctoral Project No.: 3220040.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Squadrito, G.; Maggio, G.; Nicita, A. The green hydrogen revolution. Renew. Energy 2023, 216, 119041. [Google Scholar] [CrossRef]
  2. Oliveira, A.M.; Beswick, R.R.; Yan, Y. A green hydrogen economy for a renewable energy society. Curr. Opin. Chem. Eng. 2021, 33, 100701. [Google Scholar] [CrossRef]
  3. Moses, J.K.; Tivfa, T.A.; Qasem, N.A.; Alquaity, A.B. Combustion characteristics of hydrogen, ammonia, and their blends: A review. Fuel 2025, 388, 134404. [Google Scholar] [CrossRef]
  4. Tran, D.T.; Tran, P.K.L.; Malhotra, D.; Nguyen, T.H.; Nguyen, T.T.A.; Duong, N.T.A.; Kim, N.H.; Lee, J.H. Current status of developed electrocatalysts for water splitting technologies: From experimental to industrial perspective. Nano Converg. 2025, 12, 9. [Google Scholar] [CrossRef]
  5. Greitzer, M. Iridium and Platinum Availability for Electrolyser Production up to 2030. In Powerfuels; Springer: Cham, Switzerland, 2025; pp. 257–279. [Google Scholar]
  6. Feidenhans’l, A.A.; Regmi, Y.N.; Wei, C.; Xia, D.; Kibsgaard, J.; King, L.A. Precious metal free hydrogen evolution catalyst design and application. Chem. Rev. 2024, 124, 5617–5667. [Google Scholar] [CrossRef] [PubMed]
  7. Al-Naggar, A.H.; Salah, A.; Al-Hejri, T.M.; Kamble, C.; Jadhav, V.V.; Shaikh, S.F.; Mane, R.S. Self-supported iron-doped nickel oxide multifunctional electrodes for highly efficient energy storage and overall water-splitting uses. J. Energy Storage 2024, 86, 111363. [Google Scholar] [CrossRef]
  8. Tu, J.; Zhang, M.; Li, M.; Li, J.; Zhi, L. Phosphorus-doped nickel cobalt oxide (NiCo2O4) wrapped in 3D hierarchical hollow N-doped carbon nanoflowers as highly efficient bifunctional electrocatalysts for overall water splitting. J. Colloid Interface Sci. 2024, 668, 243–251. [Google Scholar] [CrossRef]
  9. Sajid, M.; Qayyum, W.; Farhan, A.; Qamar, M.A.; Nawaz, H. Progress in the development of copper oxide-based materials for electrochemical water splitting. Int. J. Hydrogen Energy 2024, 62, 209–227. [Google Scholar] [CrossRef]
  10. Farooq, A.; Khalil, S.; Basha, B.; Habib, A.; Al-Buriahi, M.; Warsi, M.F.; Yousaf, S.; Shahid, M. Electrochemical investigation of C-doped CoFe2O4/Fe2O3 nanostructures for efficient electrochemical water splitting. Int. J. Hydrogen Energy 2024, 51, 1318–1332. [Google Scholar] [CrossRef]
  11. Gaur, A.; Sharma, J.; Lim, D.H.; Lee, H.I.; Han, H. Recent Advances in Electronic Structure Modifications of Layered Double Hydroxide (LDH) for the Water Splitting Application. ChemCatChem 2025, 17, e202401584. [Google Scholar] [CrossRef]
  12. Yang, C.; Lu, T.; Zhang, L. Probing the structural evolution of cobalt hydroxide in electrochemical water splitting. Chem. Commun. 2024, 60, 10326–10329. [Google Scholar] [CrossRef]
  13. Xu, X.; Qiao, F.; Liu, Y.; Liu, W. Preparation of Cu (OH)2/Cu2S arrays for enhanced hydrogen evolution reaction. Battery Energy 2024, 3, 20230060. [Google Scholar] [CrossRef]
  14. Liu, Z.; Zhang, T.; Tang, T.; Li, J. Phytic acid treated nickel–iron hydroxide promotes efficient electrocatalytic overall water splitting. New J. Chem. 2024, 48, 10769–10775. [Google Scholar] [CrossRef]
  15. de Almeida, J.C.; Wang, Y.; Rodrigues, T.A.; Nunes, P.H.H.; de Mendonça, V.R.; Falsetti, P.H.E.; Savazi, L.V.; He, T.; Bardakova, A.V.; Rudakova, A.V.; et al. Copper-based Materials for Photo and Electrocatalytic Process: Advancing Renewable Energy and Environmental Applications. Adv. Funct. Mater. 2025, 2502901. [Google Scholar] [CrossRef]
  16. Jena, S.S.; Tripathy, S.K.; Mandre, N.R.; Venugopal, R.; Farrokhpay, S. Sustainable use of copper resources: Beneficiation of low-grade copper ores. Minerals 2022, 12, 545. [Google Scholar] [CrossRef]
  17. Kannimuthu, K.; Sangeetha, K.; Sankar, S.S.; Karmakar, A.; Madhu, R.; Kundu, S. Investigation on nanostructured Cu-based electrocatalysts for improvising water splitting: A review. Inorg. Chem. Front. 2021, 8, 234–272. [Google Scholar] [CrossRef]
  18. Awad, M.; Nawwar, M.; Zhitomirsky, I. Synergy of Charge Storage Properties of CuO and Polypyrrole in Composite CuO-Polypyrrole Electrodes for Asymmetric Supercapacitor Devices. ACS Appl. Energy Mater. 2024, 7, 5572–5581. [Google Scholar] [CrossRef]
  19. Lu, H.; Song, S.; Jia, Q.; Liu, G.; Jiang, L. Advances in Cu2O-based photocathodes for photoelectrochemical water splitting. Acta Phys.-Chim. Sin. 2024, 40, 2304035. [Google Scholar] [CrossRef]
  20. Seo, Y.J.; Arunachalam, M.; Ahn, K.-S.; Kang, S.H. Integrating heteromixtured Cu2O/CuO photocathode interface through a hydrogen treatment for photoelectrochemical hydrogen evolution reaction. Appl. Surf. Sci. 2021, 551, 149375. [Google Scholar]
  21. Son, H.; Lee, J.-H.; Uthirakumar, P.; Van Dao, D.; Soon, A.; Lee, I.-H. Facile synthesis of flexible and scalable Cu/Cu2O/CuO nanoleaves photoelectrodes with oxidation-induced self-initiated charge-transporting platform for photoelectrochemical water splitting enhancement. J. Alloys Compd. 2023, 942, 169094. [Google Scholar] [CrossRef]
  22. Mumtazah, N.; Chan, C.-H.; Catherine, S.; Pham, M.-T.H.; Choi, J.; Yoo, J.S.; Chiang, C.Y. Modulating surface chemistry of copper oxides through annealing environment for enhanced selective HMF oxidation: A DFT-electrochemical approach. J. Environ. Chem. Eng. 2024, 12, 113444. [Google Scholar] [CrossRef]
  23. Dong, N.; Wu, J.; Wu, D.; Jiang, G.; Shi, J.; Chang, S. Cu/Cu2O/CuO hierarchical nanosheets for electrocatalytic conversion of nitrate to ammonia. ACS Appl. Nano Mater. 2024, 7, 9269–9277. [Google Scholar] [CrossRef]
  24. Liu, C.; Ma, J.; Fu, Z.; Zhao, P.; Bai, M.; Gao, Y.; Zhao, M.; He, Y.; Xiao, H.; Jia, J. Gallium-based materials for electrocatalytic and photocatalytic hydrogen evolution reaction. Int. J. Hydrogen Energy 2024, 73, 490–509. [Google Scholar] [CrossRef]
  25. Kakoria, A.; Devi, B.; Anand, A.; Halder, A.; Koner, R.R.; Sinha-Ray, S. Gallium oxide nanofibers for hydrogen evolution and oxygen reduction. ACS Appl. Nano Mater. 2018, 2, 64–74. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Hui, Z.X.; Zhou, H.Y.; Zai, S.F.; Wen, Z.; Li, J.C.; Yang, C.C.; Jiang, Q. Ga doping enables superior alkaline hydrogen evolution reaction performances of CoP. Chem. Eng. J. 2022, 429, 132012. [Google Scholar] [CrossRef]
  27. Chen, L.; Chen, J.; Fu, W.; Chen, J.; Wang, D.; Xiao, Y.; Xi, S.; Ji, Y.; Wang, L. Energy-efficient CO(2) conversion to multicarbon products at high rates on CuGa bimetallic catalyst. Nat. Commun. 2024, 15, 7053. [Google Scholar] [CrossRef]
  28. Saboor, F.H.; Ataei, A. Decoration of metal nanoparticles and metal oxide nanoparticles on carbon nanotubes. Adv. J. Chem. Sect. A 2024, 7, 122. [Google Scholar]
  29. Tafete, G.A.; Thothadri, G.; Abera, M.K. A review on carbon nanotube-based composites for electrocatalyst applications. Fuller. Nanotub. Carbon Nanostruct. 2022, 30, 1075–1083. [Google Scholar] [CrossRef]
  30. Noor, T.; Yaqoob, L.; Iqbal, N. Recent advances in electrocatalysis of oxygen evolution reaction using noble-metal, transition-metal, and carbon-based materials. ChemElectroChem 2021, 8, 447–483. [Google Scholar] [CrossRef]
  31. Wang, J. Carbon-nanotube based electrochemical biosensors: A review. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2005, 17, 7–14. [Google Scholar] [CrossRef]
  32. Playford, H.Y.; Hannon, A.C.; Tucker, M.G.; Dawson, D.M.; Ashbrook, S.E.; Kastiban, R.J.; Sloan, J.; Walton, R.I. Characterization of structural disorder in γ-Ga2O3. J. Phys. Chem. C 2014, 118, 16188–16198. [Google Scholar] [CrossRef]
  33. Tunell, G.; Posnjak, Ε.; Ksanda, C. Geometrical and optical properties, and crystal structure of tenorite. Z. Für Krist.-Cryst. Mater. 1935, 90, 120–142. [Google Scholar] [CrossRef]
  34. Goh, K.; Jiang, W.; Karahan, H.E.; Zhai, S.; Wei, L.; Yu, D.; Fane, A.G.; Wang, R.; Chen, Y. All-carbon nanoarchitectures as high-performance separation membranes with superior stability. Adv. Funct. Mater. 2015, 25, 7348–7359. [Google Scholar] [CrossRef]
  35. Zhu, H.; Li, K.; Chen, M.; Cao, H.; Wang, F. A novel metal–organic framework route to embed Co nanoparticles into multi-walled carbon nanotubes for effective oxygen reduction in alkaline media. Catalysts 2017, 7, 364. [Google Scholar] [CrossRef]
  36. Abdullah, T.A.; Juzsakova, T.; Rasheed, R.T.; Salman, A.D.; Sebestyen, V.; Domokos, E.; Sluser, B.; Cretescu, I. Polystyrene-Fe3O4-MWCNTs nanocomposites for toluene removal from water. Materials 2021, 14, 5503. [Google Scholar] [CrossRef]
  37. Oke, J.A.; Idisi, D.O.; Sarma, S.; Moloi, S.J.; Ray, S.C.; Chen, K.H.; Ghosh, A.; Shelke, A.; Pong, W.F. Electronic, electrical, and magnetic behavioral change of SiO2-NP-decorated MWCNTs. ACS Omega 2019, 4, 14589–14598. [Google Scholar] [CrossRef]
  38. Domagała, K.; Borlaf, M.; Kata, D.; Graule, T. Synthesis of copper-based multi-walled carbon nanotube composites. Arch. Metall. Mater. 2020, 65, 157–162. [Google Scholar] [CrossRef]
  39. Lavalley, J.; Daturi, M.; Montouillout, V.; Clet, G.; Areán, C.O.; Delgado, M.R.; Sahibed-Dine, A. Unexpected similarities between the surface chemistry of cubic and hexagonal gallia polymorphs. Phys. Chem. Chem. Phys. 2003, 5, 1301–1305. [Google Scholar] [CrossRef]
  40. Khadivi, A.H.; Vahdati, K.J.; Haddad, S.M. Facile synthesis of copper oxide nanoparticles using copper hydroxide by mechanochemical process. J. Ultrafine Grained Nanostruct. Mater. 2015, 48, 37–44. [Google Scholar]
  41. Neuburger, M. Präzisionsmessung der Gitterkonstante von Cuprooxyd Cu2O. Z. Für Phys. 1931, 67, 845–850. [Google Scholar] [CrossRef]
  42. Gao, S.; Xing, H.; Li, Y.; Wang, H. Synthesis of Cu2O/multi-walled carbon nanotube hybrid material and its microwave absorption performance. Res. Chem. Intermed. 2018, 44, 3425–3435. [Google Scholar] [CrossRef]
  43. Gicha, B.B.; Tufa, L.T.; Choi, Y.; Lee, J. Amorphous Ni1–xFex Oxyhydroxide Nanosheets with Integrated Bulk and Surface Iron for a High and Stable Oxygen Evolution Reaction. ACS Appl. Energy Mater. 2021, 4, 6833–6841. [Google Scholar] [CrossRef]
  44. Huseynzade, F.; Akbaba, Y.; Shawuti, S.; Xan, M.M. Investigating the photocatalytic efficacy of a porous cuprous gallate (CuGa2O4) thick film. Phys. Scr. 2024, 99, 125905. [Google Scholar] [CrossRef]
  45. Pecchi, G.; Campos, C.M.; Jiliberto, M.G.; Delgado, E.J.; Fierro, J.L.G. Effect of additive Ag on the physicochemical and catalytic properties of LaMn0.9Co0.1O3.5 perovskite. Appl. Catal. A Gen. 2009, 371, 78–84. [Google Scholar] [CrossRef]
  46. Joudeh, N.; Linke, D. Nanoparticle classification, physicochemical properties, characterization, and applications: A comprehensive review for biologists. J. Nanobiotechnol. 2022, 20, 262. [Google Scholar] [CrossRef]
  47. Ferdiana, N.A.; Bahti, H.H.; Kurnia, D.; Wyantuti, S. Synthesis, characterization, and electrochemical properties of rare earth element nanoparticles and its application in electrochemical nanosensor for the detection of various biomolecules and hazardous compounds: A review. Sens. Bio-Sens. Res. 2023, 41, 100573. [Google Scholar] [CrossRef]
  48. Wu, T.; Han, M.Y.; Xu, Z.J. Size effects of electrocatalysts: More than a variation of surface area. ACS Nano 2022, 16, 8531–8539. [Google Scholar] [CrossRef] [PubMed]
  49. Klingler, R.J.; Kochi, J.K. Electron-transfer kinetics from cyclic voltammetry. Quantitative description of electrochemical reversibility. J. Phys. Chem. 1981, 85, 1731–1741. [Google Scholar] [CrossRef]
  50. Martínez, R.; Ramírez, M.T.; González, I. Voltammetric characterization of carbon paste electrodes with a nonconducting binder. Part I: Evidence of the influence of electroactive species dissolution into the paste on the voltammetric response. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 1998, 10, 336–342. [Google Scholar] [CrossRef]
  51. Brett, C.M.; Brett, O. Principles, methods, and applications. Electrochemistry 1993, 67, 444. [Google Scholar]
  52. Bard, A.J.; Faulkner, L.R.; White, H.S. Electrochemical Methods: Fundamentals and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2022. [Google Scholar]
  53. Salinas-Torres, D.; Huerta, F.; Montilla, F.; Morallón, E. Study on electroactive and electrocatalytic surfaces of single walled carbon nanotube-modified electrodes. Electrochim. Acta 2011, 56, 2464–2470. [Google Scholar] [CrossRef]
  54. Beverskog, B.; Puigdomenech, I. Revised Pourbaix diagrams for copper at 25 to 300 °C. J. Electrochem. Soc. 1997, 144, 3476. [Google Scholar] [CrossRef]
  55. Sakong, S.; Groß, A. Dissociative adsorption of hydrogen on strained Cu surfaces. Surf. Sci. 2003, 525, 107–118. [Google Scholar] [CrossRef]
  56. Zhong, G.; Cheng, T.; Shah, A.H.; Wan, C.; Huang, Z.; Wang, S.; Leng, T.; Huang, Y.; Goddard, W.A., III; Duan, X. Determining the hydronium pK α at platinum surfaces and the effect on pH-dependent hydrogen evolution reaction kinetics. Proc. Natl. Acad. Sci. USA 2022, 119, e2208187119. [Google Scholar] [CrossRef]
  57. Carneiro-Neto, E.B.; Lopes, M.C.; Pereira, E.C. Simulation of interfacial pH changes during hydrogen evolution reaction. J. Electroanal. Chem. 2016, 765, 92–99. [Google Scholar] [CrossRef]
  58. Munoz, A.I.; Espallargas, N.; Mischler, S. Tribocorrosion; Springer International Publishing: Cham, Switzerland, 2020. [Google Scholar]
  59. Shin, S.J.; Kim, J.Y.; An, S.; Chung, T.D. Recent advances in electroanalytical methods for electroorganic synthesis. Curr. Opin. Electrochem. 2022, 35, 101054. [Google Scholar] [CrossRef]
  60. Tang, C.; Sun, A.; Xu, Y.; Wu, Z.; Wang, D. High specific surface area Mo2C nanoparticles as an efficient electrocatalyst for hydrogen evolution. J. Power Sources 2015, 296, 18–22. [Google Scholar] [CrossRef]
  61. Conway, B.; Tilak, B. Interfacial processes involving electrocatalytic evolution and oxidation of H2, and the role of chemisorbed H. Electrochim. Acta 2002, 47, 3571–3594. [Google Scholar] [CrossRef]
  62. Wan, C.; Ling, Y.; Wang, S.; Pu, H.; Huang, Y.; Duan, X. Unraveling and resolving the inconsistencies in tafel analysis for hydrogen evolution reactions. ACS Cent. Sci. 2024, 10, 658–665. [Google Scholar] [CrossRef]
  63. Hitz, C.; Lasia, A. Experimental study and modeling of impedance of the her on porous Ni electrodes. J. Electroanal. Chem. 2001, 500, 213–222. [Google Scholar] [CrossRef]
  64. Ahlberg, E.; Ásbjörnsson, J. Carbon paste electrodes in mineral processing: An electrochemical study of sphalerite. Hydrometallurgy 1994, 36, 19–37. [Google Scholar] [CrossRef]
  65. Austin, L. Tafel slopes for flooded diffusion electrodes. Trans. Faraday Soc. 1964, 60, 1319–1324. [Google Scholar] [CrossRef]
  66. Kemppainen, E.; Halme, J.; Hansen, O.; Seger, B.; Lund, P.D. Two-phase model of hydrogen transport to optimize nanoparticle catalyst loading for hydrogen evolution reaction. Int. J. Hydrogen Energy 2016, 41, 7568–7581. [Google Scholar] [CrossRef]
  67. Agüí, L.; Yáñez-Sedeño, P.; Pingarrón, J.M. Role of carbon nanotubes in electroanalytical chemistry: A review. Anal. Chim. Acta 2008, 622, 11–47. [Google Scholar] [CrossRef] [PubMed]
  68. Gong, K.; Yan, Y.; Zhang, M.; Su, L.; Xiong, S.; Mao, L. Electrochemistry and electroanalytical applications of carbon nanotubes: A review. Anal. Sci. 2005, 21, 1383–1393. [Google Scholar] [CrossRef][Green Version]
  69. Gidi, L.; Arce, R.; Ibarra, J.; Isaacs, M.; Aguirre, M.; Ramírez, G. Hydrogen evolution reaction highly electrocatalyzed by MWCNT/N-octylpyridinum hexafluorophosphate metal-free system. Electrochim. Acta 2021, 372, 137859. [Google Scholar] [CrossRef]
Figure 1. Powder X-ray diffraction (PXRD) patterns of (a) Ga2O3@C (b) CuO@C, and (c) MMO@C.
Figure 1. Powder X-ray diffraction (PXRD) patterns of (a) Ga2O3@C (b) CuO@C, and (c) MMO@C.
Ijms 26 09057 g001
Figure 2. XP spectra of O1s, Cu 2p3/2, and Ga 2p3/2 for the MMO@C electrocatalyst.
Figure 2. XP spectra of O1s, Cu 2p3/2, and Ga 2p3/2 for the MMO@C electrocatalyst.
Ijms 26 09057 g002
Figure 3. TEM images for @C, CuO@C, Ga2O3@C, and MMO@C electrocatalysts.
Figure 3. TEM images for @C, CuO@C, Ga2O3@C, and MMO@C electrocatalysts.
Ijms 26 09057 g003
Figure 4. FESEM images for @C and MMO@C electrocatalysts.
Figure 4. FESEM images for @C and MMO@C electrocatalysts.
Ijms 26 09057 g004
Figure 5. (a) Cyclic voltammograms for @C and MMO@C electrocatalysts, (b) scan rate study (0.025; 0.050; 0.075; 0.100; 0.125; 0.150; 0.175; and 0.200 V s−1), (c) plot Log (Ip) vs. Log (υ) obtained from Figure 5b, and (d) plot Ip vs. υ1/2 obtained from Figure 5b for MMO@C electrocatalyst towards a 0.1 M ferri/ferrocyanide redox couple, N2 sat.
Figure 5. (a) Cyclic voltammograms for @C and MMO@C electrocatalysts, (b) scan rate study (0.025; 0.050; 0.075; 0.100; 0.125; 0.150; 0.175; and 0.200 V s−1), (c) plot Log (Ip) vs. Log (υ) obtained from Figure 5b, and (d) plot Ip vs. υ1/2 obtained from Figure 5b for MMO@C electrocatalyst towards a 0.1 M ferri/ferrocyanide redox couple, N2 sat.
Ijms 26 09057 g005
Figure 6. (a) Linear voltammograms in PBS buffer pH = 7.0; (b) linear voltammograms in 0.1 M KOH, N2 sat., and υ = 0.01 V s−1.
Figure 6. (a) Linear voltammograms in PBS buffer pH = 7.0; (b) linear voltammograms in 0.1 M KOH, N2 sat., and υ = 0.01 V s−1.
Ijms 26 09057 g006
Figure 7. (a) Linear voltammogram towards the HER in 0.5 M H2SO4 using the MMO@C electrocatalyst. (b) Chronoamperometric study in 0.5 M H2SO4 using the MMO@C electrocatalyst at E = −0.4 and −0.6 V. (c) The Tafel slope (350 mV dec−1) for the MMO@C electrocatalyst.
Figure 7. (a) Linear voltammogram towards the HER in 0.5 M H2SO4 using the MMO@C electrocatalyst. (b) Chronoamperometric study in 0.5 M H2SO4 using the MMO@C electrocatalyst at E = −0.4 and −0.6 V. (c) The Tafel slope (350 mV dec−1) for the MMO@C electrocatalyst.
Ijms 26 09057 g007
Table 1. Binding energies (eVs) and composition (%) for each energy level and species in the MMO@C electrocatalyst.
Table 1. Binding energies (eVs) and composition (%) for each energy level and species in the MMO@C electrocatalyst.
Energy LevelSpeciesBinding Energy
(eV)
Composition
(%)
C 1sC-C284.750
C-O, C-OH286.2 28
C=O287.66
O-C=O288.916
O 1sC=O532.057
C-OH532.96
O-C=O534.037
Ga 2pGa1+ (Ga2O)1118.76
Ga3+ (Ga2O3)1119.694
Cu 2pCu1+ (Cu2O)932.733
Cu2+ (CuO)934.261
Cu2+ (Cu(OH)2)936.06
Table 2. Atomic percentages of elements obtained by XP spectra for the MMO@C electrocatalyst.
Table 2. Atomic percentages of elements obtained by XP spectra for the MMO@C electrocatalyst.
C (%)O (%)Ga (%)Cu (%)Ga/Cu
Atomic Ratio
60261226
Table 3. BET surface area (SBET), pore volume, and onset potential (Eo) at −10 mA cm−2 towards the HER at pH = 7.0 for electrocatalysts.
Table 3. BET surface area (SBET), pore volume, and onset potential (Eo) at −10 mA cm−2 towards the HER at pH = 7.0 for electrocatalysts.
ElectrocatalystSBET
(m2 g−1)
V Pore
(cm3 g−1)
EO (V) vs. RHE,
pH = 7.0
@C130.017−0.90
CuO@C220.047−0.38
Ga2O3@C410.066−0.33
MMO@C620.102+0.12
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

Barrientos, C.; Moris, S.; Arias, D.; Pecchi, G.; Ibarra, J.; Ramírez, G.; Gidi, L. New Paste Electrode Based on Copper and Gallium Mixed Metal Oxides-Decorated CNT for Highly Electrocatalyzed Hydrogen Evolution Reaction. Int. J. Mol. Sci. 2025, 26, 9057. https://doi.org/10.3390/ijms26189057

AMA Style

Barrientos C, Moris S, Arias D, Pecchi G, Ibarra J, Ramírez G, Gidi L. New Paste Electrode Based on Copper and Gallium Mixed Metal Oxides-Decorated CNT for Highly Electrocatalyzed Hydrogen Evolution Reaction. International Journal of Molecular Sciences. 2025; 26(18):9057. https://doi.org/10.3390/ijms26189057

Chicago/Turabian Style

Barrientos, Claudio, Silvana Moris, Dana Arias, Gina Pecchi, José Ibarra, Galo Ramírez, and Leyla Gidi. 2025. "New Paste Electrode Based on Copper and Gallium Mixed Metal Oxides-Decorated CNT for Highly Electrocatalyzed Hydrogen Evolution Reaction" International Journal of Molecular Sciences 26, no. 18: 9057. https://doi.org/10.3390/ijms26189057

APA Style

Barrientos, C., Moris, S., Arias, D., Pecchi, G., Ibarra, J., Ramírez, G., & Gidi, L. (2025). New Paste Electrode Based on Copper and Gallium Mixed Metal Oxides-Decorated CNT for Highly Electrocatalyzed Hydrogen Evolution Reaction. International Journal of Molecular Sciences, 26(18), 9057. https://doi.org/10.3390/ijms26189057

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