Previous Article in Journal
Effects of Graphene Quantum Dots on Thermal Properties of Epoxy Using Molecular Dynamics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electroless Nickel Plating of Magnesium Particles for Hydrogen Storage

by
Sindy Bello
1,*,
Robinson Aguirre Ocampo
2,
Julián Arias Velandia
2,
Alejandro Zuleta Gil
3,
Esteban Correa
1,
Wilber Silva
4,
Julián Andrés Lenis Rodas
2,
Carlos Arrieta
5,
Francisco Bolívar
2,
Cesar Nieto
6 and
Félix Echeverria
2
1
Grupo de Investigación Materiales con Impacto—MAT&MPAC, Facultad de Ingenierías, Universidad de Medellín UdeM, Carrera 87 No 30–65, Medellín 050026, Colombia
2
Centro de Investigación, Innovación y Desarrollo de Materiales—CIDEMAT, Facultad de Ingeniería, Universidad de Antioquia UdeA, Calle 70 No 52–21, Medellín 050010, Colombia
3
Grupo de Investigación de Estudios en Diseño—GED, Facultad de Diseño Industrial, Universidad Pontificia Bolivariana, Sede Medellín, Circular 1 No 70–01, Medellín 050010, Colombia
4
Grupo de Investigación en Óptica y Espectroscopía, Escuela de Ingenierías, Universidad Pontificia Bolivariana, Sede Medellín, Circular 1 No 70–01, Medellín 050010, Colombia
5
Grupo de Investigación en Energía—GRINEN, Facultad de Ingenierías, Universidad de Medellín UdeM, Carrera 87 No 30–65, Medellín 050026, Colombia
6
Grupo de Energía y Termodinámica, Escuela de Ingenierías, Universidad Pontificia Bolivariana, Sede Medellín, Circular 1 No 70–01, Medellín 050010, Colombia
*
Author to whom correspondence should be addressed.
Appl. Nano 2025, 6(3), 16; https://doi.org/10.3390/applnano6030016
Submission received: 25 June 2025 / Revised: 8 August 2025 / Accepted: 18 August 2025 / Published: 20 August 2025

Abstract

Hydrogen is emerging as a key energy vector for the transition toward renewable and sustainable energy sources. However, its safe and efficient storage remains a significant technical challenge in terms of cost, safety, and performance. In this study, we aimed to address the kinetic limitations of Mg by synthesizing catalyzed Mg@Ni systems using commercially available micrometric magnesium particles (~26 µm), which were decorated via electroless nickel plating under both aqueous and anhydrous conditions. Morphological and compositional characterization was carried out using SEM, EDS, and XRD. The resulting materials were evaluated through Temperature-Programmed Desorption (TPD), DSC, and isothermal hydrogen absorption/desorption kinetics. Reversibility over multiple absorption–desorption cycles was also investigated. The synthesized Mg@NiB system shows a reduction of 37 °C in the hydrogen release activation temperature at atmospheric pressure and a decrease of 167.3 °C under high vacuum conditions (4.5 × 10−7 MPa), in addition to a reversible hydrogen absorption/desorption capacity of 3.5 ± 0.09 wt.%. Additionally, the apparent activation energy for hydrogen desorption was lower (161.7 ± 21.7 kJ/mol) than that of hydrogenated commercial pure magnesium and was comparable to that of milling MgH2 systems. This research is expected to contribute to the development of efficient and low-cost processing routes for large-scale Mg catalysis.

1. Introduction

Hydrogen is a key substance in the global energy market. Although it is not a primary source of energy, it can store and produce clean energy due to its high gravimetric energy density and its capacity to generate energy without direct CO2 emissions [1,2,3]. However, the efficient utilization of hydrogen energy relies on overcoming significant storage challenges, including achieving high volumetric and gravimetric density, reversibility, safety, and operation at moderate temperatures [1,4]. Among the most promising approaches to addressing these challenges is solid-state hydrogen storage, which involves the use of specially designed materials capable of absorbing and releasing hydrogen safely and efficiently. Metal hydrides are a class of hydrogen storage materials that allow for significantly higher-density hydrogen storage compared to their gaseous or liquid forms. They also function under relatively mild temperature and pressure conditions, enhancing their economic and technical viability [4,5,6]. Magnesium-based metal hydride (MgH2) is particularly well suited for this application, offering a high gravimetric capacity (7.6 wt.%) and a high volumetric capacity (110 g/L) capable of achieving under specific conditions such as high temperature, high pressure, and particle size usually in the nanometric scale [7,8,9].
Moreover, Mg is low-cost and naturally abundant [4]. Nonetheless, its industrial application as a solid-state hydrogen storage medium is limited by several intrinsic drawbacks, including high thermal stability, slow absorption/desorption kinetics, high desorption enthalpy around 75 kJ/mol, and susceptibility to environmental contamination [10], which results in significant energy consumption for driving their hydrogen storage reaction [11,12,13,14,15]. One strategy to overcome these limitations is the use of catalysts, typically intermetallic compounds based on Ti, Nb, Ni, metal carbides, fluorides, chlorides, etc., that enhance the absorption/desorption kinetics and reduce the hydrogen release enthalpy [2,3,5,16,17].
Nickel-based compounds are among the most efficient catalysts due to their high catalytic reactivity and low cost [2,3,5,18,19]. Ni-based catalysts enhance the hydrogen sorption properties of Mg by acting as active sites for H2 dissociation. This enhancement is attributed to Ni’s high electron density and significant d-orbital occupancy, which facilitates electron transfer to the antibonding σ* orbital of the hydrogen molecule [20,21,22]. These characteristics enable the dissociation of H–H bonds, promote the diffusion of hydrogen molecules into the magnesium lattice, and effectively destabilize Mg–H and H–H bonds [3,5,18,23,24,25]—the improvement in the hydrogen sorption properties of Mg results in lower desorption temperatures and enhanced cyclic stability.
Mechanical milling is a widely employed technique to enhance hydrogen sorption properties, notably achieving up to a 50% reduction in activation energy [2,5]. This improvement is likely attributable to a decrease in crystal size or the generation of many defects. It is also a widely adopted physical processing route for producing Mg-based nano-composites [2,6,24,26]. However, a major inconvenience of this physical method is the limited control over morphology and composition of the resulting materials, as particle sizes tend to be heterogeneous, and the introduction of impurities is often unavoidable [5]. The production of Mg-based catalyzed nano-composites through mechanical milling also entails high energy consumption (frequently requiring prolonged operation times and high-impact energy for particle size reduction), leading to elevated costs for large-scale production [27].
As a result, there has been increasing interest in chemical processing routes designed to enhance the hydrogen sorption properties of Mg without relying on particle size reduction [2,3]. These chemical routes involve controlled reactions that produce nanoscale particles or thin coatings on the surface of Mg (decoration), enabling a uniform and reproducible integration of catalytic structures. Accordingly, the electroless nickel plating process offers a promising alternative to achieving such a goal. Electroless plating is a chemical processing route that involves the deposition of Ni through the reduction of metal ions in solution, without the need for an external electric current or specialized equipment, thereby simplifying the process [28]. This chemical approach offers improved uniformity in catalyst dispersion and enhanced control over the composition, chemical properties, and morphology of the synthesized particles [3,5,29].
Electroless plating of Ni on Mg involves three fundamental stages: cleaning, etching, and deposition. Initially, thorough cleaning is performed to remove organic contaminants and loose particles, using mild degreasers due to the sensitivity of Mg. Subsequently, a controlled chemical etching is applied to remove the surface oxide layer and increase the substrate’s reactivity, followed by activation using catalytic solutions, which facilitate Ni nucleation. Finally, the sample is immersed in an electroless solution containing Ni salts, a reducing agent, and stabilizing additives [30]. The electroless deposition of Ni on Mg begins with the formation of nucleation sites, or small Ni nuclei. The nuclei are formed by the reduction of Ni ions dissolved in the solution and adhere to the Mg surface, facilitated by the reducing agent. As the reaction progresses, the Ni nuclei grow as metallic nodules and later clump together to form a continuous layer [28,30,31,32]. In this way, reaction time is a determining variable in modulating the size of the Ni deposits on the Mg surface.
In this study, electroless Ni deposition is performed directly onto the micrometric Mg particles without cleaning or chemical etching, as shown in other studies [33,34,35,36,37]. Performing electroless Ni deposition improves productivity and reduces operational costs. Furthermore, compared to other chemical processing methods for enhancing the hydrogen sorption properties of Mg, which have low productivity [5], electroless Ni plating exhibits high productivity, making it a promising technique for the large-scale production of Ni-decorated micrometric Mg particles. Therefore, the main goal of the present study is to synthesize Ni-decorated Mg particles via electroless coating to improve their hydrogen sorption performance by facilitating H-H bond dissociation and reducing the hydrogen release temperature of the Mg. This research is expected to support the development of efficient and cost-effective processing routes for large-scale Mg catalysis.

2. Materials and Methods

Magnesium particles (approximately 26 µm ± 5 µm), sourced from Tangshan Weihao Magnesium Powder Co., Ltd. China, Qian’an, were used as a substrate for nickel deposition. This was achieved via an aqueous Ni–P electroless plating bath, following a formulation previously developed by the research group [31]. The aqueous Ni–P electroless plating bath was prepared under the following conditions: NH4HF2 at 2.4 g/L, bath loading of 0.15 mg/mL, temperature of 35 °C, pH of 10.5, and a plating time of 30 s. Ni-decorated Mg particles were synthesized using aqueous electroless plating. This process occurred within a gas extraction chamber, under ambient atmosphere (without inert gas), where the plating bath was prepared, and then the plating was conducted at 500 rpm with the Mg particles suspended in THF. After plating, the particles were filtered under high vacuum, washed with isopropyl alcohol, and dried at 60 °C for 60 min.
In addition, an anhydrous Ni–B electroless plating formulation was designed and proposed based on the group’s prior experience [30]. For this, ethylene glycol (≥99.9%, PANREAC APPLICHEM) was used as the solvent, anhydrous NiCl2 (≥99.9%, SIGMA ALDRICH) as the nickel source, and NaBH4 (≥99.9%, MERCK) as the reducing agent. The specific composition of the plating bath and the corresponding mixing conditions are detailed in Table 1.
The effectiveness of the anhydrous Ni–B electroless plating bath was previously assessed through its application on a carbon steel surface. Evidence of bubbling and surface discoloration indicated that a redox reaction had occurred. With these observations, the mixing protocol and key plating conditions were established, providing the basis for achieving successful Ni decoration on the surface of Mg particles. To prepare the anhydrous Ni–B electroless plating bath, the components listed in Table 1 were dissolved in ethylene glycol. A volume corresponding to 67.3% of the total bath volume was used for C4H6O6 and NiCl2. This dissolution occurred inside a sealed vessel at 60 °C and 600 rpm for 15 min (Step 1). Concurrently with the initial preparation, NaOH was dissolved under the same conditions to adjust the pH to 10.5 [38]. This NaOH solution comprised 13.5% of the total bath volume (Step 2). Following this, the reducing agent NaBH4 was introduced, contributing 9.6% to the total bath volume (Step 3). All reagents were introduced via syringes to maintain the system’s hermetic conditions. Each component needed to be added as quickly as possible, allowing for a 5-min mixing consolidation time between Steps 1, 2, and 3. Any delay in addition, especially of the NaOH or NaBH4 suspensions, could lead to the premature reduction of bath salts. This would result in a mint-green, gel-like compound, invalidating the process and requiring a complete restart.
Once the anhydrous electroless bath was prepared, a continuous nitrogen flow was introduced to create bubbling, thereby promoting the initiation of the redox reaction [32]. The Mg particles were then subjected to a 5-min plating reaction (Step 6). These particles were also introduced via syringe after being suspended in ethylene glycol within a sealed vessel inside the glovebox, corresponding to the remaining 9.6% of the total bath volume (Step 5). The resulting product was centrifuged at 6000 rpm for 45 min. For this, the sealed vessel containing the reaction product was transferred back into the glovebox (Vigor Tech, Houston, TX, USA O2 and H2O < 1 ppm with argon atmosphere) and poured into Falcon tubes (Step 7). Finally, the synthesized particles were washed three times with acetone by centrifugation at 6000 rpm for 5 min (15 min in total) and then dried under vacuum using hot air (Step 8). Throughout this process, environmental exposure was carefully controlled by using a continuous nitrogen flow during the reaction since the electroless plating process had to be performed outside the glovebox due to the need for heating and the release of gases during the process, which could compromise the inert and controlled atmosphere within the glovebox. To maintain an inert nitrogen atmosphere while operating outside the glovebox, silicone stoppers were used to keep the nitrogen flow contained within the reaction vessel. Figure 1 outlines the step-by-step procedure for preparing the anhydrous Ni–B electroless bath formulated by the research group.
After the deposition process, the morphology and microstructure of the samples were analyzed using a thermionic Scanning Electron Microscope JEOL-JSM 6490LV., JPN (SEM) and a ThermoFisher Scientific Apreo 2 Field Emission Scanning Electron Microscope, USA (FESEM), both microscopes outfitted with energy-dispersive X-ray (EDX) microprobes. Phase identification was performed using an X-ray diffractometer (XRD) XPert PANalytical Empyrean Series II—Alpha1., NL., equipped with Cu Kα radiation, scanning over a 2θ range of 5 to 90°. A domed polycarbonate sample holder with a 16 mm Ø zero-background diffraction plate and 2 mm thickness was used for the safe handling and transport of air-sensitive samples. According to the HighScore Plus database, the identified phases included Mg (CIDC 96-151-2520), MgO (CIDC 96-101-1118), MgH2 (CIDC 01-074-0934), Ni4B3 (CIDC 00-012-0417), Mg2NiH4 (CIDC 00-040-1204), and CB4 (CIDC 00-006-0555).
Hydrogen storage properties were evaluated using a Sieverts-type pressure–composition–temperature (PCT) apparatus designed by the research group. Using Temperature-Programmed Desorption (TPD), isothermal hydrogen absorption/desorption kinetics at 350 °C, and multiple absorption/desorption cycles. Samples were activated (i.e., subjected to expansion and contraction cycles) using a custom protocol of activation involving six hydrogen absorption/desorption cycles of 30 min each, followed by a prolonged absorption cycle of 720 min and a desorption cycle of 540 min at 350 °C. Hydrogen absorption tests were carried out under an initial hydrogen pressure of 2.1 MPa, while desorption tests were performed under an initial vacuum pressure of 4.5 × 10−7 MPa and a final vacuum pressure of about 0.17 MPa. Hydrogen absorption/desorption values were calculated using the NIST Reference Fluid Thermodynamic and Transport Properties Database (REFPROP)® database [39]. Activation energy (Ea) was estimated from kinetic curves at 325, 350, and 375 °C using the Arrhenius equation:
ln k = E a R T + ln k 0
where R is the constant molar gas, T is the temperature in K, and k0 is a constant. The reaction rate constant (k) is calculated through the Johnson–Mehl–Avrami–Kolmogorov (JMAK) equation:
ln ln 1 = n ln k + n ln t
After desorbing at 325 °C, the sample was heated to 350 °C to release all the hydrogen that it failed to release at 325 °C. n is the Arrhenius exponent, and α is the reaction rate fraction at time t [18,40,41,42,43,44,45,46]. Although the classical linearized form of the JMAK equation is ln[−ln(1 − α)] = ln(k) + n·ln(t), in this study we adopted the regrouped expression ln[−ln(1 − α)] = n·ln(k) + n·ln(t). This formulation is widely reported in hydrogen storage studies involving MgH2-based systems [47,48,49,50,51], allowing for methodological consistency and direct comparison with related works. Furthermore, it provides a more comprehensive representation of the transformation kinetics under non-ideal experimental conditions, where the rate constant k may encompass geometric, thermodynamic, and kinetic contributions. As noted in the studies of Hristov [52], Jin et al. [53], and Mooij and Dam [54], this regrouped form appropriately reflects the interplay between nucleation, growth dimensionality, and interfacial or diffusion-limited mechanisms.
In addition, differential scanning calorimetry (DSC) analyses were performed using a Thermal Analysis—TA Instruments, TGA/DTA 5500, DSC550., USA., system under a nitrogen flow of 50 mL/min with a heating rate of 2, 3, and 5 °C/min, from 100 to 400 °C, to support the findings regarding the thermal behavior and desorption characteristics of the hydrogenated samples.

3. Results

3.1. Aqueous Ni–P Electroless Bath

An initial evaluation was performed on uncoated magnesium particles to assess the synthesis effectiveness. These results were then compared with those obtained from the particles following the electroless Ni–P process. Figure 2a,b shows the initial morphology of the magnesium particles, and following the electroless plating process (c, d, and e). Figure 2a displays the spherical morphology (~24 µm in diameter) of the magnesium particles before the electroless process. On their surface, small, agglomerated particles, composed of Mg and O, are visible (Figure 2b). Following the electroless procedure on these magnesium particles, the SEM images (Figure 2c,d) reveal the presence of nickel aggregates (brighter deposits) on the surface of the micrometer-sized magnesium particles, with sizes ranging from 225 to 853 nm, while preserving the spherical geometry of the Mg particles. EDS elemental mapping (Figure 2e) confirmed that the nodules were composed of nickel (orange) and phosphorus (pink), confirming the successful deposition of Ni–P particles on the Mg surface. The average elemental composition of the nickel-decorated magnesium particles was as follows: 3.2 wt.% ± 0.084 fluorine; 1.5 wt.% ± 0.029 sodium; 72.5 wt.% ± 0.025 magnesium; 2.0 wt.% ± 0.020 phosphorus; 6.6 wt.% ± 0.054 nickel. Although the plating reaction was conducted outside the glovebox, the oxygen content remained relatively low, only 14.1% ± 0.086 wt.%), suggesting that the particles remained partially protected during the electroless plating process.

3.2. Anhydrous Ni–B Electroless Bath

Figure 3 presents the SEM images of magnesium particles following the electroless Ni–B process, using the anhydrous formulation detailed in Table 1. Figure 3a shows that the formulated anhydrous electrolytic bath does not deteriorate the initial morphology of the Mg particles, indicating that the particles are stable under the given operating conditions. High-contrast areas corresponding to the Ni deposits (red arrows) are evident in Figure 3b. This anhydrous electroless bath allows the decoration of the micrometric Mg particles with Ni nodules that are smaller in size than those obtained using the previously mentioned aqueous nickel-plating bath; in this case, it is impossible to determine the size with this technique of characterization. Therefore, through elemental mapping, it was identified that the Ni–B nodules are homogeneously distributed over the surface of the magnesium particles (Figure 3c,d), along with carbon deposits originating from bath residues and the acetone used during washing. The average elemental composition of the nickel-decorated magnesium particles was as follows: 85.7 wt.% magnesium ± 0.30; 2.8 wt.% ± 0.04 nickel; 0.16 wt.% ± 0.01 boron; 5.3 wt.% ± 0.07 carbon; 6.1 wt.% ± 0.1 oxygen.

3.3. Thermal and Hydrogen Storage Properties

To evaluate the performance of the nickel-decorated magnesium particles, their thermal behavior and hydrogen absorption/desorption properties were analyzed. As a first step, the modified particles were subjected to an activation protocol consisting of six hydrogen absorption/desorption cycles of 30 min each, followed by an absorption cycle of 720 min and a desorption cycle of 540 min at 350 °C (see Figure 4). Figure 5 shows the comparison of hydrogen absorption/desorption isothermal kinetic curves at 350 °C for Mg particles decorated with aqueous Ni–P electroless plating (Mg@NiP), Mg particles decorated with anhydrous Ni–B electroless plating (Mg@NiB), and Mg particles without electroless plating (MgH2). In 90 min, Mg absorbs 3.2 wt.% of hydrogen, Mg@NiB 4.3 wt.% of hydrogen, and Mg@NiP 2.2 wt.% of hydrogen. In 60 min, the samples desorb 1.5, 4.41, and 0.03 wt.%, respectively. The aqueous Ni–P electroless decoration process worsens the hydrogen absorption kinetics of Mg particles by 31% and the desorption kinetics by 98%. On the contrary, the anhydrous Ni–B electroless decoration process improves the hydrogen sorption of Mg particles by 34% for absorption and by 173% for desorption.
Figure 6 shows the isothermal absorption/desorption kinetic curves at 350 °C for MgH2 and the Mg@NiB sample. For the latter, three consecutive activation protocol cycles were evaluated: cycle 1 (black curve), cycle 2 (orange curve), and cycle 3 (aqua green curve). In each cycle, the Mg@NiB sample absorbs, in 20 min, 3.7 wt.%, 3.3 wt.%, and 3.1 wt.% of hydrogen, respectively, compared to the MgH2 sample, which only absorbs 1.6 wt.% of hydrogen. The maximum absorption time was 90 min, during which the highest hydrogen absorption for the Mg@NiB sample was 4.3 wt.% in cycle one and the lowest was 3.6 wt.% in cycle three; for the MgH2 sample, the maximum was 3.2 wt.%. Regarding the desorption of the absorbed hydrogen, the Mg@NiB sample is capable of completely releasing the absorbed hydrogen in all three evaluated cycles within 60 min, unlike the MgH2 sample, which only releases 1.7 wt.% of hydrogen.
Figure 7 presents the results of TPD and DSC analyses. Figure 7a shows the temperature-programmed desorption (TPD) test at a heating rate of 5 °C/min for the Mg@NiB sample (sample of particles decorated via the anhydrous electroless Ni–B process). The onset of hydrogen release occurs at 330 °C, with two endothermic peaks suggesting the presence of multiple desorption phases, associated with the decomposition of Mg2NiH4 (240.6 °C) and MgH2 (330 °C),. Figure 7b shows the DSC analysis at different heating rates: 2 °C/min, 3 °C/min, and 5 °C/min. All curves exhibit two prominent endothermic peaks, confirming the thermal events observed in the TPD test. The desorption peak temperatures shift to higher values with increasing heating rates, occurring at 365.4 °C, 389.7 °C, and 391.3 °C, respectively.
Figure 8a shows the initial state of the Mg@NiB particles exhibiting a smooth surface and spherical morphology with a diameter of 32 µm. As the cycling progresses, visible surface cracking, roughening, and partial exfoliation become evident (Figure 8b–g), suggesting cumulative mechanical stress due to volumetric changes during hydride formation and decomposition, accompanied by a contraction of 50%. Also, it was observed that despite the deformation of the particle during the activation process, the nickel particles remained adhered to the surface of the Mg.
Figure 9 shows the XRD of commercial Mg particles (black spectrum), Mg@NiB sample (red spectrum), and the Mg@NiB activated sample that has been activated according to the protocol of activation (blue spectrum). Initially, only signals corresponding to magnesium are observed. After the particles are decorated, characteristic peaks corresponding to boron carbide and nickel–boron nodules appear. Following activation in the Sieverts apparatus, additional peaks are detected corresponding to magnesium oxide, magnesium hydride, and magnesium–nickel hydride.
The aim is to identify the structural transformations of Mg phases in their natural state (hexagonal compact crystal system) and their hydrogenated state (tetragonal crystal system). An XRD analysis was conducted to identify Ni catalysis of hydrogen sorption through the presence of Mg2NiH4 and to corroborate that the Mg@NiB sample absorbs hydrogen and can desorb. The results are coherent with the onset values observed in Figure 7. Figure 10 Over the following samples: MgH2 (gray spectrum), Mg@NiB hydrogenated (red spectrum), and Mg@NiB dehydrogenated (fuchsia spectrum), they are identically decorated particles after dehydrogenation. The compounds identified in Figure 9 remain present, and this comparison highlights the structural changes that the sample undergoes during the hydrogen absorption and desorption process.
Finally, Figure 11a presents the JMAK and Arrhenius plots for the Mg@NiB sample. Ea provides a measure of the energy barrier for the hydrogen desorption reaction; the lower Ea value improves desorption kinetics. The energy activation (Ea) for the sample is 94.56 ± 4.43 kJ/mol.
Figure 11b shows the hydrogen absorption/desorption cycles at 350 °C for the Mg@NiB sample. After 10 additional cycles following the initial three hydrogen absorption/desorption cycles (see Figure 6). The reversible hydrogen absorption/desorption capacity remains at 3.5% ± 0.09% by weight, with a capacity retention of 100%.

4. Discussion

The present results demonstrate that micrometric magnesium particles can be successfully decorated using aqueous Ni–P electroless baths, even under corrosive conditions and environmental exposure. The distribution of the deposited nickel on the surface of these particles is consistent with previous findings reported for Mg substrate surfaces [28,31]. However, for magnesium particles decorated via the aqueous Ni–P electroless bath, no improvement in hydrogen storage properties was observed. The decoration of Mg particles using an aqueous Ni–P electroless bath does not catalyze the Mg surface and significantly impairs hydrogen sorption kinetics at 350 °C (see Figure 5). Mg@NiP cannot release hydrogen, unlike Mg@NiB and the MgH2 sample. However, elements such as F, Na, P, and O have been individually reported as effective catalysts for magnesium [17,19,55,56,57]. The particles synthesized through aqueous electroless plating incorporate all of them simultaneously. This simultaneous incorporation leads to the blockage of active sites, the formation of physical barriers, or the development of passive phases, all of which hinder hydrogen diffusion into the magnesium matrix. Selim Kazaz [58], and M. Sherif El-Eskandarany [59] found a similar phenomenon in their studies, increasing the catalyst significantly impaired hydrogen sorption.
On the other hand, micrometric magnesium particles decorated with the anhydrous Ni–B electroless bath exhibited a significant improvement in their hydrogen sorption kinetics. SEM analysis revealed that the activation of the coated Ni–B Mg particles (based on cyclic volume expansion/contraction) induces structural destabilization of the crystalline structure of magnesium through cyclic plastic deformation, triggering a flaking effect on the particles [60]. Furthermore, XRD analysis in conjunction with EDS reveals the presence of CB4 and Ni–B phases on the surface of the nanometer-scale magnesium particles. These two phases enhance the conductivity and cyclic stability of magnesium, as reported in [61,62]. This, in turn, promotes hydrogen diffusion pathways and supports the cyclic stability of the Mg@NiB sample, even after 10 cycles of hydrogen absorption and desorption.
Moreover, the XRD and TPD results demonstrate that the Mg@NiB sample can release nearly all the stored hydrogen at 350 °C under a pressure of 4.5 × 10−7 MPa, despite the micrometric size. The residual presence of magnesium hydride and magnesium–nickel hydride is attributed to the micrometric particle size, which offers greater resistance to complete hydrogen release [60]. Several authors define the onset temperature from TPD in hydrogen desorption [63,64] as the point at which desorption begins, regardless of the amount of hydrogen released.
Catalyzing micrometric magnesium particles through the anhydrous Ni–B electroless coating enables a reduction in the onset temperature for release of hydrogen under a pressure of 4.5 × 10−7 MPa (TPD) from 441 °C to 330 °C (a decrease of 111 °C) [65]. At atmospheric pressure (DSC), the hydrogen release temperature is reduced by 37 °C, from 428.3 °C to 391.3 °C [19], and the apparent activation energy (Ea) for the desorption process, determined from the linear fit of the JMAK equation, was calculated as 161.7 ± 21.7 kJ/mol. This value is typically reported for commercial MgH2 to be about 160–245 kJ/mol [66,67,68]. Compared to commercial MgH2 without ball milling, which has values in some cases close to 245 kJ/mol [67], there is a reduction in energy activation with electroless decoration.
The catalytic mechanism of magnesium with surface-deposited nickel is illustrated in Figure 12. Nickel, distributed on the surface of the Mg particles, plays a crucial role in facilitating the dissociation of hydrogen molecules due to its favorable electronic properties [2,5,18,20,22,23]. This surface configuration allows H2 molecules to dissociate before interacting directly with the Mg matrix, thereby enhancing hydrogen diffusion and reducing the desorption temperature. This explains the behavior of the Mg@NiB sample.
The results obtained in the present work are notable for micrometric particles, achieving superior values in hydrogen sorption kinetics and storage capacity compared to other studies carried out using materials with particles in the nanometric scale for Mg and Ni catalysts [3,24,27,69,70]. In addition, the electroless decoration process is reproducible, economical, and versatile, compared to other methods, yielding good kinetic results for micrometric scale particles [71]. With only the anhydrous electroless decoration as a catalytic method, a reversible hydrogen absorption/desorption capacity of 3.5 wt.% ± 0.09 wt.% between 90 and 60 min at 350 °C represents an improvement of nearly 10% up to over 100% in hydrogen absorption and desorption kinetics, respectively, compared to untreated MgH2. This marks a significant advancement in improving hydrogen sorption kinetics for micrometric particles, which are typically not the focus of such studies, as size reduction techniques are generally preferred to enhance performance [3,4,10,16,17,19,24,25,27,70].

5. Conclusions

In this study, magnesium particles (~26 µm ± 5 µm) from Tangshan Weihao Magnesium Powder Co. were successfully decorated with nickel using an electroless bath formulated by the research group, resulting in a material with high hydrogen storage capacity and fast absorption/desorption kinetics compared to untreated MgH2. The magnesium particles decorated with the anhydrous Ni–B electroless bath (Mg@NiB) were catalyzed not only through the incorporation of Ni–B nodules but also by the presence of CB4 deposits. Additionally, the study demonstrated the critical role of volumetric expansion and contraction cycles in activating the sample, specifically, in increasing the available diffusion pathways for hydrogen molecules to reach the magnesium. The synthesized Mg@NiB sample exhibited favorable kinetic behavior, apparent activation energy, and cyclic stability, comparable to results reported in studies involving samples with micrometric particle sizes. These findings highlight the potential of our method as a more straightforward and potentially scalable alternative to ball milling, while still achieving desirable kinetic improvements and avoiding the limitations and costs associated with using ball-milled Mg nanomaterials for solid-state hydrogen storage. We hope this research contributes to the development of more energy-efficient techniques to enhance hydrogen sorption in magnesium hydride, with greater potential for industrial scalability.

Author Contributions

S.B.: writing—review and editing, writing—original draft, investigation, formal analysis; R.A.O.: visualization, data curation, conceptualization; J.A.V.: software, resources, formal analysis; J.A.L.R.: visualization, data curation, conceptualization; A.Z.G.: writing—review and editing, supervision, project administration; E.C.: writing—review and editing, supervision, project administration; C.A.: supervision, project administration; F.B.: supervision, project administration; W.S.: writing—review and editing, supervision; C.N.: supervision, project administration; F.E.: supervision, project administration, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the “Sistema General de Regalías—SGR” through the project BPIN 2022000100089, and the M.Sc. student Sindy Bello is grateful to the Universidad de Antioquia and the Sede de Investigación Universitaria (SIU), NIT 890980040-8., for their support through the CONVENIO BECA POSGRADO 20230046-11-2024.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Acknowledgments

The authors are pleased to acknowledge the financial assistance of the “Sistema General de Regalías—SGR” through the project BPIN 2022000100089, Centro de Investigación para el Desarrollo de la Innovación (CIDI) from Universidad Pontificia Bolivariana (Rad: 822C- 06/23–35), Centro de Investigación en Ingenierías (CEIN) from Universidad de Medellín, “Estrategia de Sostenibilidad de la Universidad de Antioquia” and the M.Sc. student Sindy Bello is grateful to the Universidad de Antioquia and the Sede de Investigación Universitaria (SIU) for their support through the CONVENIO BECA POSGRADO 20230046-11-2024.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work report.

References

  1. Shang, Y.; Pistidda, C.; Gizer, G.; Klassen, T.; Dornheim, M. Mg-Based Materials for Hydrogen Storage. J. Magnes. Alloys 2021, 9, 1837–1860. [Google Scholar] [CrossRef]
  2. Liu, Y.; Wang, S.; Li, Z.; Gao, M.; Liu, Y.; Sun, W.; Pan, H. Enhanced Hydrogen Storage Performance of MgH2 by the Catalysis of a Novel Intersected Y2O3/NiO Hybrid. Processes 2021, 9, 892. [Google Scholar] [CrossRef]
  3. Lu, C.; Zou, J.; Shi, X.; Zeng, X.; Ding, W. Synthesis and Hydrogen Storage Properties of Core–Shell Structured Binary Mg@Ti and Ternary Mg@Ti@Ni Composites. Int. J. Hydrogen Energy 2017, 42, 2239–2247. [Google Scholar] [CrossRef]
  4. Moradi, R.; Groth, K.M. Hydrogen Storage and Delivery: Review of the State of the Art Technologies and Risk and Reliability Analysis. Int. J. Hydrogen Energy 2019, 44, 12254–12269. [Google Scholar] [CrossRef]
  5. Zhang, J.; Zhu, Y.; Yao, L.; Xu, C.; Liu, Y.; Li, L. State of the Art Multi-Strategy Improvement of Mg-Based Hydrides for Hydrogen Storage. J. Alloys Compd. 2019, 782, 796–823. [Google Scholar] [CrossRef]
  6. Dai, M.; Bolarin, J.A.; Lei, G.; Li, Z.; He, T.; Cao, H.; Chen, P. Potassium Hydride Reduced Black TiO2−x for Boosting the Hydrogenation of Magnesium at Room Temperature. J. Alloys Compd. 2022, 897, 162750. [Google Scholar] [CrossRef]
  7. Shtuckmeyster, D.; Bereznitsky, M.; Shelly, L.; Maman, N.; Hanuka, E.; Moskovich, N.; Shneck, R.Z.; Jacob, I. First Hydrogenation Kinetics and Properties of Mechanically Treated Pure Mg. ACS Appl. Energy Mater. 2025, 8, 2145–2157. [Google Scholar] [CrossRef]
  8. Yartys, V.A.; Lototskyy, M.V.; Akiba, E.; Albert, R.; Antonov, V.E.; Ares, J.R.; Baricco, M.; Bourgeois, N.; Buckley, C.E.; Bellosta von Colbe, J.M.; et al. Magnesium Based Materials for Hydrogen Based Energy Storage: Past, Present and Future. Int. J. Hydrogen Energy 2019, 44, 7809–7859. [Google Scholar] [CrossRef]
  9. Marzouki, M.; Ghodbane, O.; Abdellaoui, M. Study of Hydrogen Absorption Kinetics of Mg2Ni-Based Powders Produced by High-Injected Shock Power Mechanical Alloying and Subsequent Annealing. Mater. Renew. Sustain. Energy 2013, 2, 9. [Google Scholar] [CrossRef]
  10. Crivello, J.C.; Dam, B.; Denys, R.V.; Dornheim, M.; Grant, D.M.; Huot, J.; Jensen, T.R.; de Jongh, P.; Latroche, M.; Milanese, C.; et al. Review of Magnesium Hydride-Based Materials: Development and Optimisation. Appl. Phys. A Mater. Sci. Process 2016, 122, 97. [Google Scholar] [CrossRef]
  11. Zhang, X.; Li, C.; Ye, J.; Hu, X.; Chen, W.; Fang, F.; Sun, D.; Liu, Y.; Yu, X.; Xia, G. Light-Enabled Reversible Hydrogen Storage of Borohydrides Activated by Photogenerated Vacancies. J. Am. Chem. Soc. 2025, 147, 2786–2796. [Google Scholar] [CrossRef]
  12. Hu, X.; Chen, X.; Zhang, X.; Meng, Y.; Xia, G.; Yu, X.; Sun, D.; Fang, F. In Situ Construction of Interface with Photothermal and Mutual Catalytic Effect for Efficient Solar-Driven Reversible Hydrogen Storage of MgH2. Adv. Sci. 2024, 11, 2400274. [Google Scholar] [CrossRef]
  13. Zhang, X.; Sun, Y.; Ju, S.; Ye, J.; Hu, X.; Chen, W.; Yao, L.; Xia, G.; Fang, F.; Sun, D.; et al. Solar-Driven Reversible Hydrogen Storage. Adv. Mater. 2023, 35, 2206946. [Google Scholar] [CrossRef]
  14. Zhang, X.; Ju, S.; Li, C.; Hao, J.; Sun, Y.; Hu, X.; Chen, W.; Chen, J.; He, L.; Xia, G.; et al. Atomic Reconstruction for Realizing Stable Solar-Driven Reversible Hydrogen Storage of Magnesium Hydride. Nat. Commun. 2024, 15, 2815. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, Y.; Chen, X.; Zhang, H.; Xia, G.; Sun, D.; Yu, X. Heterostructures Built in Metal Hydrides for Advanced Hydrogen Storage Reversibility. Adv. Mater. 2020, 32, 2202647. [Google Scholar] [CrossRef] [PubMed]
  16. Zheng, Z.; Peng, C.; Zhang, Q. Hydrogen Desorption Kinetics of Magnesium Hydride Co-Doped with Titanium Trifluoride and Nickelocene. Int. J. Hydrogen Energy 2024, 50, 1574–1580. [Google Scholar] [CrossRef]
  17. Czujko, T.; Varin, R.A.; Wronski, Z.; Zaranski, Z.; Durejko, T. Synthesis and Hydrogen Desorption Properties of Nanocomposite Magnesium Hydride with Sodium Borohydride (MgH2 + NaBH4). J. Alloys Compd. 2007, 427, 291–299. [Google Scholar] [CrossRef]
  18. Liu, Y.; Zou, J.; Zeng, X.; Wu, X.; Li, D.; Ding, W. Hydrogen Storage Properties of a Mg-Ni Nanocomposite Coprecipitated from Solution. J. Phys. Chem. C 2014, 118, 18401–18411. [Google Scholar] [CrossRef]
  19. Varin, R.A.; Czujko, T.; Wasmund, E.B.; Wronski, Z.S. Hydrogen Desorption Properties of MgH2 Nanocomposites with Nano-Oxides and Inco Micrometric- and Nanometric-Ni. J. Alloys Compd. 2007, 446–447, 63–66. [Google Scholar] [CrossRef]
  20. Houari, A.; Matar, S.F.; Eyert, V. Electronic Structure and Magnetic Ordering of NiN and Ni2N from First Principles. Electron. Struct. 2019, 1, 015002. [Google Scholar] [CrossRef]
  21. Suraj, K.S.; Assadi, M.H.N. Detrimental 2p-3d Hybridisation in Ni Nanosheets Supported on Strontium Dioxide for Catalytic H2 Production, Necessitating Thickness Optimisation. Energy Transit. Towar. Carbon Neutrality 2024, 44, 2–5. [Google Scholar] [CrossRef]
  22. Huo, L.; Jin, C.; Jiang, K.; Bao, Q.; Hu, Z.; Chu, J. Applications of Nickel-Based Electrocatalysts for Hydrogen Evolution Reaction. Adv. Energy Sustain. Res. 2022, 3, 2100189. [Google Scholar] [CrossRef]
  23. El-Eskandarany, M.S.; Al-Nasrallah, E.; Banyan, M.; Al-Ajmi, F. Bulk Nanocomposite MgH2/10 Wt% (8 Nb2O5/2 Ni) Solid-Hydrogen Storage System for Fuel Cell Applications. Int. J. Hydrogen Energy 2018, 43, 23382–23396. [Google Scholar] [CrossRef]
  24. Jeon, K.J.; Theodore, A.; Wu, C.Y.; Cai, M. Hydrogen Absorption/Desorption Kinetics of Magnesium Nano-Nickel Composites Synthesized by Dry Particle Coating Technique. Int. J. Hydrogen Energy 2007, 32, 1860–1868. [Google Scholar] [CrossRef]
  25. Song, M.; Zhang, L.; Wu, F.; Zhang, H.; Zhao, H.; Chen, L.; Li, H. Recent Advances of Magnesium Hydride as an Energy Storage Material. J. Mater. Sci. Technol. 2023, 149, 99–111. [Google Scholar] [CrossRef]
  26. Tarasov, B.P.; Arbuzov, A.A.; Mozhzhuhin, S.A.; Volodin, A.A.; Fursikov, P.V.; Lototskyy, M.V.; Yartys, V.A. Hydrogen Storage Behavior of Magnesium Catalyzed by Nickel-Graphene Nanocomposites. Int. J. Hydrogen Energy 2019, 44, 29212–29223. [Google Scholar] [CrossRef]
  27. El-Eskandarany, M.S.; Al-Ajmi, F.; Banyan, M. Mechanically-Induced Catalyzation of MgH2 Powders with Zr2Ni-Ball Milling Media. Catalysts 2019, 9, 382. [Google Scholar] [CrossRef]
  28. Loto, C.A. Electroless Nickel Plating–A Review. Silicon 2016, 8, 177–186. [Google Scholar] [CrossRef]
  29. Sun, Y.; Shen, C.; Lai, Q.; Liu, W.; Wang, D.W.; Aguey-Zinsou, K.F. Tailoring Magnesium Based Materials for Hydrogen Storage through Synthesis: Current State of the Art. Energy Storage Mater. 2018, 10, 168–198. [Google Scholar] [CrossRef]
  30. Correa, E.; Zuleta, A.A.; Guerra, L.; Castaño, J.G.; Echeverría, F.; Baron-Wiecheć, A.; Skeldon, P.; Thompson, G.E. Formation of Electroless Ni-B on Bifluoride-Activated Magnesium and AZ91D Alloy. J. Electrochem. Soc. 2013, 160, D327–D336. [Google Scholar] [CrossRef]
  31. Carrillo, D.F.; Bermudez, A.; Gómez, M.A.; Zuleta, A.A.; Castaño, J.G.; Mischler, S. Fretting-Corrosion Behavior of Electroless Ni-P/Ni-P-TiO2 Coatings Obtained on AZ91D Magnesium Alloy by a Chromium-Free Process. Surf. Interfaces 2020, 21, 100733. [Google Scholar] [CrossRef]
  32. Olsen, J.D.; Romankiw, L.T. Electroless Ni-Fe Deposition Process. U.S. Patent US3382079A, 7 May 1968. Available online: https://patents.google.com/patent/US3382079A/en1964 (accessed on 20 May 2025).
  33. Cheng, A.Y.; Huang, Y.C.; You, J.L.; Liu, Y.M.; Hou, K.H.; Ger, M. Der Evaluation of Electroless Plating on Magnesium Alloy Micro-Arc Oxidation Layer with Different Electroless Ni-P Plating Baths. J. Alloys Compd. 2025, 1011, 178398. [Google Scholar] [CrossRef]
  34. Lee, H.B.; Sheu, H.H.; Jian, J.S.; Chang, S.Y.; Yen, C.H.; Lin, H.E. Supercritical-CO2 Electroless Nickel Plating Enhanced Anti-Corrosion Properties of Micro-Arc Oxidized AZ31 Magnesium Alloy. Mater. Today Commun. 2022, 33, 104475. [Google Scholar] [CrossRef]
  35. Zhang, J.; Xie, Z.H.; Chen, H.; Hu, C.; Li, L.; Hu, B.; Song, Z.; Yan, D.; Yu, G. Electroless Deposition and Characterization of a Double-Layered Ni-B/Ni-P Coating on AZ91D Mg Alloy from Eco-Friendly Fluoride-Free Baths. Surf. Coat. Technol. 2018, 342, 178–189. [Google Scholar] [CrossRef]
  36. Shang, W.; Zhan, X.; Wen, Y.; Li, Y.; Zhang, Z.; Wu, F.; Wang, C. Deposition Mechanism of Electroless Nickel Plating of Composite Coatings on Magnesium Alloy. Chem. Eng. Sci. 2019, 207, 1299–1308. [Google Scholar] [CrossRef]
  37. Thakur, A.; Gharde, S.; Kandasubramanian, B. Electroless Nickel Fabrication on Surface Modified Magnesium Substrates. Def. Technol. 2019, 15, 636–644. [Google Scholar] [CrossRef]
  38. Makar, G.L.; Kruger, J. Corrosion of Magnesium. Int. Mater. Rev. 1993, 38, 138–153. [Google Scholar] [CrossRef]
  39. Huber, M.L.; Lemmon, E.W.; Bell, I.H.; McLinden, M.O. The NIST REFPROP Database for Highly Accurate Properties of Industrially Important Fluids. Ind. Eng. Chem. Res. 2022, 61, 15449–15472. [Google Scholar] [CrossRef]
  40. Zhu, X.; Yang, M.; Mu, D.; Gao, Y.; He, L.; Ma, L. Synergistic Combination of Ni Nanoparticles and Ti3C2 MXene Nanosheets with MgH2 Particles for Hydrogen Storage. ACS Appl. Nano Mater. 2023, 6, 21521–21531. [Google Scholar] [CrossRef]
  41. Yang, X.; Hou, Q.; Yu, L.; Zhang, J. Improvement of the Hydrogen Storage Characteristics of MgH2 with a Flake Ni Nano-Catalyst Composite. Dalton Trans. 2021, 50, 1797–1807. [Google Scholar] [CrossRef] [PubMed]
  42. Zhu, Y.; Qin, Z.; Shi, X.; Ding, X.; Dao, V.D.; Li, Y. Size-Dependent Activity Modulation of Supported Ni Nanocatalysts for Efficient Solid-State Hydrogen Storage in Magnesium. Chem. Eng. J. 2024, 498, 155285. [Google Scholar] [CrossRef]
  43. Yuan, Q.; Peng, C.; Yang, C.; Li, Y.; Zhang, Q.; Lv, Y.; Liu, G.; Liu, D. Facilitated Hydrogen Storage Properties of MgH2 by Ni Nanoparticles Anchored on Mo2C@C Nanosheets. Int. J. Hydrogen Energy 2024, 85, 12–19. [Google Scholar] [CrossRef]
  44. Yin, C.; Qiu, S.; Xia, Y.; Xu, F.; Sun, L.; Chu, H. Ni-Based Catalyst Assisted by MnO to Boost the Hydrogen Storage Performance of Magnesium Hydride. Int. J. Hydrogen Energy 2024, 55, 512–520. [Google Scholar] [CrossRef]
  45. Khan, D.; Ong, W.J. Tailoring Hydrogen Storage Materials Kinetics and Thermodynamics Through Nanostructuring, and Nanoconfinement With In-Situ Catalysis. Interdiscip. Mater. 2025, 4, 249–283. [Google Scholar] [CrossRef]
  46. Suárez-Alcántara, K.; Flores-Jacobo, N.I.; del Pilar Osorio-García, M.; Cabañas-Moreno, J.G. Fast Hydrogen Sorption Kinetics in Mg-VCl3 Produced by Cryogenic Ball-Milling. Materials 2023, 16, 2526. [Google Scholar] [CrossRef]
  47. Zhou, C.; Hu, C.; Li, Y.; Zhang, Q. Crystallite Growth Characteristics of Mg during Hydrogen Desorption of MgH2. Prog. Nat. Sci. Mater. Int. 2020, 30, 246–250. [Google Scholar] [CrossRef]
  48. Khan, D.; Zou, J.; Muhammad, S.; Khan, N.A.; Saud, S.; Panda, S. The Adaptable Effect of Ru on Hydrogen Sorption Characteristics of the MgH2 System. Mater. Chem. Phys. 2023, 301, 127583. [Google Scholar] [CrossRef]
  49. Nyahuma, F.M.; Zhang, L.; Song, M.; Lu, X.; Xiao, B.; Zheng, J.; Wu, F. Significantly Improved Hydrogen Storage Behaviors in MgH2 with Nb Nanocatalyst. Int. J. Miner. Metall. Mater. 2022, 29, 1788–1797. [Google Scholar] [CrossRef]
  50. Galey, B.; Auroux, A.; Sabo-Etienne, S.; Dhaher, S.; Grellier, M.; Postole, G. Improved Hydrogen Storage Properties of Mg/MgH2 Thanks to the Addition of Nickel Hydride Complex Precursors. Int. J. Hydrogen Energy 2019, 44, 28848–28862. [Google Scholar] [CrossRef]
  51. Tian, T.; Liu, F.; Zhang, X.; Wang, J.; Li, S.; Chen, W.; Song, Y.; Li, J.; Liu, G. Effect of Porous Nanosheet NiO on Hydrogen Storage Performance of MgH2. Int. J. Hydrogen Energy 2025, 135, 2–9. [Google Scholar] [CrossRef]
  52. Hristov, J. A Note on the Johnson–Mehl–Avrami–Kolmogorov Kinetic Model: An Attempt Aiming to Introduce Time Non-Locality. Eng 2025, 6, 24. [Google Scholar] [CrossRef]
  53. Jin, O.; Shang, Y.; Huang, X.; Szabó, D.V.; Le, T.T.; Wagner, S.; Klassen, T.; Kübel, C.; Pistidda, C.; Pundt, A. Transformation Kinetics of LiBH4–MgH2 for Hydrogen Storage. Molecules 2022, 27, 7005. [Google Scholar] [CrossRef]
  54. Mooij, L.; Dam, B. Nucleation and Growth Mechanisms of Nano Magnesium Hydride from the Hydrogen Sorption Kinetics. Phys. Chem. Chem. Phys. 2013, 15, 11501–11510. [Google Scholar] [CrossRef]
  55. Jain, P.; Dixit, V.; Jain, A.; Srivastava, O.N.; Huot, J. Effect of Magnesium Fluoride on Hydrogenation Properties of Magnesium Hydride. Energies 2015, 8, 12546–12556. [Google Scholar] [CrossRef]
  56. Zhang, X.L.; Liu, Y.F.; Zhang, X.; Hu, J.J.; Gao, M.X.; Pan, H.G. Empowering Hydrogen Storage Performance of MgH2 by Nanoengineering and Nanocatalysis. Mater. Today Nano 2020, 9, 100064. [Google Scholar] [CrossRef]
  57. Liang, C.; Wang, Z.; Zhang, M.; Ma, C. Enhanced Hydrogen Storage Property of MgH2 Caused by a BaCrO4 Nanocatalyst. RSC Adv. 2024, 14, 23930–23942. [Google Scholar] [CrossRef]
  58. Kazaz, S.; Billeter, E.; Longo, F.; Borgschulte, A.; Łodziana, Z. Why Hydrogen Dissociation Catalysts Do Not Work for Hydrogenation of Magnesium. Adv. Sci. 2024, 11, 2304603. [Google Scholar] [CrossRef] [PubMed]
  59. El-Eskandarany, M.S.; Shaban, E.; Ali, N.; Aldakheel, F.; Alkandary, A. In-Situ Catalyzation Approach for Enhancing the Hydrogenation/Dehydrogenation Kinetics of MgH2 Powders with Ni Particles. Sci. Rep. 2016, 6, 37335. [Google Scholar] [CrossRef] [PubMed]
  60. Yu, Q.; Qi, L.; Mishra, R.K.; Zeng, X.; Minor, A.M. Size-Dependent Mechanical Properties of Mg Nanoparticles Used for Hydrogen Storage. Appl. Phys. Lett. 2015, 106, 261903. [Google Scholar] [CrossRef]
  61. Chinthamani, S.; Kannan, G.; George, G.D.; Sreedharan, C.E.; Rajagopal, K.S. Effect of Nano B4C on the Tribological Behaviour of Magnesium Alloy Prepared Through Powder Metallurgy. Mater. Sci. 2020, 26, 392–400. [Google Scholar] [CrossRef]
  62. Öztop, H.F.; Gürgenç, E.; Gür, M. Thermophysical Properties and Enhancement Behavior of Novel B4C-Nanoadditive RT35HC Nanocomposite Phase Change Materials: Structural, Morphological, Thermal Energy Storage and Thermal Stability. Sol. Energy Mater. Sol. Cells 2024, 272, 112909. [Google Scholar] [CrossRef]
  63. Zhang, Q.; Huang, Y.; Ma, T.; Li, K.; Ye, F.; Wang, X.; Jiao, L.; Yuan, H.; Wang, Y. Facile Synthesis of Small MgH2 Nanoparticles Confined in Different Carbon Materials for Hydrogen Storage. J. Alloys Compd. 2020, 825, 153953. [Google Scholar] [CrossRef]
  64. Ma, Z.; Zhang, J.; Zhu, Y.; Lin, H.; Liu, Y.; Zhang, Y.; Zhu, D.; Li, L. Facile Synthesis of Carbon Supported Nano-Ni Particles with Superior Catalytic Effect on Hydrogen Storage Kinetics of MgH2. ACS Appl. Energy Mater. 2018, 1, 1158–1165. [Google Scholar] [CrossRef]
  65. Babić, B.; Prvulović, M.; Filipović, N.; Mravik, Ž.; Sekulić, Z.; Govedarović, S.M.; Milanović, I. Hydrogen Storage Properties of MgH2–Tm: Ni-Catalysis vs. Mechanical Milling. Int. J. Hydrogen Energy 2024, 54, 446–456. [Google Scholar] [CrossRef]
  66. Pourabdoli, M.; Raygan, S.; Abdizadeh, H.; Uner, D. Determination of Kinetic Parameters and Hydrogen Desorption Characteristics of MgH2-10 Wt% (9Ni–2Mg–Y) Nano-Composite. Int J Hydrogen Energy 2013, 38, 11910–11919. [Google Scholar] [CrossRef]
  67. Czujko, T.; Oleszek, E.E.; Szot, M. New Aspects of MgH2 Morphological and Structural Changes during High-Energy Ball Milling. Materials 2020, 13, 4550. [Google Scholar] [CrossRef] [PubMed]
  68. Liu, H.; Wang, X.; Liu, Y.; Dong, Z.; Cao, G.; Li, S.; Yan, M. Improved Hy-drogen Storage Properties of MgH2 by Ball Milling with AlH3: Prepara-tions, de/Rehydriding Properties, and Reaction Mechanisms. J Mater Chem A Mater 2013, 1, 12527–12535. [Google Scholar] [CrossRef]
  69. Ocampo, R.A.; Arias-Velandia, J.; Lenis, J.A.; Zuleta Gil, A.A.; Bello, S.; Correa, E.; Arrieta, C.; Bolívar, F.J.; Echeverría, F.E. Catalytic Effect of Commercial Carbon-Coated Nickel Nanoparticles on the Hydrogen Storage Performance of Magnesium Hydride. Int. J. Hydrogen Energy 2025, 129, 91–100. [Google Scholar] [CrossRef]
  70. Zou, J.; Long, S.; Chen, X.; Zeng, X.; Ding, W. Preparation and Hydrogen Sorption Properties of a Ni Decorated Mg Based Mg@Ni Nano-Composite. Int. J. Hydrogen Energy 2015, 40, 1820–1828. [Google Scholar] [CrossRef]
  71. Fadonougbo, J.O.; Kim, H.J.; Suh, B.C.; Suh, J.Y.; Lee, Y.S.; Shim, J.H.; Yim, C.D.; Cho, Y.W. Kinetics and Thermodynamics of near Eutectic Mg-Mg2Ni Composites Produced by Casting Process. Int. J. Hydrogen Energy 2020, 45, 29009–29022. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the process for obtaining anhydrous Ni–B electroless bath and for decorating the magnesium particles with the formulated electroless bath.
Figure 1. Schematic representation of the process for obtaining anhydrous Ni–B electroless bath and for decorating the magnesium particles with the formulated electroless bath.
Applnano 06 00016 g001
Figure 2. SEM images of Mg particles (Tangshan Weihao Magnesium Powder Co., ~26 µm). (a,b) without plating, (c,d) with Ni–P plating; (b,d) are a zoom section of (a,c); (e) EDS mapping analysis of Mg particles with Ni–P plating. (Cyan color assigned to magnesium, pink color assigned to phosphorus, orange color assigned to nickel).
Figure 2. SEM images of Mg particles (Tangshan Weihao Magnesium Powder Co., ~26 µm). (a,b) without plating, (c,d) with Ni–P plating; (b,d) are a zoom section of (a,c); (e) EDS mapping analysis of Mg particles with Ni–P plating. (Cyan color assigned to magnesium, pink color assigned to phosphorus, orange color assigned to nickel).
Applnano 06 00016 g002
Figure 3. SEM images of Mg particles (Tangshan Weihao Magnesium Powder Co., ~26 µm) with plating. (a) Secondary electron beam; (b) backscattered electron beam; (c) EDS mapping analysis of Mg particles with Ni–B plating, and (d) is a zoom of (c). (Cyan color assigned to magnesium, orange color assigned to nickel, blue color assigned to boron).
Figure 3. SEM images of Mg particles (Tangshan Weihao Magnesium Powder Co., ~26 µm) with plating. (a) Secondary electron beam; (b) backscattered electron beam; (c) EDS mapping analysis of Mg particles with Ni–B plating, and (d) is a zoom of (c). (Cyan color assigned to magnesium, orange color assigned to nickel, blue color assigned to boron).
Applnano 06 00016 g003
Figure 4. Diagram of the sample activation protocol in Sieverts-type equipment to evaluate their hydrogen sorption properties.
Figure 4. Diagram of the sample activation protocol in Sieverts-type equipment to evaluate their hydrogen sorption properties.
Applnano 06 00016 g004
Figure 5. Isothermal hydrogen absorption/desorption kinetic curves of the Mg@NiB (black curve), Mg@NiP (lilac curve), and MgH2 “sample without plating” (gray curve).
Figure 5. Isothermal hydrogen absorption/desorption kinetic curves of the Mg@NiB (black curve), Mg@NiP (lilac curve), and MgH2 “sample without plating” (gray curve).
Applnano 06 00016 g005
Figure 6. Isothermal hydrogen absorption/desorption kinetic curves of the Mg@NiB sample in its first (black curve), second (orange curve), and third (green curve) cycles, and the MgH2 sample.
Figure 6. Isothermal hydrogen absorption/desorption kinetic curves of the Mg@NiB sample in its first (black curve), second (orange curve), and third (green curve) cycles, and the MgH2 sample.
Applnano 06 00016 g006
Figure 7. (a) Temperature programmed desorption (TPD) at 5 °C/min and (b) differential scanning calorimetry (DSC) at different heating rates, 2 °C/min (red signal), 3 °C/min (green curve), and 5 °C/min for Mg@NiB sample.
Figure 7. (a) Temperature programmed desorption (TPD) at 5 °C/min and (b) differential scanning calorimetry (DSC) at different heating rates, 2 °C/min (red signal), 3 °C/min (green curve), and 5 °C/min for Mg@NiB sample.
Applnano 06 00016 g007
Figure 8. SEM images of magnesium particles decorated with an aqueous Ni–B electroless bath during the activation process (contraction and expansion cycles). Hydrogen absorption was carried out at 350 °C and 2.1 MPa, and hydrogen desorption was carried out at 350 °C and 0.000045 MPa. (a) cycle 1, (b) cycle 2, (c) cycle 3, (d) cycle 4, (e) cycle 5, (f) cycle 6, and (g) cycle 7.
Figure 8. SEM images of magnesium particles decorated with an aqueous Ni–B electroless bath during the activation process (contraction and expansion cycles). Hydrogen absorption was carried out at 350 °C and 2.1 MPa, and hydrogen desorption was carried out at 350 °C and 0.000045 MPa. (a) cycle 1, (b) cycle 2, (c) cycle 3, (d) cycle 4, (e) cycle 5, (f) cycle 6, and (g) cycle 7.
Applnano 06 00016 g008
Figure 9. XRD spectra of samples: commercial Mg particles, decorated particles, and activated particles.
Figure 9. XRD spectra of samples: commercial Mg particles, decorated particles, and activated particles.
Applnano 06 00016 g009
Figure 10. XRD spectra of samples: hydrogenated commercial Mg particles, hydrous decorated particles, and dehydrogenated decorated particles.
Figure 10. XRD spectra of samples: hydrogenated commercial Mg particles, hydrous decorated particles, and dehydrogenated decorated particles.
Applnano 06 00016 g010
Figure 11. (a) JMAK graph of Mg@NiB sample, and (b) hydrogen desorption/absorption cycles plot at 350 °C of Mg@NiB sample.
Figure 11. (a) JMAK graph of Mg@NiB sample, and (b) hydrogen desorption/absorption cycles plot at 350 °C of Mg@NiB sample.
Applnano 06 00016 g011
Figure 12. Schematic representation of metal hydride formation at the structural level, the surface-scale interaction between a Ni catalyst deposited on the Mg particle surface, and the hydrogen absorption/desorption processes in the presence of the catalyst.
Figure 12. Schematic representation of metal hydride formation at the structural level, the surface-scale interaction between a Ni catalyst deposited on the Mg particle surface, and the hydrogen absorption/desorption processes in the presence of the catalyst.
Applnano 06 00016 g012
Table 1. Anhydrous Ni–B electroless bath formulation for 1 L in ethylene glycol.
Table 1. Anhydrous Ni–B electroless bath formulation for 1 L in ethylene glycol.
Chemical Composition
ComponentComposition (g/L)
C4H6O62.5
NiCl211.3
NaOH9.3
NaBH41.7
Mixing conditions
Stirring speed>600 RPM
Temperature (°C)70 ± 5
Time (minutes)25
Plating conditions
Load bath4 mg/mL
Stirring speed>600 RPM
Temperature (°C)60 ± 1
Time (minutes)5
Centrifugation-assisted washing conditions (≥6000 RPM)
Centrifugal separation time (minutes)45
Centrifugal wash acetone time (minutes)15
Components are added in the order of the table.
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

Bello, S.; Aguirre Ocampo, R.; Arias Velandia, J.; Zuleta Gil, A.; Correa, E.; Silva, W.; Lenis Rodas, J.A.; Arrieta, C.; Bolívar, F.; Nieto, C.; et al. Electroless Nickel Plating of Magnesium Particles for Hydrogen Storage. Appl. Nano 2025, 6, 16. https://doi.org/10.3390/applnano6030016

AMA Style

Bello S, Aguirre Ocampo R, Arias Velandia J, Zuleta Gil A, Correa E, Silva W, Lenis Rodas JA, Arrieta C, Bolívar F, Nieto C, et al. Electroless Nickel Plating of Magnesium Particles for Hydrogen Storage. Applied Nano. 2025; 6(3):16. https://doi.org/10.3390/applnano6030016

Chicago/Turabian Style

Bello, Sindy, Robinson Aguirre Ocampo, Julián Arias Velandia, Alejandro Zuleta Gil, Esteban Correa, Wilber Silva, Julián Andrés Lenis Rodas, Carlos Arrieta, Francisco Bolívar, Cesar Nieto, and et al. 2025. "Electroless Nickel Plating of Magnesium Particles for Hydrogen Storage" Applied Nano 6, no. 3: 16. https://doi.org/10.3390/applnano6030016

APA Style

Bello, S., Aguirre Ocampo, R., Arias Velandia, J., Zuleta Gil, A., Correa, E., Silva, W., Lenis Rodas, J. A., Arrieta, C., Bolívar, F., Nieto, C., & Echeverria, F. (2025). Electroless Nickel Plating of Magnesium Particles for Hydrogen Storage. Applied Nano, 6(3), 16. https://doi.org/10.3390/applnano6030016

Article Metrics

Back to TopTop