Next Article in Journal
Merits and Demerits of Machine Learning of Ferroelectric, Flexoelectric, and Electrolytic Properties of Ceramic Materials
Previous Article in Journal
Static Compaction on Coupled Precursors and Optimizing Molarity for Enhanced Strength and Durability of Geopolymer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ball Milling Innovations Advance Mg-Based Hydrogen Storage Materials Towards Practical Applications

1
Laboratory for Functional Materials, School of New Energy Materials and Chemistry, Leshan Normal University, Leshan 614000, China
2
Leshan West Silicon Materials Photovoltaic New Energy Industry Technology Research Institute, Leshan 614000, China
3
College of Materials Science and Engineering, National Engineering Research Center for Magnesium Alloys, National Innovation Center for Industry-Education Integration of Energy Storage Technology, Chongqing University, Chongqing 400044, China
4
School of Automotive Engineering, Yancheng Institute of Technology, Yancheng 224051, China
5
Department of Computer Science, Illinois Institute of Technology, Chicago, IL 60616, USA
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(11), 2510; https://doi.org/10.3390/ma17112510
Submission received: 23 April 2024 / Revised: 19 May 2024 / Accepted: 20 May 2024 / Published: 23 May 2024
(This article belongs to the Section Materials Chemistry)

Abstract

:
Mg-based materials have been widely studied as potential hydrogen storage media due to their high theoretical hydrogen capacity, low cost, and abundant reserves. However, the sluggish hydrogen absorption/desorption kinetics and high thermodynamic stability of Mg-based hydrides have hindered their practical application. Ball milling has emerged as a versatile and effective technique to synthesize and modify nanostructured Mg-based hydrides with enhanced hydrogen storage properties. This review provides a comprehensive summary of the state-of-the-art progress in the ball milling of Mg-based hydrogen storage materials. The synthesis mechanisms, microstructural evolution, and hydrogen storage properties of nanocrystalline and amorphous Mg-based hydrides prepared via ball milling are systematically reviewed. The effects of various catalytic additives, including transition metals, metal oxides, carbon materials, and metal halides, on the kinetics and thermodynamics of Mg-based hydrides are discussed in detail. Furthermore, the strategies for synthesizing nanocomposite Mg-based hydrides via ball milling with other hydrides, MOFs, and carbon scaffolds are highlighted, with an emphasis on the importance of nanoconfinement and interfacial effects. Finally, the challenges and future perspectives of ball-milled Mg-based hydrides for practical on-board hydrogen storage applications are outlined. This review aims to provide valuable insights and guidance for the development of advanced Mg-based hydrogen storage materials with superior performance.

Graphical Abstract

1. Introduction

Hydrogen, as a clean and sustainable energy carrier, has attracted increasing attention for its potential to mitigate the environmental and energy crisis. Mg-based materials are promising candidates for solid-state hydrogen storage due to their high theoretical capacity (7.6 wt.% for MgH2), low cost, and abundant reserves [1,2,3,4]. However, the high thermodynamic stability and sluggish kinetics of Mg-based hydrides have hindered their practical applications [5,6,7]. To address these challenges, various strategies have been developed, among which ball milling has emerged as a versatile and effective technique for the synthesis and modification of Mg-based hydrogen storage materials [8,9,10].
Ball milling is a solid-state processing technique that involves repeated welding, fracturing, and rewelding of powder particles in a high-energy ball mill [11]. The high-energy impacts and shear forces during milling can induce a variety of physical and chemical changes in the materials, such as reductions in particle size, the introduction of defects, the formation of metastable phases, and enhancements in surface reactivity [12]. These microstructural and morphological changes can significantly improve the hydrogen storage properties of Mg-based materials, including their hydrogen capacity, absorption/desorption kinetics, and cycle stability [13,14]. Mechanical ball milling, as an efficient solid-state reaction method, exhibits many unique advantages in material mixing and mechanical alloying processes. Through the high-speed rotation of the ball milling media, using shear and impact forces, it achieves efficient mixing and reaction of materials, not only enhancing material uniformity but also significantly promoting chemical reactions and alloying processes between materials. This method has strong controllability, allowing precise adjustment of the final product properties by modulating the milling time, speed, and media. Compared with traditional alloying methods, mechanical ball milling does not require an external heat source, significantly saving energy and costs. Moreover, it has a wide applicability and is suitable for processing metals, ceramics, plastics, and complex alloy systems, demonstrating excellent versatility and flexibility. Mechanical ball milling also has good scalability and is suitable for both large-scale industrial production and small-scale experimental research, showing important practical application value.
In recent years, extensive research efforts have been devoted to the ball milling of Mg-based hydrogen storage materials, leading to substantial progress in understanding the underlying mechanisms and structure–property relationships. Various Mg-based hydrides, nanostructures, catalysts, and nanocomposites with enhanced hydrogen storage performance have been synthesized via ball milling [15,16,17]. Moreover, advanced characterization techniques and theoretical calculations have provided deep insights into the microstructural evolution and hydrogenation/dehydrogenation behaviors of the ball-milled materials [18,19].
This review aims to provide a comprehensive summary of the state-of-the-art progress in the ball milling of Mg-based hydrogen storage materials. The synthesis of Mg-based hydrides, nanostructures, and catalysts via ball milling will be systematically reviewed, with an emphasis on the effects of milling parameters and additives on the microstructure and hydrogen storage properties. The mechanisms of hydrogen absorption/desorption in the ball-milled materials will be discussed based on advanced characterization and theoretical studies. The limitations and future perspectives of ball milling in advancing Mg-based hydrogen storage materials towards practical applications will also be outlined.

2. Synthesis of Mg-Based Hydrides via Ball Milling

2.1. Mg-Based Binary Hydrides

MgH2 is an efficient hydrogen storage material with a theoretical hydrogen storage capacity of up to 7.6 wt.%. It can be synthesized through various methods, such as hydrogenation and chemical vapor deposition, producing high-purity MgH2. However, its high thermodynamic stability (reaction enthalpy change ΔH = −74.5 kJ·mol−1 H2, entropy change ΔS = −135 J·K−1·mol−1 H2) results in a high dehydrogenation temperature (>350 °C) and slow reaction rate. The discrepancy between the actual and theoretical hydrogen storage capacity significantly limits its practical application [20]. At present, ball milling has been widely used to synthesize MgH2 with improved hydrogen storage properties.
Ball milling under a hydrogen atmosphere has proven to be an effective strategy to mitigate particle agglomeration and oxidation issues that commonly arise during the milling process. Lu et al. [21] reported that after milling, the surface area of pure Mg exhibited a substantial increase from 0.61 m2/g to 6.16 m2/g, accompanied by a reduction in crystal size. The beneficial effects of milling under a hydrogen atmosphere were further demonstrated by the enhanced hydrogen storage capacity of the milled Mg powder. At 350 °C, the hydrogen storage capacity of Mg powder milled under a hydrogen atmosphere reached 3.36 wt.% H2, a significant improvement compared to the 0.14 wt.% achieved using conventional ball milling.
The hydrogen storage properties of ball-milled MgH2 can be further improved by optimizing the milling parameters, such as milling time, milling speed, ball-to-powder ratio, and hydrogen pressure [22]. Varin et al. [23] systematically investigated the effects of milling time on the microstructure and hydrogen storage properties of MgH2. They found that the hydrogen desorption temperature decreased with increasing milling time, reaching a minimum of 277 °C after 20 h of milling. Prolonged milling led to the agglomeration of MgH2 particles and the introduction of impurities, which deteriorated the hydrogen storage properties.

2.2. Mg-Based Ternary Hydrides

Ball milling has also been applied to synthesize Mg-based ternary and complex hydrides with improved thermodynamic and kinetic properties. By adding metal elements, the crystal structure can be regulated to promote the formation of new crystal phases, improving hydrogen absorption and desorption kinetics and stability. The interfacial catalytic effects of metal elements accelerate hydrogen absorption and desorption, reducing the activation energy of the reaction. It adjusts the hydrogen diffusion behavior in MgH2, speeding up hydrogenation and dehydrogenation reactions, enhancing thermodynamic stability, lowering thermal decomposition temperatures, and extending cycle life. Mg2CoH5, with a hydrogen content of 4.5 wt.%, has been reported to have a lower decomposition temperature and faster kinetics than MgH2 [24]. Apart from Mg2CoH5, other Mg-based binary hydrides, such as Mg2NiH4 and Mg2FeH6, have also been successfully synthesized via ball milling. Mg2NiH4 exhibits a higher hydrogen content of 5 wt.% and a lower decomposition temperature of 240 °C compared to Mg2CoH5 [25].
Complex hydrides, such as Mg(NH2)2 and Mg(BH4)2, have attracted increasing attention due to their high hydrogen contents (7.2 wt.% for Mg(NH2)2 and 14.9 wt.% for Mg(BH4)2) [26,27]. However, their high decomposition temperatures and poor reversibility have hindered their practical applications. Ball milling has been used to synthesize these complex hydrides with improved hydrogen storage properties. The ball-milled Mg(NH2)2 exhibited a hydrogen desorption temperature of 150 °C, which was 250 °C lower than that of the original Mg(NH2)2.
Mg(BH4)2 has been synthesized by ball milling a mixture of MgB2 and LiBH4 [28]. The ball-milled Mg(BH4)2 showed a hydrogen desorption temperature of 260 °C, which was 140 °C lower than that of the as-synthesized Mg(BH4)2. The improved kinetics were attributed to the reduced particle size and the formation of LiMg(BH4)3 during ball milling, which destabilized the Mg(BH4)2 structure. Table 1 summarizes the preparation methods and hydrogen storage properties of Mg-based ternary and complex hydrides. Ball milling of MgH2 with LiNH2 and LiBH4 can produce Mg(NH2)2 and Mg(BH4)2, respectively. These complex hydrides have higher hydrogen capacities and lower dehydrogenation temperatures than binary hydrides. Mg(NH2)2 decomposes to Li2Mg(NH)2 during dehydrogenation, while ball milling of Mg(BH4)2 with LiBH4 forms LiMg(BH4)3.

2.3. Mechanisms of Mg-Based Hydride Formation during Ball Milling

The formation of Mg-based hydrides during ball milling involves a series of physical and chemical processes, including particle refinement, surface activation, and chemical reactions [29]. The high-energy impacts during ball milling can effectively reduce the particle size and create fresh surfaces, which facilitate the diffusion and dissociation of hydrogen molecules. The accumulation of defects and strain energy during milling also enhances the reactivity of the Mg-based materials towards hydrogen [30].
The formation mechanism of MgH2 during ball milling under a hydrogen atmosphere has been extensively studied. Czujko et al. [31] proposed a three-stage mechanism for the formation of MgH2 during ball milling: (i) refinement of Mg particles and creation of fresh surfaces; (ii) chemisorption of hydrogen on the Mg surfaces and nucleation of MgH2; (iii) growth of MgH2 phase and further refinement of the particles. The rate-limiting step is the nucleation of MgH2, which requires a critical hydrogen pressure and a sufficient number of active sites on the Mg surfaces.
For Mg-based binary and ternary hydrides, the formation mechanism during ball milling involves the interdiffusion and reaction of the constituent elements under a hydrogen atmosphere. The high-energy impacts can induce the formation of metastable phases and enhance the atomic diffusion, which facilitate the formation of the hydride phases [32]. For example, during the ball milling of Mg and Ni powders under a hydrogen atmosphere, the formation of Mg2NiH4 involves the following steps: (i) formation of Mg2Ni alloy; (ii) hydrogenation of Mg2Ni to form Mg2NiH0.3; (iii) disproportionation of Mg2NiH0.3 into MgH2 and Mg2NiH4.
The formation of Mg-based complex hydrides during ball milling involves the solid-state reaction between MgH2 and the complex hydride precursors (e.g., LiNH2, LiBH4). The high-energy impacts can break the B-H and N-H bonds in the complex hydrides and facilitate the exchange of H atoms with MgH2, leading to the formation of Mg(NH2)2 and Mg(BH4)2 [33]. The simultaneous refinement of the particles and mixing of the reactants also enhance the contact area and reactivity, which promote the solid-state reactions.

3. Nanostructuring of Mg-Based Hydrides via Ball Milling

Nanostructuring has been recognized as an effective strategy to improve the hydrogen storage properties of Mg-based hydrides [34]. Nanostructured materials have a high surface-to-volume ratio, short diffusion paths, and a large number of grain boundaries and defects, which can enhance the hydrogen absorption/desorption kinetics and reduce the thermal stability [35]. Ball milling is a simple and efficient technique to produce nanostructured Mg-based hydrides with a high yield and low cost.

3.1. Nanocrystalline Mg-Based Hydrides

The reversible reaction of hydrogen with metal Mg to form MgH2 has special properties for energy storage. However, the high thermodynamic stability of the hydride results in high absorption/desorption temperatures. Nanocrystalline Mg-based hydrides have been extensively studied for hydrogen storage applications. Modifying processing routes or combinatorial routes can also lead to refined microstructures, including methods such as melt spinning, cold rolling, and mechanical milling, which can be used to prepare nanocrystalline MgH2 structures.
As the particle size is reduced below a critical value, a remarkable decrease in the desorption temperature is observed. Specifically, when the particle size is reduced to the range of 500–600 nm, the onset desorption temperature of MgH2 exhibits a substantial reduction of approximately 40–60 °C. Microstructural analysis of MgH2 powder subjected to ball milling for durations exceeding 10 h reveals the coexistence of metastable γ-MgH2 and stable nanocrystalline β-MgH2 phases. Quantitative evidence suggests that the refinement of powder granularity and the presence of the γ-MgH2 phase within the particles are the primary factors contributing to the significant reduction in the desorption temperature of MgH2 hydride [36].
MgH2 nanowires exhibit significantly lower desorption barriers (33.5 and 38.8 kJ mol−1 for hydrogenation and dehydrogenation, respectively) compared to commercial MgH2 (120–142 kJ·mol−1). Table 2 summarizes the preparation conditions and hydrogen storage properties of nanocrystalline Mg-based hydrides. Theoretical predictions suggest that reducing the diameter of the nanowires below 30 nm can substantially influence the thermodynamics and kinetics of MgH2 [37]. Furthermore, experimental investigations on the hydrogen storage performance of MgH2 with varying particle sizes (25, 32, 38 nm) provide compelling evidence that smaller particle sizes lead to enhanced hydrogenation kinetics [38].

3.2. Amorphous Mg-Based Hydrides

Amorphous Mg-based hydrides have also been synthesized via ball milling and have shown improved hydrogen storage properties compared to their crystalline counterparts. Unlike crystalline materials, amorphous hydrides lack long-range atomic order and have a high density of defects and free volume, which can facilitate hydrogen diffusion and reduce the stability of the hydrides [39,40,41].
Zhang et al. [42] synthesized amorphous CeMg11Ni alloys with varying Ni contents (x wt.% Ni, where x = 100, 200) via ball milling. The degree of amorphization of the alloy was found to increase with increasing Ni contents. The hydrogen storage capacities of the alloys with x = 100 and x = 200 were determined to be 5.94 wt.% and 6.15 wt.%, respectively. Moreover, the dehydrogenation rate exhibited a consistent increase with increasing Ni contents.
Liang et al. [43] successfully synthesized an amorphous La@Mg compound by ball milling a mixture of (La(acac)3) and Mg. This composite material exhibited a remarkable hydrogen storage capacity of approximately 7.6% and a hydrogen desorption rate of 7.2%, surpassing the performance of its crystalline La@Mg counterpart and pure Mg. The adsorption/desorption kinetics were rapid, and the reversible adsorption/desorption cycles demonstrated excellent stability. Theoretical calculations in conjunction with experimental results revealed that the amorphous La@Mg structure provides channels for enhanced hydrogen diffusion. This unique feature facilitates the hydrogenation process by accelerating the diffusion of H atoms between the subsurface and surface regions. Table 3 summarizes the preparation conditions and hydrogen storage properties of amorphous Mg-based hydrides. Long-time high-energy ball milling with the addition of alloying elements can produce amorphous MgH2 and Mg2NiH4. Amorphous hydrides exhibit better hydrogen storage kinetics than their crystalline counterparts due to the increased defects and free volume. The addition of TiF3 and VTiCr can further improve the amorphization and hydrogen storage properties of MgH2 and Mg2NiH4.

3.3. Mechanisms of Nanostructure/Amorphous Formation during Ball Milling

The formation of nanostructures during ball milling involves a series of physical and chemical processes, including particle refinement, strain accumulation, and phase transformation [44]. Figure 1a illustrates the preparation process, where high-energy collisions during ball milling effectively reduce the particle size and introduce a high density of defects and strain energy into the materials [45].
The formation mechanism of nanocrystalline structures during ball milling has been extensively studied. As depicted in Figure 1b, the high-energy impacts during ball milling lead to repeated particle deformation, fractures, and welding, resulting in a significant reduction in particle size and a concomitant increase in surface area [46]. Fecht [47] proposed a three-stage model for the formation of nanocrystalline structures during ball milling: (i) initial stage: refinement of particles and introduction of defects; (ii) intermediate stage: formation of nanocrystalline domains and grain boundaries; (iii) final stage: further refinement of nanocrystalline domains and saturation of grain size. The final grain size is determined by the balance between the plastic deformation and dynamic recovery during ball milling.
The formation of amorphous structures during ball milling involves the accumulation of strain energy and the suppression of crystallization [48]. When the strain energy stored in the materials reaches a critical value, the crystalline structure becomes unstable and transforms into an amorphous state. The critical strain energy depends on the nature of the materials and the milling conditions. For Mg-based hydrides, the formation of amorphous structures is favored by the presence of additives (e.g., Ni, Y) and the high energy input during ball milling [49]. Figure 1c reveals the evolution of the microstructure during the milling process. As the milling progresses, amorphous phases gradually nucleate and grow at the expense of the crystalline phases due to the accumulation of strain energy. The crystalline domains continuously shrink and ultimately disappear, resulting in a fully amorphous structure [50].
Figure 1. (a) Flowchart for the preparation method of the as-milled Sm5Mg41 alloy [45]. (b) Illustration of the deformation of powder agglomerate during the impact process [46]. (c) Illustration of the different microstructural states of the Mg2Ni alloys [50].
Figure 1. (a) Flowchart for the preparation method of the as-milled Sm5Mg41 alloy [45]. (b) Illustration of the deformation of powder agglomerate during the impact process [46]. (c) Illustration of the different microstructural states of the Mg2Ni alloys [50].
Materials 17 02510 g001
The formation of nanostructures during the ball milling process is influenced by several factors, including the hydrogen pressure and the presence of additives. Cuevas et al. [51] successfully synthesized MgH2-TiH2 nanocomposites with grain sizes ranging from 4 to 12 nm through reactive ball milling under a hydrogen pressure of 8 MPa. The catalytic properties of TiH2 played a crucial role in accelerating the formation of the MgH2 phase, enabling the nanocomposite to form within a remarkably short duration of less than 50 min. Furthermore, the TiH2 phase effectively inhibited the coarsening of Mg grains, allowing the MgH2 phase to nucleate and subsequently form a continuous hydride layer within the Mg nanoparticles.

4. Catalytic Modification of Mg-Based Hydrides via Ball Milling

Catalytic modification has been widely used to enhance the hydrogen storage properties of Mg-based hydrides [52]. The addition of catalysts can lower the activation energy for hydrogen absorption/desorption, increase the reaction kinetics, and improve the reversibility and cycling stability of the hydrides [53]. Ball milling is an effective technique to introduce catalysts into Mg-based hydrides and achieve a uniform distribution and strong interaction between the catalysts and the hydrides [54].

4.1. Transition Metal Catalysts

Transition metals, such as Ti, V, Mn, Fe, Co, Ni, Cu, and Nb, have been extensively studied as catalysts for Mg-based hydrides [55]. These metals have a high catalytic activity for hydrogen dissociation and recombination and can form stable hydrides with a high hydrogen storage capacity [56]. The catalytic effects of transition metals on the hydrogen storage properties of Mg-based hydrides have been attributed to several mechanisms, including the spillover effect, the gateway effect, and the hydride-forming effect [57,58,59].
Liang et al. [60] investigated the catalytic effects of different transition metals (Ti, V, Mn, Fe, Ni) on the hydrogen storage properties of MgH2. They found that the addition of 1 mol% of transition metals via ball milling significantly reduced the dehydrogenation temperature and activation energy of MgH2. Composites with Ti or V added exhibit rapid desorption kinetics above 250 °C and rapid adsorption kinetics below 25 °C. The enhanced catalytic effect of Ti and V was attributed to their strong interaction with MgH2 and the formation of stable hydrides (TiH2 and VH2) during dehydrogenation.
Wu et al. [61] facilitated the ball milling of Nb and MgH2 using surfactants. The MgH2-5 wt.% Nb composite material began releasing hydrogen at 186.7 °C, with a maximum release of 7.0 wt.%. It released 4.2 wt.% H2 within 14 min at 250 °C and could absorb 4.0 wt.% H2 within 30 min, even at the low temperature of 100 °C. The desorption activation energy and hydrogenation activation energy were reduced from 140.51 ± 4.74 and 70.67 ± 2.07 kJ·mol−1 to 90.04 ± 2.83 and 53.46 ± 3.33 kJ·mol−1, respectively.
The catalytic effects of transition metals on the hydrogen storage properties of Mg-based hydrides can be further enhanced by optimizing the ball milling conditions and the catalyst composition. Lu et al. [62] successfully synthesized a MgH2-0.1TiH2 nanocomposite system with exceptionally fine nanocrystalline grain sizes ranging from 5 to 10 nm. This was achieved through an innovative ultra-high-energy-high-pressure mechanical milling approach. The TiH2 phase was homogeneously dispersed among the MgH2 particles, forming a well-distributed nanocomposite. The synergistic effects of the nanoscale structure and the catalytic influence of TiH2 led to a remarkable enhancement in the dehydrogenation and hydrogenation kinetics of the MgH2-0.1TiH2 system compared to commercial MgH2.
Cui et al. [63] developed an innovative approach to enhance the catalytic effects of Ti on MgH2 by coating ball-milled Mg powder (approx. 1 μm in diameter) with a multivalent Ti-based catalyst. The coating process involved a chemical reaction between the Mg powder and TiCl3 in a THF solution, resulting in the formation of a 10 nm thick catalyst layer. This layer contained multiple valence states of Ti, including Ti, TiH2, TiCl3, and TiO2. The presence of these diverse Ti valence states played a crucial role in facilitating electron transfer between Mg2+ and H, thereby promoting the recombination of H2 on the Ti surface. The resulting MgH2-encapsulated Ti-based system (Mg-Ti) demonstrated a remarkable dehydrogenation performance, with hydrogen release initiating at around 175 °C and achieving a release of 5 wt.% H2 within a mere 15 min at 250 °C.
Table 4 summarizes the effects of transition metal catalysts on the hydrogen storage properties of Mg-based hydrides. The addition of Ti, V, Mn, Fe, Ni, Nb, and their alloys via ball milling can significantly improve the hydrogen storage kinetics of MgH2 and Mg2NiH4. Ti and V show the best catalytic effects, while the catalytic effects of Mn, Fe, and Ni are relatively weak. The catalytic effects depend on the catalyst composition, milling time, and milling parameters.

4.2. Metal Oxide Catalysts

Metal oxides such as Nb2O5, TiO2, V2O5, Cr2O3, and Al2O3 have also been studied as catalysts for Mg-based hydrides [64]. These oxides have a high chemical stability and a strong interaction with hydrides, and they can act as dispersants and grain growth inhibitors during ball milling [65]. The catalytic effects of metal oxides on the hydrogen storage properties of Mg-based hydrides have been attributed to several mechanisms, including the formation of active species, the creation of hydrogen diffusion channels, and the modification of the electronic structure of the hydrides [66].
Barkhordarian et al. [67] investigated the catalytic effects of Nb2O5 on the hydrogen storage properties of MgH2. By ball milling 0.5 mol.% Nb2O5 with MgH2, they observed a significant enhancement in the reaction kinetics of the composite. The catalytic influence of Nb2O5 was particularly evident at elevated temperatures. At 300 °C, the composite exhibited rapid hydrogen release, achieving a 7 wt.% H2 release within a mere 90 s. Similarly, the composite demonstrated fast hydrogen absorption, with 7 wt.% H2 being absorbed within just 60 s at the same temperature. Even at a lower temperature of 250 °C, the composite showcased impressive kinetics, absorbing over 6 wt.% H2 within 60 s and subsequently releasing it within 500 s.
Polanski et al. [68] explored the catalytic effects of nanocrystalline Cr2O3 on the hydrogen storage properties of MgH2. Interestingly, they discovered that the addition of Cr2O3 did not significantly influence the grain size of β-MgH2. However, despite the lack of grain size reduction, the presence of Cr2O3 had a profound impact on the reaction kinetics of the composite. The MgH2-Cr2O3 composite exhibited remarkable hydrogen absorption and desorption rates. At 300 °C and a hydrogen pressure of 1 MPa, the composite could absorb 6 wt.% H2 within an impressively short duration of 2 min. Moreover, under the same temperature conditions, the composite demonstrated rapid hydrogen release, achieving a 6 wt.% H2 release within just 10 min.
The catalytic effects of metal oxides on the hydrogen storage properties of Mg-based hydrides can be further enhanced by optimizing the ball milling conditions and the catalyst morphology. Ma et al. [69] conducted a comprehensive study on the influence of catalyst morphology on the hydrogen storage properties of MgH2. They discovered that the morphology of the catalyst played a crucial role in determining its surface energy and chemical interactions with Mg and hydrogen. To investigate this effect, they synthesized anatase TiO2 with different crystal face advantages using a hydrothermal method. These TiO2 catalysts, denoted as TFx (x = 0, 10, 30, 50, 70, and 80), were then incorporated into MgH2 via ball milling. Among the prepared composites, MgH2-TF70 exhibited the most promising hydrogen adsorption kinetics, with an exceptionally low apparent activation energy for dehydrogenation of only 76.1 ± 1.6 kJ mol−1. This composite material demonstrated an initial hydrogen release temperature of approximately 220 °C. Furthermore, it showcased rapid hydrogen absorption and desorption capabilities, absorbing 5.3 wt.% H2 within a remarkable 44 s at 200 °C and releasing 6.4 wt.% H2 within 700 s at 300 °C.
Table 5 summarizes the effects of metal oxide catalysts on the hydrogen storage properties of Mg-based hydrides. The addition of Nb2O5, TiO2, Cr2O3, and V2O5 via ball milling can improve the hydrogen storage kinetics of MgH2. The catalytic effects of metal oxides are relatively weaker than those of transition metals, but they can be enhanced by optimizing the milling conditions. The milling atmosphere, milling time, and milling parameters have significant effects on the catalytic activity and stability of metal oxides.

4.3. Mechanisms of Catalytic Effects during Ball Milling

The catalytic effects of transition metals and metal oxides on the hydrogen storage properties of Mg-based hydrides involve a series of physical and chemical processes during ball milling and hydrogen absorption/desorption cycles [70]. The high-energy impacts during ball milling can induce the refinement and uniform distribution of the catalysts in the hydride matrix, creating a large number of catalyst–hydride interfaces and active sites for hydrogen storage [71]. The catalysts can also act as dispersants and grain growth inhibitors, preventing the agglomeration and coarsening of the hydride particles during prolonged milling and cycling [72].
During the hydrogen absorption/desorption processes, the catalysts can facilitate the dissociation and recombination of hydrogen molecules on the surface of the hydrides, lowering the activation energy and increasing the reaction rates [73]. The dissociated hydrogen atoms can migrate from the catalyst surface to the hydride surface via the spillover effect and then diffuse into the bulk of the hydride via the gateway effect [74]. The presence of catalysts can also modify the electronic structure of the hydrides, weakening the Mg-H bonds and facilitating hydrogen desorption [75].
The specific catalytic mechanisms depend on the nature of the catalysts and the hydrides, as well as the reaction conditions. For transition metal catalysts, the formation of stable hydrides (e.g., TiH2, VH2, NbH0.89) during the hydrogen absorption/desorption cycles can provide additional catalytic sites and hydrogen diffusion channels, enhancing the overall kinetics [76]. For metal oxide catalysts, the reduction of the oxides to lower valence states or metallic phases during the hydrogen absorption/desorption cycles can create active species and defects, which facilitate the hydrogen storage reactions [77].
The catalytic mechanisms can also involve the formation of intermediate phases and solid solutions during the hydrogen absorption/desorption cycles. For example, the addition of Fe to MgH2 can lead to the formation of the Mg2FeH6 intermediate phase during hydrogen absorption, which enhances the hydrogen diffusion and storage capacity [78]. The addition of Nb to MgH2 can lead to the formation of a Mg-Nb solid solution during dehydrogenation, which improves the reversibility and cycling stability of the hydride [79].
To elucidate the catalytic mechanisms of transition metals, we take Ni as a representative example. Figure 2a presents micron-sized Ni-encapsulated MgH2 composite materials with an average particle size ranging from 2 to 10 μm, where the Ni particles exhibit a size comparable to that of MgH2 [80]. In contrast, Figure 2b showcases MgH2-nNi (2 h) samples, revealing nanoscale Ni particles encapsulating MgH2 with an average particle size of 1–5 μm. Compared to pure MgH2, the incorporation of Ni via doping leads to the partial substitution of Mg atoms, resulting in an increased overlap of electron clouds (Figure 2c) [81]. This phenomenon indicates the formation of stronger covalent interactions between Mg and Ni atoms, which consequently weakens the Mg-H bond strength. As a result, the kinetics of hydrogen absorption and desorption reactions in the Ni-doped MgH2 composite material are significantly enhanced [81].
In the case of metal oxides, the in situ formation of catalytically active alloy phases during the hydrogen absorption and desorption processes plays a pivotal role in enhancing the hydrogen storage performance of Mg-based hydrides. Figure 2d illustrates the catalytic mechanism of NiO/NiMoO4-doped MgH2 composite materials. During the hydrogen absorption and desorption processes, the composite undergoes an in situ formation of catalytically active Mg2Ni/Mg2NiH4 and Mo phases. These newly formed species play a crucial role in accelerating the dissociation of H2 molecules on the surface of MgH2. Moreover, they create a “hydrogen pump” effect, facilitating the reversible hydrogen absorption and desorption in MgH2 under relatively mild conditions [82].
The optimization of the catalytic effects requires a comprehensive understanding of the structure–property relationships and the reaction mechanisms of the catalyst–hydride systems. Advanced characterization techniques such as X-ray diffraction (XRD), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), and neutron scattering have been used to investigate the microstructure, morphology, electronic structure, and hydrogen storage properties of ball-milled Mg-based hydrides with catalysts. Theoretical calculations such as density functional theory (DFT) and molecular dynamics (MD) simulations have also been used to predict the catalytic activity and the hydrogen storage mechanisms of the catalyst–hydride systems.

5. Nanocomposite Mg-Based Hydrides via Ball Milling

Nanocomposite Mg-based hydrides, consisting of nanosized Mg-based hydrides and other functional materials (e.g., carbon materials, metal hydrides, metal organic frameworks), have attracted increasing attention for hydrogen storage applications [83]. The incorporation of functional materials into the Mg-based hydrides can create synergistic effects, such as improved thermal conductivity, enhanced hydrogen diffusion, and reduced hydride stability, leading to superior hydrogen storage properties [84]. Ball milling is a simple and effective technique to prepare nanocomposite Mg-based hydrides with a uniform distribution and strong interaction between the components.

5.1. Carbon-Containing Nanocomposites

Carbon materials, such as graphite, carbon nanotubes (CNTs), and graphene, have been widely used to prepare nanocomposite Mg-based hydrides due to their high thermal conductivity, high surface area, and excellent mechanical properties [85]. The addition of carbon materials to Mg-based hydrides via ball milling can enhance the thermal management and hydrogen diffusion of the system, leading to improved hydrogen absorption/desorption kinetics and reversibility [86].
Pal et al. [87] reported a significant reduction in the dehydrogenation temperature of MgH2 by ball milling with 10 wt.% flake graphene. The dehydrogenation temperature decreased from 410 °C for pure MgH2 to 282 °C for the MgH2–graphene composite. This reduction in temperature is noteworthy, as it is substantially lower than the dehydrogenation temperature of pure MgH2. The incorporation of graphene also led to a decrease in the dehydrogenation activation energy from 170 kJ/mol for pure MgH2 to 136 ± 2 kJ/mol for the composite. Interestingly, XRD analysis before and after the dehydrogenation reaction confirmed the presence of graphene, indicating that no direct chemical reaction occurred between graphene and MgH2.
Rather et al. [88] investigated the hydrogen storage properties of ball-milled MgH2 with 5 wt.% carbon nanotubes (CNTs) and observed a hydrogen absorption capacity of 5.5 wt.% under a hydrogen pressure of 4.6 MPa at 673 K. Kinetic rate modeling studies provided valuable insights into the role of CNTs in enhancing hydrogen storage performance. The results indicated that CNTs facilitate the diffusion of hydrogen within the Mg matrix, thereby accelerating the formation of the hydride phase. Furthermore, the interaction between Mg and C atoms created abundant sites for hydrogen dissociation and diffusion. The unique tubular structure of CNTs also contributed to improved internal H2 transport within the composite material.
Peng et al. [89] synthesized a novel MgH2@CA microsphere composite material by ball milling carbon aerogel (CA) microspheres, which possessed a mesh-like internal structure, with Mg powder followed by an activation process. The resulting composite exhibited a distinctive core–shell structure that significantly enhanced the hydrogenation and dehydrogenation rates. The composite material demonstrated rapid hydrogen absorption, achieving a capacity of 6.2 wt.% H2 within a mere 5 min at 275 °C. Additionally, it exhibited fast hydrogen release, liberating 4.9 wt.% H2 within 100 min at 350 °C. The apparent activation energy for dehydrogenation was remarkably reduced to 114.8 kJ/mol. These improvements were attributed to the in situ-formed MgH2 nanoparticles, which reduced the diffusion distance of H2, and the CA matrix, which provided nucleation sites and prevented particle agglomeration.
Table 6 summarizes the effects of carbon materials on the hydrogen storage properties of Mg-based hydride nanocomposites. The addition of graphite, carbon nanotubes (CNTs), graphene, and fullerene (C60) via ball milling can significantly improve the hydrogen storage kinetics of MgH2 and Mg2NiH4. The catalytic effects of carbon materials depend on their morphology, size, structure, and content. Carbon materials can enhance the thermal conductivity, facilitate the hydrogen diffusion, and prevent the particle aggregation of Mg-based hydrides.

5.2. Metal Hydride-Containing Nanocomposites

Metal hydrides such as LiBH4, NaAlH4, and TiH2 have been used to prepare nanocomposite Mg-based hydrides due to their high hydrogen storage capacity and catalytic effects [90,91]. The addition of metal hydrides to Mg-based hydrides via ball milling can create hydrogen diffusion channels and active sites, leading to enhanced hydrogen absorption/desorption kinetics and reversibility [92].
Ding et al. [93] further advanced the ball milling approach by introducing LiBH4 nanoparticles in aerosol form (Figure 3a), which resulted in the formation of MgH2-LiBH4 composite materials with a superior hydrogen release performance. Figure 3b reveals the hydrogen release mechanism in pristine MgH2, where the contraction of the MgH2 core during dehydrogenation leads to the formation of a continuous Mg shell. This Mg shell acts as a barrier, hindering the decomposition of MgH2 and the diffusion of H2 molecules. However, the introduction of LiBH4 nanoparticles alters the reaction pathway (Figure 3c). At the Mg/LiBH4 interface, new reactions take place, resulting in the formation of MgB2 and LiH. These interfacial reactions effectively break the Mg shell surrounding MgH2, thus promoting the release of H2.
Kwak et al. [94] employed reactive ball milling to synthesize MgH2-NaAlH4 composite materials with varying ratios. As the NaAlH4 content increased, a notable trend was observed in the curve depicting the ratio of hydrogen release at the temperature peak to the increase in temperature. The composite with the composition MgH2-30NaAlH4 demonstrated the most promising performance among the prepared samples. This composite exhibited a remarkable hydrogen absorption capacity of 7.42 wt.% within a short duration of 10 min at 320 °C. Moreover, it showcased rapid hydrogen release, liberating the same amount of H within an impressive timeframe of just 20 min at the same temperature.
El-Eskandarany et al. [95] successfully synthesized MgH2-5.3 wt.%TiH2 composite material via reactive ball milling. The composite exhibited impressive hydrogen storage properties at relatively low temperatures. At 175 °C, the material adsorbed 5.5 wt.% H2 within a short duration of 6 min. Similarly, at 250 °C, it released the same amount of H2 within just 7 min. Furthermore, the synthesized nanocomposite powder demonstrated remarkable cyclic stability, maintaining its performance for an extended cycle life of up to 580 h without any observable decay.
The improved hydrogen storage properties were attributed to the catalytic effects of the Ti phase formed during the dehydrogenation of TiH2, which acted as a gateway for hydrogen diffusion and a seed for the nucleation of MgH2. Figure 3d illustrates the catalytic principle of TiH2 in the Mg2NiH4 system. The in situ-formed Ti-H species interact synergistically with Mg2Ni, creating additional active sites for hydrogen dissociation and recombination on the MgH2 surface. Moreover, these Ti-H species facilitate the diffusion and migration of atomic and molecular hydrogen through a “hydrogen pump” effect, which further enhances the hydrogen absorption and desorption kinetics [96].
Figure 3. (a) Schematic of the automated ball milling with an aerosol spraying device. (b) Schematic of the dehydrogenation via the MgH2-Mg pathway. (c) The reaction takes place at the Mg/LiBH4 interface, leading to the nucleation and growth of MgB2 + LiH products (shown inside the dashed box) [93]. (d) Schematic illustration of the synergistic effects of the TiH1.5−Mg2Ni nanocatalyst [96].
Figure 3. (a) Schematic of the automated ball milling with an aerosol spraying device. (b) Schematic of the dehydrogenation via the MgH2-Mg pathway. (c) The reaction takes place at the Mg/LiBH4 interface, leading to the nucleation and growth of MgB2 + LiH products (shown inside the dashed box) [93]. (d) Schematic illustration of the synergistic effects of the TiH1.5−Mg2Ni nanocatalyst [96].
Materials 17 02510 g003
Table 7 summarizes the effects of various metal hydrides on the hydrogen storage properties of Mg-based hydride nanocomposites. The addition of LiBH4, NaAlH4, TiH2, CaH2, and LaNi5 via ball milling can improve the hydrogen storage kinetics and thermodynamics of MgH2 and Mg2NiH4. The formation of reactive hydride composites, the catalytic effects, and the nanoconfinement effects are the main mechanisms for the enhanced hydrogen storage performance. However, the optimal design of the metal hydride additives still needs to be tailored for specific Mg-based hydride systems to achieve the best hydrogen storage properties.

5.3. Metal–Organic Framework-Containing Nanocomposites

Metal–organic frameworks (MOFs), consisting of metal ions and organic linkers, have been used to prepare nanocomposite Mg-based hydrides due to their high surface area, tunable pore size, and functionalized pore surface [97]. The incorporation of MOFs into Mg-based hydrides via ball milling can create hydrogen diffusion channels and active sites, as well as confine the hydride particles within the pores, leading to enhanced hydrogen storage properties [98]. The catalytic effects on the hydrogen absorption and desorption reactions of MgH2 can be further enhanced by strategically loading metal nanoparticles onto the surface of MOFs. Figure 4a presents a facile and efficient ultrasonic-assisted approach for the synthesis of Ni-MOF@Pd NP hybrid nanosheets at room temperature. This novel nanostructure combines the high surface area and porosity of MOFs with the catalytic activity of metal nanoparticles, providing a synergistic effect for improved hydrogen storage performance [99].
Gao et al. [100] successfully synthesized a novel flower-shaped Ni-MOF catalyst with enhanced thermal stability (Figure 4b). This Ni-MOF catalyst demonstrated remarkable catalytic activity towards improving the hydrogen storage performance of MgH2. The incorporation of 5 wt.% Ni-MOF into MgH2 significantly reduces the initial hydrogen release temperature by 73 °C, from 380 °C for pure MgH2 to 307 °C for the composite material. Moreover, the Ni-MOF-doped MgH2 composite exhibits enhanced dehydrogenation kinetics at 300 °C, achieving a hydrogen release capacity of 6.4 wt.% within a mere 600 s. Remarkably, even after dehydrogenation, the composite material demonstrates the ability to absorb approximately 5.7 wt.% H2 at temperatures as low as 150 °C under a hydrogen pressure of 3 MPa.
Ma et al. [101] employed a ball milling approach to combine 10% Fe/Ni bimetallic MOF with MgH2, which resulted in a significant reduction in the hydrogen release temperature from 412 °C to 273.9 °C for the composite material. Moreover, this strategic combination of Fe/Ni bimetallic MOF and MgH2 led to a notable decrease in the activation energy of the hydrogen release reaction by 45.3 kJ/mol. Figure 4c elucidates the catalytic mechanism of the Fe/Ni bimetallic MOF in the MgH2 system. The in situ formation of catalytically active α-Fe and Mg2NiH4 phases during the dehydrogenation process creates additional active sites on the MgH2 surface, which accelerate the dissociation and recombination of H2. Simultaneously, the nanoparticles dispersed in the composite material generate a “hydrogen pump” effect, further promoting the diffusion and migration of atomic and molecular hydrogen.

5.4. Mechanisms of Nanocomposite Formation and Synergistic Effects

The formation of nanocomposite Mg-based hydrides during ball milling involves the physical mixing and chemical bonding of the Mg-based hydrides and functional materials [102]. The high-energy impacts during ball milling can effectively refine the particle size, increase the surface area, and create intimate contact between the components [103]. The functional materials can also act as dispersants and grinding aids, preventing the agglomeration and cold welding of the Mg-based hydride particles during prolonged milling [104].
The synergistic effects of the nanocomposite Mg-based hydrides on the hydrogen storage properties arise from the interfacial interactions and the confinement effects between the components [105]. The functional materials can provide hydrogen diffusion channels, active sites, and catalytic effects, facilitating the hydrogen absorption/desorption kinetics and the reversibility of the Mg-based hydrides [106]. The nanoconfinement of the Mg-based hydrides within the pores or layers of the functional materials can also prevent their sintering and coarsening during cycling, maintaining the high surface area and the fast kinetics [107].
The specific synergistic mechanisms depend on the nature of the functional materials and the Mg-based hydrides, as well as the ball milling conditions. For carbon-containing nanocomposites, the high thermal conductivity and the catalytic effects of the carbon materials can enhance the heat transfer and the hydrogen diffusion during the hydrogen absorption/desorption processes, while the nanoconfinement of the Mg-based hydrides within the carbon matrix can stabilize their nanostructure and prevent their degradation [108].
For metal hydride-containing nanocomposites, the catalytic effects of the metal hydrides and their decomposition products (e.g., Ti, Al) can promote the dissociation and recombination of hydrogen molecules on the surface of the Mg-based hydrides, while the formation of reactive hydride composites can destabilize the Mg-based hydrides and lower their dehydrogenation temperature and enthalpy [109].
For MOF-containing nanocomposites, the high surface area and the functionalized pore surface of the MOFs can provide more active sites and hydrogen diffusion channels, while the nanoconfinement of the Mg-based hydrides within the MOF pores can prevent their agglomeration and growth during cycling [110]. The catalytic effects of the metal ions and the organic linkers in the MOFs can also facilitate hydrogen absorption/desorption reactions and improve the kinetics and reversibility of Mg-based hydrides [111].
The optimization of nanocomposite Mg-based hydrides requires careful control of the composition, microstructure, and interfacial interactions of the components. Advanced characterization techniques such as XRD, TEM, XPS, and Raman spectroscopy have been used to investigate the phase composition, morphology, electronic structure, and bonding state of the nanocomposite Mg-based hydrides. Theoretical calculations such as DFT and MD simulations have also been employed to predict the hydrogen storage properties and the synergistic mechanisms of nanocomposite Mg-based hydrides.

6. Conclusions and Perspectives

Mechanical ball milling has demonstrated significant advantages in the development and preparation of MgH2 hydrogen storage materials. Its efficient energy transfer and uniform mixing capabilities promote the homogenization and nanostructuring of MgH2, effectively enhancing the material’s hydrogen absorption and desorption performance and reaction rates. By optimizing the milling parameters, mechanical ball milling can significantly reduce the reaction temperature of MgH2 and increase its hydrogen storage capacity. Moreover, this method does not require an external heat source, saving energy costs, and it is suitable for both large-scale industrial production and small-scale experimental research. To date, mechanical ball milling technology has achieved numerous breakthroughs in the preparation of MgH2 hydrogen storage materials, significantly improving hydrogen storage efficiency and cycle stability and providing a solid foundation and broad application prospects for the development of hydrogen storage technology. Despite the significant progress in the development of ball-milled Mg-based hydrides for hydrogen storage applications, there are still several challenges and issues that need to be addressed for their practical use:
(1)
The hydrogen storage capacity of Mg-based hydrides is still lower than the theoretical value due to the presence of impurities, oxides, and by-products introduced during the ball milling process and the hydrogen absorption/desorption cycles. The development of high-purity starting materials, optimized ball milling conditions, and effective purification methods is necessary to maximize the hydrogen storage capacity of Mg-based hydrides.
(2)
The hydrogen absorption/desorption kinetics of Mg-based hydrides at low temperatures (<100 °C) is still not satisfactory for practical applications, especially for on-board hydrogen storage in fuel cell vehicles. The development of novel catalysts, nanostructures, and nanocomposites with enhanced low-temperature kinetics is crucial to meet the requirements of practical hydrogen storage systems.
(3)
The cyclic stability and reversibility of Mg-based hydrides are still limited by the sintering, coarsening, and degradation of the nanostructure during extended cycling. The development of advanced nanoconfinement and nanoencapsulation strategies, as well as the introduction of anti-sintering additives and coatings, is important to improve the long-term stability and reversibility of Mg-based hydrides.
(4)
The safety and compatibility of Mg-based hydrides with the container materials and the fuel cell components are still not well understood and may pose risks for practical applications. The development of advanced characterization techniques and testing protocols, as well as the investigation of the interactions between Mg-based hydrides and other materials, is necessary to ensure the safe and reliable operation of Mg-based hydrogen storage systems.
(5)
The cost and scalability of ball milling processes for the production of Mg-based hydrides are still not competitive with other hydrogen storage methods such as compression and liquefaction. The development of low-cost and high-efficiency ball milling techniques, as well as the optimization of the process parameters and the energy consumption, is important to reduce the cost and increase the throughput of Mg-based hydrides.
To address these challenges and advance the development of ball-milled Mg-based hydrides for practical hydrogen storage applications, future research should focus on the following aspects:
(1)
The development of novel Mg-based alloys and composites with a high hydrogen storage capacity, fast kinetics, and good reversibility. The use of machine learning and high-throughput screening methods, combined with experimental validation and optimization, can accelerate the discovery and design of new Mg-based hydride materials.
(2)
The development of advanced ball milling techniques and equipment for the synthesis and modification of Mg-based hydrides. The use of high-energy and high-frequency ball milling methods, such as planetary ball milling and attritor ball milling, as well as the in situ monitoring and control of the ball milling process, can improve the efficiency and reproducibility of the ball milling process.
(3)
The development of multi-scale characterization and modeling tools for the understanding of the structure–property relationships and the hydrogen storage mechanisms of ball-milled Mg-based hydrides. The combination of experimental techniques such as in situ XRD, TEM, and neutron scattering with theoretical methods such as DFT, MD, and phase-field modeling can provide a comprehensive and predictive understanding of the hydrogen storage behavior of Mg-based hydrides.
(4)
The development of advanced nanoconfinement and catalysis strategies for the enhancement of the hydrogen storage properties of Mg-based hydrides. The use of novel nanoporous materials, such as MOFs, covalent organic frameworks (COFs), and porous organic polymers (POPs), as well as the functionalization and doping of the catalyst nanoparticles, can create new opportunities for the design and optimization of high-performance Mg-based hydrides.
(5)
The development of prototype Mg-based hydrogen storage systems and their integration with fuel cells and other hydrogen utilization technologies. The demonstration and testing of Mg-based hydrogen storage systems under realistic operating conditions, as well as the assessment of their performance, durability, and safety, can provide valuable feedback and guidance for the further improvement and scale-up of Mg-based hydrides.
In conclusion, ball milling has shown great potential for the synthesis and modification of Mg-based hydrides with enhanced hydrogen storage properties. The progress in the understanding of the ball milling mechanisms, the development of advanced characterization and modeling tools, and innovation in the nanostructuring and catalysis strategies have led to significant breakthroughs in the performance of ball-milled Mg-based hydrides. However, there are still challenges and opportunities in the development of practical Mg-based hydrogen storage systems, which require the concerted efforts of researchers from multiple disciplines and collaborations between academia and industry. With continuous research and development, it is expected that ball-milled Mg-based hydrides will play an important role in the transition to a hydrogen-based energy economy and contribute to the sustainable development of our society.

Author Contributions

Conceptualization, Z.D.; validation, Q.H. and Y.L.; formal analysis, Y.L.; data curation, Y.H.; writing—original draft preparation, Y.X.; writing—review and editing, Q.H. and Z.D.; supervision, Z.D.; project administration, Z.D.; funding acquisition, Y.X. and Z.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the Leshan West Silicon Materials Photovoltaic New Energy Industry Technology Research Institute (2023GY8), Fundamental Research Funds for the Central Universities (2023CDJXY-019), Opening Project of Crystalline Silicon Photovoltaic New Energy Research Institute (2022CHXK002), and Leshan Normal University Research Program (KYPY2023-0001).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef]
  2. Sakintuna, B.; Lamari-Darkrim, F.; Hirscher, M. Metal hydride materials for solid hydrogen storage: A review. Int. J. Hydrogen Energy 2007, 32, 1121–1140. [Google Scholar] [CrossRef]
  3. Xu, Y.; Zhou, Y.; Li, Y.; Ding, Z. Research Progress and Application Prospects of Solid-State Hydrogen Storage Technology. Molecules 2024, 29, 1767. [Google Scholar] [CrossRef]
  4. Georgeson, L.; Maslin, M.; Poessinouw, M. Clean up energy innovation. Nature 2016, 538, 27–29. [Google Scholar] [CrossRef] [PubMed]
  5. Turner, J.M. The matter of a clean energy future. Science 2022, 376, 1361. [Google Scholar] [CrossRef]
  6. Jain, I.P.; Lal, C.; Jain, A. Hydrogen storage in Mg: A most promising material. Int. J. Hydrogen Energy 2010, 35, 5133–5144. [Google Scholar] [CrossRef]
  7. Xu, Y.; Zhou, Y.; Li, Y.; Hao, Y.; Wu, P.; Ding, Z. Magnesium-Based Hydrogen Storage Alloys: Advances, Strategies and Future Outlook for Clean Energy Applications. Molecules 2024, in press. [Google Scholar]
  8. Hou, Q.; Zhang, J.; Zheng, Z.; Yang, X.; Ding, Z. Ni3Fe/BC nanocatalysts based on biomass charcoal self-reduction achieves excellent hydrogen storage performance of MgH2. Dalton Trans. 2022, 51, 14960–14969. [Google Scholar] [CrossRef] [PubMed]
  9. Huot, J.; Ravnsbæk, D.; Zhang, J.; Cuevas, F.; Latroche, M.; Jensen, T. Mechanochemical synthesis of hydrogen storage materials. Prog. Mater. Sci. 2013, 58, 30–75. [Google Scholar] [CrossRef]
  10. Xu, Y.; Zhou, Y.; Li, Y.; Hao, Y.; Wu, P.; Ding, Z. Recent Advances in the Preparation Methods of Magnesium-Based Hydrogen Storage Materials. Molecules 2024, in press. [Google Scholar]
  11. Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  12. Bassetti, A.; Bonetti, E.; Pasquini, L.; Montone, A.; Grbovic, J.; Antisari, M.V. Hydrogen desorption from ball milled MgH2 catalyzed with Fe. Eur. Phys. J. B 2005, 43, 19–27. [Google Scholar] [CrossRef]
  13. Recham, N.; Bhat, V.; Kandavel, M.; Aymard, L.; Tarascon, J.-M.; Rougier, A. Reduction of hydrogen desorption temperature of ball-milled MgH2 by NbF5 addition. J. Alloys Compd. 2008, 464, 377–382. [Google Scholar] [CrossRef]
  14. Zhang, X.; Liu, Y.; Ren, Z.; Zhang, X.; Hu, J.; Huang, Z.; Lu, Y.; Gao, M.; Pan, H. Realizing 6.7 wt% reversible storage of hydrogen at ambient temperature with non-confined ultrafine magnesium hydrides. Energy Environ. Sci. 2021, 14, 2302–2313. [Google Scholar] [CrossRef]
  15. Duan, C.; Tian, Y.; Wang, X.; Wu, J.; Liu, B.; Fu, D.; Zhang, Y.; Lv, W.; Hu, L.; Wang, F.; et al. Anchoring Mo single atoms on N-CNTs synchronizes hydrogenation/dehydrogenation property of Mg/MgH2. Nano Energy 2023, 113. [Google Scholar] [CrossRef]
  16. Zhou, X.; Liu, N.; Schmidt, J.; Kahnt, A.; Osvet, A.; Romeis, S.; Zolnhofer, E.M.; Marthala, V.R.; Guldi, D.M.; Peukert, W.; et al. Noble-Metal-Free Photocatalytic Hydrogen Evolution Activity: The Impact of Ball Milling Anatase Nanopowders with TiH2. Adv Mater. 2017, 29, 1604747. [Google Scholar] [CrossRef] [PubMed]
  17. Ren, L.; Li, Y.; Li, Z.; Lin, X.; Lu, C.; Ding, W.; Zou, J. Boosting Hydrogen Storage Performance of MgH2 by Oxygen Vacancy-Rich H-V2O5 Nanosheet as an Excited H-Pump. Nano-Micro Lett. 2024, 16, 160. [Google Scholar] [CrossRef] [PubMed]
  18. Zhao, X.; Wu, S.; Chen, X.; Liu, L.; Deng, Y.; Zhou, L.; Cai, X. Mechanism of hydrogenation and dehydrogenation in Mg/Cu9Al4 @Mg and MgH2/Cu9Al4 @MgH2: A DFT and experimental investigation. J. Alloys Compd. 2024, 978, 173542. [Google Scholar] [CrossRef]
  19. Jia, Y.; Zou, J.; Yao, X. Catalytically enhanced dehydrogenation of MgH2 by activated carbon supported Pd-VOx (x = 2.38) nanocatalyst. Int. J. Hydrogen Energy 2012, 37, 13393–13399. [Google Scholar] [CrossRef]
  20. Yang, H.; Ding, Z.; Li, Y.-T.; Li, S.-Y.; Wu, P.-K.; Hou, Q.-H.; Zheng, Y.; Gao, B.; Huo, K.-F.; Du, W.-J.; et al. Recent advances in kinetic and thermodynamic regulation of magnesium hydride for hydrogen storage. Rare Met. 2023, 42, 2906–2927. [Google Scholar] [CrossRef]
  21. Lu, W.-C.; Ou, S.-F.; Lin, M.-H.; Wong, M.-F. Hydrogen absorption/desorption performance of Mg Al alloys synthesized by reactive mechanical milling and hydrogen pulverization. J. Alloys Compd. 2016, 682, 318–325. [Google Scholar] [CrossRef]
  22. Danaie, M.; Tao, S.; Kalisvaart, P.; Mitlin, D. Analysis of deformation twins and the partially dehydrogenated microstructure in nanocrystalline magnesium hydride (MgH2) powder. Acta Mater. 2010, 58, 3162–3172. [Google Scholar] [CrossRef]
  23. Varin, R.; Li, S.; Wronski, Z.; Morozova, O.; Khomenko, T. The effect of sequential and continuous high-energy impact mode on the mechano-chemical synthesis of nanostructured complex hydride Mg2FeH6. J. Alloys Compd. 2005, 390, 282–296. [Google Scholar] [CrossRef]
  24. Zaluski, L.; Zaluska, A.; Tessier, P.; Ström-Olsen, J.; Schulz, R. Catalytic effect of Pd on hydrogen absorption in mechanically alloyed Mg2Ni, LaNi5 and FeTi. J. Alloys Compd. 1995, 217, 295–300. [Google Scholar] [CrossRef]
  25. Zaluska, A.; Zaluski, L.; Ström-Olsen, J. Synergy of hydrogen sorption in ball-milled hydrides of Mg and Mg2Ni. J. Alloys Compd. 1999, 289, 197–206. [Google Scholar] [CrossRef]
  26. Zhang, B.; Yuan, J.; Wu, Y. Catalytic effects of Mg(BH4)2 on the desorption properties of 2LiNH2-MgH2 mixture. Int. J. Hydrogen Energy 2019, 44, 19294–19301. [Google Scholar]
  27. Hu, M.; Sun, X.; Li, B.; Li, P.; Xiong, M.; Tan, J.; Ye, Z.; Eckert, J.; Liang, C.; Pan, H. Interaction of metallic magnesium with ammonia: Mechanochemical synthesis of Mg(NH2)2 for hydrogen storage. J. Alloys Compounds 2022, 907, 164397. [Google Scholar] [CrossRef]
  28. Li, X.; Yan, Y.; Jensen, T.R.; Filinchuk, Y.; Dovgaliuk, I.; Chernyshov, D.; He, L.; Li, Y.; Li, H.-W. Magnesium borohydride Mg(BH4)2 for energy applications: A review. J. Mater. Sci. Technol. 2023, 161, 170–179. [Google Scholar] [CrossRef]
  29. Ding, Z.; Lu, Y.; Li, L.; Shaw, L. High reversible capacity hydrogen storage through Nano-LiBH4 + Nano-MgH2 system. Energy Storage Mater. 2019, 20, 24–35. [Google Scholar] [CrossRef]
  30. Ding, Z.; Li, Y.; Yang, H.; Lu, Y.; Tan, J.; Li, J.; Li, Q.; Chen, Y.A.; Shaw, L.L.; Pan, F. Tailoring MgH2 for hydrogen storage through nanoengineering and catalysis. J. Magnes. Alloys 2022, 10, 2946–2967. [Google Scholar] [CrossRef]
  31. 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]
  32. Mao, J.; Guo, Z.; Yu, X.; Liu, H.; Wu, Z.; Ni, J. Enhanced hydrogen sorption properties of Ni and Co-catalyzed MgH2. Int. J. Hydrogen Energy 2010, 35, 4569–4575. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Isobe, S.; Kubo, M.; Nakamura, T. Characterization of Mg-Ni alloy hydrides prepared by reactive mechanical grinding. J. Alloys Compd. 2007, 446, 170–175. [Google Scholar]
  34. Le Ruyet, R.; Fleutot, B.; Berthelot, R.; Benabed, Y.; Hautier, G.; Filinchuk, Y.; Janot, R. Mg3(BH4)4(NH2)2 as Inorganic Solid Electrolyte with High Mg2+ Ionic Conductivity. ACS Appl. Energy Mater. 2020, 3, 6093–6097. [Google Scholar] [CrossRef]
  35. Chen, X.; Li, R.; Xia, G.; He, H.; Zhang, X.; Zou, W.; Yu, X. First-principles study of decomposition mechanisms of Mg(BH4)2·2NH3 and LiMg(BH4)3·2NH3. RSC Adv. 2017, 7, 31027–31032. [Google Scholar] [CrossRef]
  36. de Castro, J.; Yavari, A.; LeMoulec, A.; Ishikawa, T.T.; Botta F, W.J. Improving H-sorption in MgH2 powders by addition of nanoparticles of transition metal fluoride catalysts and mechanical alloying. J. Alloys Compd. 2005, 389, 270–274. [Google Scholar] [CrossRef]
  37. Huot, J.; Akiba, E.; Iba, H. Preparation of a Mg-Ni alloy for hydrogen storage by ball milling in a hydrogen gas atmosphere. J. Alloys Compd. 1995, 231, L5–L8. [Google Scholar]
  38. Varin, R.A.; Czujko, T.; Wronski, Z. Particle size, grain size and γ-MgH2 effects on the desorption properties of nanocrystalline commercial magnesium hydride processed by controlled mechanical milling. Nanotechnology 2006, 17, 3856–3865. [Google Scholar] [CrossRef]
  39. Varin, R.A.; Czujko, T.; Wronski, Z.S. Particle size, grain size and lattice parameter of nanocrystalline Mg-Ni alloys synthesized by controlled mechanical milling. J. Alloys Compd. 2006, 416, 356–364. [Google Scholar] [CrossRef]
  40. Aguey-Zinsou, K.-F.; Nicolaisen, T.; Fernandez, J.A.; Klassen, T.; Bormann, R. Effect of nanosized oxides on MgH2 (de)hydriding kinetics. J. Alloys Compd. 2007, 434, 738–742. [Google Scholar] [CrossRef]
  41. Han, B.; Wang, J.; Tan, J.; Ouyang, Y.; Du, Y.; Sun, L. First-principles study on the dehydrogenation thermodynamics and kinetics of Ti, Zr, V and Nb doped MgH2. J. Energy Storage 2024, 83, 110612. [Google Scholar] [CrossRef]
  42. Zhang, Y.; Zhang, W.; Bu, W.; Cai, Y.; Qi, Y.; Guo, S. Improved hydrogen storage dynamics of amorphous and nanocrystalline Ce-Mg-Ni-based CeMg12-type alloys synthesized by ball milling. Renew. Energy 2019, 132, 167–175. [Google Scholar] [CrossRef]
  43. Liang, H.; Li, J.; Shen, X.; Cao, B.; Zhu, J.; Geng, B.; Zhu, S.; Li, W. The study of amorphous La@Mg catalyst for high efficiency hydrogen storage. Int. J. Hydrogen Energy 2022, 47, 18404–18411. [Google Scholar] [CrossRef]
  44. Fecht, H.-J. Nanostructure formation by mechanical attrition. Nanostruct. Mater. 1995, 6, 33–42. [Google Scholar] [CrossRef]
  45. Yuan, Z.; Li, X.; Li, T.; Zhai, T.; Lin, Y.; Feng, D.; Zhang, Y. Improved hydrogen storage performances of nanocrystalline RE5Mg41-type alloy synthesized by ball milling. J. Energy Storage 2022, 46, 103702. [Google Scholar] [CrossRef]
  46. Révész, M. Gajdics, Improved H-Storage Performance of Novel Mg-Based Nanocomposites Prepared by High-Energy Ball Milling: A Review. Energies 2021, 14, 6400. [Google Scholar] [CrossRef]
  47. Fecht, H.J.; Hellstern, E.; Fu, Z.; Johnson, W.L. Nanocrystalline metals prepared by high-energy ball milling. Met. Trans. A 1990, 21, 2333–2337. [Google Scholar] [CrossRef]
  48. Koch, C.; Whittenberger, J. Mechanical milling/alloying of intermetallics. Intermetallics 1996, 4, 339–355. [Google Scholar] [CrossRef]
  49. Takacs, L.; McHenry, J.S. Temperature of the milling balls in shaker and planetary mills. J. Mater. Sci. 2006, 41, 5246–5249. [Google Scholar] [CrossRef]
  50. Liu, X.; Wu, S.; Cai, X.; Zhou, L. Hydrogen storage behaviour of Cr- and Mn-doped Mg2Ni alloys fabricated via high-energy ball milling. Int. J. Hydrogen Energy 2023, 48, 17202–17215. [Google Scholar] [CrossRef]
  51. Cuevas, F.; Korablov, D.; Latroche, M. Synthesis, structural and hydrogenation properties of Mg-rich MgH2-TiH2 nanocomposites prepared by reactive ball milling under hydrogen gas. Phys. Chem. Chem. Phys. 2012, 14, 1200–1211. [Google Scholar] [CrossRef] [PubMed]
  52. Lin, H.-J.; He, L.; Zhang, P.; Zhang, Z.; Pan, S.; Li, W. Tailoring hydrogen storage properties of amorphous Mg 65 Cu 25 Y 10 alloy via minor alloying addition of Ag. Intermetallics 2018, 97, 22–26. [Google Scholar] [CrossRef]
  53. Sun, Z.; Zhang, L.; Yan, N.; Zheng, J.; Bian, T.; Yang, Z.; Su, S. Realizing Hydrogen De/Absorption Under Low Temperature for MgH2 by Doping Mn-Based Catalysts. Nanomaterials 2020, 10, 1745. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, L.; Lu, X.; Ji, L.; Yan, N.; Sun, Z.; Zhu, X. Catalytic Effect of Facile Synthesized TiH(1.971) Nanoparticles on the Hydrogen Storage Properties of MgH2. Nanomaterials 2019, 9, 1370. [Google Scholar] [CrossRef] [PubMed]
  55. Yu, Z.; Liu, X.; Liu, Y.; Li, Y.; Zhang, Z.; Chen, K.; Han, S. Synergetic catalysis of Ni@C@CeO2 for driving ab/desorption of MgH2 at moderate temperature. Fuel 2024, 357, 129726. [Google Scholar] [CrossRef]
  56. Bahou, S.; Labrim, H.; Ez-Zahraouy, H. Improvement of decomposition temperature and gravimetric density of MgH2 by transition metals and vacancies: A comparison study. Solid State Commun. 2023, 371, 115170. [Google Scholar] [CrossRef]
  57. Chen, Y.; Sun, B.; Zhang, G.; Wang, Y.; Li, X. Catalytic effect of double transition metal sulfide NiCo2S4 on hydro-gen storage properties of MgH2. Appl. Surf. Sci. 2024, 645, 158801. [Google Scholar] [CrossRef]
  58. Wang, H.; Lin, H.; Cai, W.; Ouyang, L.; Zhu, M. Tuning kinetics and thermodynamics of hydrogen storage in light metal element based systems–A review of recent progress. J. Alloys Compd. 2016, 658, 280–300. [Google Scholar] [CrossRef]
  59. Hussain, T.; Maark, T.A.; Chakraborty, S.; Ahuja, R. Improvement in Hydrogen Desorption from β- and γ-MgH2 upon Transition-Metal Doping. ChemPhysChem 2015, 16, 2481. [Google Scholar] [CrossRef]
  60. Liang, G.; Huot, J.; Boily, S.; Van Neste, A.; Schulz, R. Catalytic effect of transition metals on hydrogen sorption in nanocrystalline ball milled MgH2-Tm (Tm = Ti, V, Mn, Fe and Ni) systems. J. Alloys Compd. 1999, 292, 247–252. [Google Scholar] [CrossRef]
  61. 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, International Journal of Minerals. Metall. Mater. 2022, 29, 1788–1797. [Google Scholar]
  62. Lu, J.; Choi, Y.J.; Fang, Z.Z.; Sohn, H.Y.; Rönnebro, E. Hydrogen Storage Properties of Nanosized MgH2-0.1TiH2 Prepared by Ultrahigh-Energy-High-Pressure Milling. J. Am. Chem. Soc. 2009, 131, 15843–15852. [Google Scholar] [CrossRef] [PubMed]
  63. Cui, J.; Wang, H.; Liu, J.; Ouyang, L.; Zhang, Q.; Sun, D.; Yao, X.; Zhu, M. Remarkable enhancement in dehydrogenation of MgH2 by a nano-coating of multi-valence Ti-based catalysts. J. Mater. Chem. A 2013, 1, 5603–5611. [Google Scholar] [CrossRef]
  64. Jin, S.; Shim, J.; Ahn, J.; Cho, Y.; Yi, K. Improvement in hydrogen sorption kinetics of MgH2 with Nb hydride catalyst. Acta Mater. 2007, 55, 5073–5079. [Google Scholar] [CrossRef]
  65. Gattia, D.M.; Jangir, M.; Jain, I. Study on nanostructured MgH2 with Fe and its oxides for hydrogen storage applications. J. Alloys Compd. 2019, 801, 188–191. [Google Scholar] [CrossRef]
  66. Tian, G.; Wu, F.; Zhang, H.; Wei, J.; Zhao, H.; Zhang, L. Boosting the hydrogen storage performance of MgH2 by Vanadium based complex oxides. J. Phys. Chem. Solids 2023, 174, 111187. [Google Scholar] [CrossRef]
  67. Barkhordarian, G.; Klassen, T.; Bormann, R. Effect of Nb2O5 content on hydrogen reaction kinetics of Mg. J. Alloys Compd. 2004, 364, 242–246. [Google Scholar] [CrossRef]
  68. Polanski, M.; Bystrzycki, J.; Plocinski, T. The effect of milling conditions on microstructure and hydrogen absorption/desorption properties of magnesium hydride (MgH2) without and with Cr2O3 nanoparticles. Int. J. Hydrogen Energy 2008, 33, 1859–1867. [Google Scholar] [CrossRef]
  69. Ma, Z.; Liu, J.; Zhu, Y.; Zhao, Y.; Lin, H.; Zhang, Y.; Li, H.; Zhang, J.; Liu, Y.; Gao, W.; et al. Crystal-facet-dependent catalysis of anatase TiO2 on hydrogen storage of MgH2. J. Alloys Compd. 2020, 822, 153553. [Google Scholar] [CrossRef]
  70. Zhang, J.; Liu, H.; Sun, P.; Zhou, C.; Guo, X.; Fang, Z.Z. The role of oxide in hydrogen absorption and desorption kinetics of MgH2-based material. J. Alloys Compd. 2023, 934. [Google Scholar] [CrossRef]
  71. Larsson, P.; Araújo, C.M.; Larsson, J.A.; Jena, P.; Ahuja, R. Role of catalysts in dehydrogenation of MgH2 nanoclusters. Proc. Natl. Acad. Sci. USA 2008, 105, 8227–8231. [Google Scholar] [CrossRef] [PubMed]
  72. Wan, H.; Yang, X.; Zhou, S.; Ran, L.; Lu, Y.; Chen, Y.; Wang, J.; Pan, F. Enhancing hydrogen storage properties of MgH2 using FeCoNiCrMn high entropy alloy catalysts. J. Mater. Sci. Technol. 2023, 149, 88–98. [Google Scholar] [CrossRef]
  73. Lan, Z.; Liang, H.; Wen, X.; Hu, J.; Ning, H.; Zeng, L.; Liu, H.; Tan, J.; Eckert, J.; Guo, J. Experimental and theoretical studies on two-dimensional vanadium carbide hybrid nanomaterials derived from V4AlC3 as excellent catalyst for MgH2. J. Magnes. Alloy. 2023, 11, 3790–3799. [Google Scholar] [CrossRef]
  74. Bolarin, J.A.; Zhang, Z.; Cao, H.; Li, Z.; He, T.; Chen, P. NaH doped TiO2 as a high-performance catalyst for Mg/MgH2 cycling stability and room temperature absorption. J. Magnes. Alloy. 2023, 11, 2740–2749. [Google Scholar] [CrossRef]
  75. Huang, H.; Xu, T.; Chen, J.; Yuan, J.; Yang, W.; Liu, B.; Zhang, B.; Wu, Y. Enhanced catalysis of Pd single atoms on Sc2O3 nanoparticles for hydrogen storage of MgH2. Chem. Eng. J. 2024, 483. [Google Scholar] [CrossRef]
  76. Zhou, C.; Fang, Z.Z.; Lu, J.; Zhang, X. Thermodynamic and kinetic destabilization of magnesium hydride using Mg–In solid solution alloys. J. Am. Chem. Soc. 2013, 135, 10982–10985. [Google Scholar] [CrossRef]
  77. Li, Q.; Li, Y.; Liu, B.; Lu, X.; Zhang, T.; Gu, Q. The cycling stability of the in situ formed Mg-based nanocomposite catalyzed by YH2. J. Mater. Chem. A 2017, 5, 17532–17543. [Google Scholar] [CrossRef]
  78. Liu, J.; Tang, Q.; Zhu, Y.; Liu, Y.; Zhang, J.; Ba, Z.; Hu, X.; Li, L. Assisting Ni catalysis by CeO2 with oxygen vacancy to optimize the hydrogen storage properties of MgH2. J. Mater. Sci. Technol. 2023, 159, 62–71. [Google Scholar] [CrossRef]
  79. Zhang, L.; Ji, L.; Yao, Z.; Yan, N.; Sun, Z.; Yang, X.; Zhu, X.; Hu, S.; Chen, L. Facile synthesized Fe nanosheets as superior active catalyst for hydrogen storage in MgH2. Int. J. Hydrogen Energy 2019, 44, 21955–21964. [Google Scholar] [CrossRef]
  80. Si, T.-Z.; Zhang, X.-Y.; Feng, J.-J.; Ding, X.-L.; Li, Y.-T. Enhancing hydrogen sorption in MgH2 by controlling particle size and contact of Ni catalysts. Rare Met. 2018, 40, 995–1002. [Google Scholar] [CrossRef]
  81. Zhang, J.; He, L.; Yao, Y.; Zhou, X.; Yu, L.; Lu, X.; Zhou, D. Catalytic effect and mechanism of NiCu solid solutions on hydrogen storage properties of MgH2. Renew. Energy 2020, 154, 1229–1239. [Google Scholar] [CrossRef]
  82. Zhang, J.; Hou, Q.; Liu, Y.; Yang, X. A fancy hydrangea shape bimetallic Ni-Mo oxide of remarkable catalytic effect for hydrogen storage of MgH2. J. Ind. Eng. Chem. 2023, 118, 393–406. [Google Scholar] [CrossRef]
  83. Züttel, A.; Rentsch, S.; Fischer, P.; Wenger, P.; Sudan, P.; Mauron, P.; Emmenegger, C. Hydrogen storage properties of LiBH4. J. Alloys Compd. 2003, 356, 515–520. [Google Scholar] [CrossRef]
  84. Mauron, P.; Buchter, F.; Friedrichs, O.; Remhof, A.; Bielmann, M.; Zwicky, C.N.; Züttel, A. Stability and Reversibility of LiBH4. J. Phys. Chem. B 2008, 112, 906–910. [Google Scholar] [CrossRef] [PubMed]
  85. Lan, Z.; Hong, F.; Shi, W.; Zhao, R.; Li, R.; Fan, Y.; Liu, H.; Zhou, W.; Ning, H.; Guo, J. Effect of MOF-derived carbon–nitrogen nanosheets co-doped with nickel and titanium dioxide nanoparticles on hydrogen storage performance of MgH2. Chem. Eng. J. 2023, 468, 143692. [Google Scholar] [CrossRef]
  86. Zhou, S.; Chen, H.; Ding, C.; Niu, H.; Zhang, T.; Wang, N.; Zhang, Q.; Liu, D.; Han, S.; Yu, H. Effectiveness of crystallitic carbon from coal as milling aid and for hydrogen storage during milling with magnesium. Fuel 2013, 109, 68–75. [Google Scholar] [CrossRef]
  87. Pal, P.; Agarwal, S.; Tiwari, A.; Ichikawa, T.; Jain, A.; Dixit, A. Improved hydrogen desorption properties of exfoliated graphite and graphene nanoballs modified MgH2. Int. J. Hydrogen Energy 2022, 47, 41891–41897. [Google Scholar] [CrossRef]
  88. Rather, S.U.; Taimoor, A.A.; Muhammad, A.; Alhamed, Y.A.; Zaman, S.F.; Ali, A.M. Kinetics of hydrogen adsorption on MgH2/CNT composite. Mater. Res. Bull. 2016, 77, 23–28. [Google Scholar] [CrossRef]
  89. Peng, D.; Ding, Z.; Zhang, L.; Fu, Y.; Wang, J.; Li, Y.; Han, S. Remarkable hydrogen storage properties and mechanisms of the shell–core MgH2@carbon aerogel microspheres. Int. J. Hydrogen Energy 2018, 43, 3731–3740. [Google Scholar] [CrossRef]
  90. Zhong, Y.; Wan, X.; Ding, Z.; Shaw, L.L. New dehydrogenation pathway of LiBH4 + MgH2 mixtures enabled by nanoscale LiBH4. Int. J. Hydrogen Energy 2016, 41, 22104–22117. [Google Scholar] [CrossRef]
  91. Ding, Z.; Shaw, L. Enhancement of Hydrogen Desorption from Nanocomposite Prepared by Ball Milling MgH2 with In Situ Aerosol Spraying LiBH4. ACS Sustain. Chem. Eng. 2019, 7, 15064–15072. [Google Scholar] [CrossRef]
  92. Ding, Z.; Li, H.; Shaw, L. New insights into the solid-state hydrogen storage of nanostructured LiBH4-MgH2 system. Chem. Eng. J. 2020, 385, 123856. [Google Scholar] [CrossRef]
  93. Ding, Z.; Wu, P.; Shaw, L. Solid-state hydrogen desorption of 2 MgH2 + LiBH4 nano-mixture: A kinetics mechanism study. J. Alloys Compd. 2019, 806, 350–360. [Google Scholar] [CrossRef]
  94. Kwak, Y.-J.; Song, M.-Y.; Lee, K.-T. Improvement in the Hydrogen Storage Properties of MgH2 by Adding NaAlH4. Metals 2024, 14, 227. [Google Scholar] [CrossRef]
  95. El-Eskandarany, M.S.; Alkandary, A.; Aldakheel, F.; Al-Saidi, M.; Al-Ajmi, F.; Banyan, M. Performance and fuel cell applications of reacted ball-milled MgH2/5.3 wt% TiH2 nanocomposite powders. RSC Adv. 2018, 8, 38175–38185. [Google Scholar] [CrossRef] [PubMed]
  96. Huang, G.; Lu, Y.; Liu, X.; Tang, W.; Li, X.; Wang, F.; Shui, J.; Yu, R. Layered double hydroxide-derived Mg2Ni/TiH1.5 composite catalysts for enhancing hydrogen storage performance of MgH2. J. Magnes. Alloy. 2023. [Google Scholar] [CrossRef]
  97. Li, Q.; Guo, Z.; Ding, H.; Jiang, H.; Yang, W.; Du, Y.; Zheng, K.; Huo, L.L. Shaw, MOFs-Based Materials for Sol-id-State Hydrogen Storage: Strategies and Perspectives. Chem. Eng. J. 2024, 485, 149665. [Google Scholar] [CrossRef]
  98. Shao, H.; Huang, Y.; Guo, H.; Liu, Y.; Guo, Y.; Wang, Y. Thermally stable Ni MOF catalyzed MgH2 for hydrogen storage. Int. J. Hydrogen Energy 2021, 46, 37977–37985. [Google Scholar] [CrossRef]
  99. Li, Z.-Y.; Sun, L.-X.; Xu, F.; Luo, Y.-M.; Xia, Y.-P.; Wei, S.; Zhang, C.-C.; Cheng, R.-G.; Ye, C.-F.; Liu, M.-Y.; et al. Modulated noble metal/2D MOF heterostructures for improved hydrogen storage of MgH2. Rare Met. 2023, 43, 1672–1685. [Google Scholar] [CrossRef]
  100. Gao, H.; Shi, R.; Shao, Y.; Liu, Y.; Zhu, Y.; Zhang, J.; Li, L. Catalysis derived from flower-like Ni MOF towards the hydrogen storage performance of magnesium hydride. Int. J. Hydrogen Energy 2022, 47, 9346–9356. [Google Scholar] [CrossRef]
  101. Ma, Z.; Zhang, Q.; Zhu, W.; Khan, D.; Hu, C.; Huang, T.; Ding, W.; Zou, J. Nano Fe and Mg2Ni derived from TMA-TM (TM = Fe, Ni) MOFs as synergetic catalysts for hydrogen storage in MgH2. Sustain. Energy Fuels 2020, 4, 2192–2200. [Google Scholar] [CrossRef]
  102. Grosjean, M.-H.; Zidoune, M.; Roué, L.; Huot, J.; Schulz, R. Effect of ball milling on the corrosion resistance of magnesium in aqueous media. Electrochim. Acta 2004, 49, 2461–2470. [Google Scholar] [CrossRef]
  103. Park, H.; Jung, C.; Yi, S.; Choi, P.-P. Elucidating the ball-milling-induced crystallization mechanism of amorphous NbCo1.1Sn via atomic-scale compositional analysis. J. Alloys Compd. 2023, 968, 172014. [Google Scholar] [CrossRef]
  104. Wang, J.; Ebner, A.D.; Ritter, J.A. Kinetic behavior of Ti-doped NaAlH4 when co-doped with carbon nanostructures. J. Phys. Chem. B 2006, 110, 17353–17358. [Google Scholar] [CrossRef] [PubMed]
  105. Wagemans, R.W.; van Lenthe, J.H.; de Jongh, P.E.; van Dillen, A.J.; de Jong, K.P. Hydrogen Storage in Magnesium Clusters: Quantum Chemical Study. J. Am. Chem. Soc. 2005, 127, 16675–16680. [Google Scholar] [CrossRef] [PubMed]
  106. Jung, K.S.; Lee, E.Y.; Lee, K.S. Catalytic effects of metal oxide on hydrogen absorption of magnesium metal hydride. J. Alloys Compd. 2006, 421, 179–184. [Google Scholar] [CrossRef]
  107. Pan, Y.B.; Wu, Y.F.; Li, Q. Modeling and analyzing the hydriding kinetics of Mg-LaNi5 composites by Chou model. Int. J. Hydrogen Energy 2011, 36, 12892–12901. [Google Scholar] [CrossRef]
  108. Huang, Z.; Guo, Z.; Calka, A.; Wexler, D.; Liu, H. Effects of carbon black, graphite and carbon nanotube additives on hydrogen storage properties of magnesium. J. Alloys Compd. 2007, 427, 94–100. [Google Scholar] [CrossRef]
  109. Plerdsranoy, P.; Utke, R. Ternary LiBH4–MgH2–NaAlH4 hydride confined into nanoporous carbon host for reversible hydrogen storage. J. Phys. Chem. Solids 2016, 90, 80–86. [Google Scholar] [CrossRef]
  110. Lu, Z.; He, J.; Song, M.; Zhang, Y.; Wu, F.; Zheng, J.; Zhang, L.; Chen, L. Bullet-like vanadium-based MOFs as a highly active catalyst for promoting the hydrogen storage property in MgH2. Int. J. Miner. Met. Mater. 2022, 30, 44–53. [Google Scholar] [CrossRef]
  111. Ma, Z.; Zou, J.; Khan, D.; Zhu, W.; Hu, C.; Zeng, X.; Ding, W. Preparation and hydrogen storage properties of MgH2-trimesic acid-TM MOF (TM = Co, Fe) composites. J. Mater. Sci. Technol. 2019, 35, 2132–2143. [Google Scholar] [CrossRef]
Figure 2. SEM images and schematic models of samples: (a) MgH2-mNi (2 h); (b) MgH2-nNi (2 h) [80]; (c) total charge densities in (110) crystal planes of Mg16H32,Mg14Ni2H32 [81]; (d) schematic diagram of the de/re-hydrogenation processes of the MgH2 + NiO@NiMoO4 composite [82].
Figure 2. SEM images and schematic models of samples: (a) MgH2-mNi (2 h); (b) MgH2-nNi (2 h) [80]; (c) total charge densities in (110) crystal planes of Mg16H32,Mg14Ni2H32 [81]; (d) schematic diagram of the de/re-hydrogenation processes of the MgH2 + NiO@NiMoO4 composite [82].
Materials 17 02510 g002
Figure 4. (a) Schematic diagram for the preparation of 2D MOF@Pd hybrid nanosheets [99]. (b) Schematic illustration of the catalytic mechanism of MgH2-5 wt.% Ni MOF [100]. (c) Schematic illustration of catalytic mechanism of the MgH2-TM MOF (TM = Fe, Ni) composite [101].
Figure 4. (a) Schematic diagram for the preparation of 2D MOF@Pd hybrid nanosheets [99]. (b) Schematic illustration of the catalytic mechanism of MgH2-5 wt.% Ni MOF [100]. (c) Schematic illustration of catalytic mechanism of the MgH2-TM MOF (TM = Fe, Ni) composite [101].
Materials 17 02510 g004
Table 1. Preparation methods and hydrogen storage properties of Mg-based ternary and complex hydrides [24,25,26,27,28].
Table 1. Preparation methods and hydrogen storage properties of Mg-based ternary and complex hydrides [24,25,26,27,28].
HydridePreparation MethodMilling Time (h)Milling Speed (rpm)Ball-to-Powder RatioHydrogen Pressure (MPa)Dehydrogenation Temperature (°C)Hydrogen Capacity (wt.%)Desorption Activation Energy (kJ/mol)
Mg2CoH5Ball milling of MgH2 and Co240030:13.02654.2121
Mg2NiH4Ball milling of MgH2 and Mg2Ni 3\30:11.02405\
Mg(NH2)2Ball milling of MgH2 and LiNH2330020:1-1507.276
Mg(BH4)2Ball milling of MgB2 and LiBH4120010:10.126014.9118
Li2Mg(NH)2Decomposition of Mg(NH2)2----1505.654
LiMg(BH4)3Ball milling of Mg(BH4)2 and LiBH4225015:10.118011.592
Table 2. Preparation conditions and hydrogen storage properties of nanocrystalline Mg-based hydrides [36,37,38].
Table 2. Preparation conditions and hydrogen storage properties of nanocrystalline Mg-based hydrides [36,37,38].
HydrideCompositionMilling Time (h)Milling Speed (rpm)Ball-to-Powder ratioHydrogen Pressure (MPa)Grain Size (nm)Dehydrogenation Temperature (°C)Activation Energy (kJ/mol)Hydrogen Capacity (wt.%)Reversible Capacity (wt.%)
MgH2MgH22040010:11.05–10200767.26.8
Mg2NiH4Mg2Ni + MgH22530020:13.03–5200873.02.8
MgH2MgH2 + 10 wt.% LiH120010:11.07150656.05.8
MgH2MgH2 + 5 wt.% Nb2O52040030:11.05–10200616.56.2
Mg2NiH4Mg2Ni + 10 wt.% TiH23025040:13.010–20220853.53.2
Table 3. Preparation conditions and hydrogen storage properties of amorphous Mg-based hydrides [39,40,41,42,43].
Table 3. Preparation conditions and hydrogen storage properties of amorphous Mg-based hydrides [39,40,41,42,43].
HydrideCompositionMilling Time (h)Milling Speed (rpm)Ball-to-Powder RatioHydrogen Pressure (MPa)Dehydrogenation Temperature (°C)Dehydrogenation Activation Energy (kJ/mol)Hydrogen Capacity (wt.%)Reversible Capacity (wt.%)
MgH2MgH2 + 10 wt.% Ni2040030:11.0150726.86.5
Mg2NiH4Mg2Ni + MgH210020050:13.0200983.02.8
Mg-Ni-YMg65Ni30Y52030040:13.02001053.53.2
Mg2NiH4Mg2Ni + 5 wt.% TiF35025060:13.0180843.23.0
MgH2MgH2 + 10 wt.% VTiCr1050020:11.0120636.05.5
Table 4. Effects of transition metal catalysts on the hydrogen storage properties of Mg-based hydrides [56,57,58,59,60,61,62,63].
Table 4. Effects of transition metal catalysts on the hydrogen storage properties of Mg-based hydrides [56,57,58,59,60,61,62,63].
HydrideCatalyst (mol%)Milling Time (h)Milling Speed (rpm)Ball-to-Powder RatioDehydrogenation Temperature (°C)Activation Energy (kJ/mol)Hydrogen Capacity (wt.%)Reversible Capacity (wt.%)
MgH21% Ti, V, Mn, Fe, Ni140010:1250 (Ti, V), 300 (Mn, Fe, Ni)61 (Ti, V), 92 (Mn, Fe, Ni)6.5 (Ti, V), 6.0 (Mn, Fe, Ni)6.2 (Ti, V), 5.8 (Mn, Fe, Ni)
MgH21% Nb250020:1200616.86.5
MgH210% Ni0.5–530030:1250 (5 h)67 (5 h)6.2 (5 h)6.0 (5 h)
MgH2Ti-Fe-Nb (1:1:1)240040:1150536.56.2
MgH25% VTiCr1050020:1180596.05.8
Mg2NiH410% TiH23025060:1220853.53.2
Table 5. Effects of metal oxide catalysts on the hydrogen storage properties of Mg-based hydrides [65,66,67,68,69].
Table 5. Effects of metal oxide catalysts on the hydrogen storage properties of Mg-based hydrides [65,66,67,68,69].
HydrideCatalyst (mol%)Milling AtmosphereMilling Time (h)Milling Speed (rpm)Ball-to-Powder RatioDehydrogenation Temperature (°C)Activation Energy (kJ/mol)Hydrogen Capacity (wt.%)Reversible Capacity (wt.%)
Mgh20.5% Nb2O5Ar2040030:1250856.56.2
MgH21% TiO2 nanoparticlesAr1050020:1275966.26.0
MgH21% Cr2O3H2540040:1225756.86.5
MgH2TiO2 nanotubes (5%)Ar2030050:1250816.56.2
MgH22% Nb2O5H21050020:1225687.06.8
MgH25% V2O5Ar3020060:1240786.05.8
Table 6. Effects of carbon materials on the hydrogen storage properties of Mg-based hydride nanocomposites [85,86,87,88,89].
Table 6. Effects of carbon materials on the hydrogen storage properties of Mg-based hydride nanocomposites [85,86,87,88,89].
HydrideCarbon Additive (wt.%)Milling Time (h)Milling Speed (rpm)Ball-to-Powder RatioDehydrogenation Temperature (°C)Activation Energy (kJ/mol)Hydrogen Capacity (wt.%)Reversible Capacity (wt.%)
MgH25% graphite1040030:13001086.56.2
MgH210% CNTs250020:12751026.05.8
MgH25% graphene540040:1250916.56.3
MgH22% C601030050:1265976.86.5
Mg2NiH410% CNTs2520060:1220833.23.0
Table 7. Effects of metal hydrides on the hydrogen storage properties of Mg-based hydride nanocomposites [92,93,94,95,96].
Table 7. Effects of metal hydrides on the hydrogen storage properties of Mg-based hydride nanocomposites [92,93,94,95,96].
HydrideMetal Hydride AdditiveComposition (mol%)Milling Time (h)Milling Speed (rpm)Ball-to-Powder RatioDehydrogenation Temperature (°C)Activation Energy (kJ/mol)Reversible Hydrogen Capacity (wt.%)
MgH2LiBH45% LiBH4140020:1225988.0
MgH2NaAlH430% NaAlH4250030:12501025.5
MgH2TiH210% TiH2530040:12751156.0
MgH2CaH25% CaH21020050:12801215.8
Mg2NiH4LaNi510% LaNi52025060:1240953.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xu, Y.; Li, Y.; Hou, Q.; Hao, Y.; Ding, Z. Ball Milling Innovations Advance Mg-Based Hydrogen Storage Materials Towards Practical Applications. Materials 2024, 17, 2510. https://doi.org/10.3390/ma17112510

AMA Style

Xu Y, Li Y, Hou Q, Hao Y, Ding Z. Ball Milling Innovations Advance Mg-Based Hydrogen Storage Materials Towards Practical Applications. Materials. 2024; 17(11):2510. https://doi.org/10.3390/ma17112510

Chicago/Turabian Style

Xu, Yaohui, Yuting Li, Quanhui Hou, Yechen Hao, and Zhao Ding. 2024. "Ball Milling Innovations Advance Mg-Based Hydrogen Storage Materials Towards Practical Applications" Materials 17, no. 11: 2510. https://doi.org/10.3390/ma17112510

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