Next Article in Journal
Thermohydraulic Performance and Entropy Generation of a Triple-Pass Solar Air Heater with Three Inlets
Previous Article in Journal
Analysis of Reservoir Fluid Migration in the Process of CO2 Sequestration in a Partially Depleted Gas Reservoir
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Improved H-Storage Performance of Novel Mg-Based Nanocomposites Prepared by High-Energy Ball Milling: A Review

Department of Materials Physics, Eötvös University, P.O. Box 32, H-1518 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Energies 2021, 14(19), 6400; https://doi.org/10.3390/en14196400
Submission received: 14 September 2021 / Revised: 1 October 2021 / Accepted: 4 October 2021 / Published: 7 October 2021

Abstract

:
Hydrogen storage in magnesium-based composites has been an outstanding research area including a remarkable improvement of the H-sorption properties of this system in the last 5 years. Numerous additives of various morphologies have been applied with great success to accelerate the absorption/desorption reactions. Different combinations of catalysts and preparation conditions have also been explored to synthesize better hydrogen storing materials. At the same time, ball milling is still commonly and effectively applied for the fabrication of Mg-based alloys and composites in order to reduce the grain size to nanometric dimensions and to disperse the catalyst particles over the surface of the host material. In this review, we present the very recent progress, from 2016 to 2021, on catalyzing the hydrogen sorption of Mg-based materials by ball milling. The various catalyzing routes enhancing the hydrogenation performance, including in situ formation of catalysts and synergistic improvement achieved by using multiple additives, will also be summarized. At the end of this work, some thoughts on the prospects for future research will be highlighted.

1. Introduction

According to the Key World Energy Statistics 2020 report of the International Energy Agency, the worldwide total energy consumption has increased by 50% in the last 20 years [1]. The majority of the consumed energy is produced from non-renewable sources, mainly oil and natural gas. The associated pollution puts a significant burden on the environment as we see more and more evidence of climate change, since CO2 emissions increased by 50% in the last 20 years. To counterbalance this issue, renewable energy sources have gained significant attention in the last few decades. For example, the energy production using renewables and biofuel has continuously increased in the EU since the early 2000s, and in 2018, these two sources accounted for 34.2% of the total energy production [2]. With the significant increment in the renewable production capacity, the inherent difficulties of these sources (location dependency, non-continuous supply) become a more pressing issue. Hence, energy storage technologies have to be developed to cope with future demands.
One of the promising prospects is the utilization of hydrogen as an energy vector [3,4]. Hydrogen has several positive features, including its outstanding energy density per mass unit (120–140 MJ/kg) and the absence of harmful byproducts when used in a fuel cell [4,5,6,7]. Although the production of hydrogen is currently largely based on techniques using fossil sources, such as hydrocarbon reforming and gasification, there are some promising methods utilizing renewable energy sources as well [5,8,9,10]. For example, photocatalytic water splitting and hydrogen generation from biomass are particularly attractive methods for green hydrogen generation; furthermore, there are also attempts to apply biological processes for such purpose [5,8,9]. Hydrogen can be employed in a number of transportation and stationary applications; for example, fuel cell electric vehicles may be an alternative to internal combustion vehicles and also to electric vehicles using Li-ion batteries [6,11]. Buses, trains, trams and other similar vehicles are particularly suitable for the introduction of hydrogen fuel cells, since these applications do not need an extensive refueling infrastructure [11]. Hydrogen can also be used as a stationary energy storage medium, such as a complementary power supply for solar powered households [12] and off-grid charging stations [13], in case of the absence of sunlight. There are proposals for large scale storage of renewable energy on a daily and even seasonal basis [14,15].
One of the major problems of widespread commercialization of hydrogen-based technologies is the lack of efficient and application-friendly hydrogen storage solutions [16,17]. Depending on the use case, the hydrogen storage system should possess high storage density (either volumetrically and/or gravimetrically), fast charge/discharge rate, adequate operational cycle life and delivery temperature/pressure [4]. As an example, for light-duty vehicles, the US Department of Energy (DOE) set up technical targets for a range of system parameters [18].
There are various methods to store hydrogen, including the traditional solutions, i.e., storage of hydrogen in high-pressure vessels and storage in liquid form. Although special composite high-pressure vessels can handle pressures up to 70 MPa [19,20] and they are currently utilized in several vehicular systems [11], the size of these containers still implies a limitation for some applications. While liquid storage mitigates this issue, the considerable energy required for cooling hydrogen to 21 K and the unavoidable boil-off of H2 make this method impractical in most cases [3,5]. Solid state hydrogen storage can be a promising alternative to the above-mentioned techniques, as they generally offer larger volumetric hydrogen densities [4,17,20,21]. Hydrogen can be adsorbed in molecular form on the surface of several carbon-based materials, such as graphite, carbon nanotubes, graphene and metal-organic frameworks (MOF) [22,23,24]. These systems offer good H-sorption rates and reversibility; however, higher capacities can usually be achieved only at lower temperatures (~77 K). Alternatively, hydrogen can be stored in the form of different metal hydrides (LaNi5H6, TiFeH2, MgH2, Mg2NiH4, etc.) [16,25,26,27,28], complex hydrides (alanates, such as LiAlH4; borohydrides, such as LiBH4) [4,29,30] and chemical hydrides (NH3BH3; amides, such as LiNH2) [4,30,31]. The main advantage of hydrides, compared to other storage methods, is their excellent storage capacities. However, reversibility can be an issue, particularly in case of complex hydrides and chemical hydrides. Additionally, complex hydrides and some metal hydrides possess high desorption temperatures.
Among metallic hydrides, MgH2 is considered to be one of the best candidates to be incorporated into the hydrogen-based economy due to its high storage capacity and low cost. Recent research and development targeting new approaches for improving the H-storage performance of Mg-based systems is comprehensively expounded [32]. It has undoubtedly been confirmed that high-energy ball milling is an ideal tool to improve the hydrogenation/dehydrogenation performance of these materials via the creation of abundant lattice defects [33].
In this review paper, first a brief summary will be given on the physics of the thermodynamics and kinetics of hydrogenation of elemental magnesium (Section 2). Since nanostructuring is an essential tool to considerably improve the kinetics of Mg-based systems, details of the most efficient top-down technique, i.e., ball milling will be highlighted in Section 3. The main body of this study (Section 4) will be dedicated to the very recent research trends (covering years from 2016 through 2021) of Mg-based hydrogen storage systems doped by conventional and innovative additives including the following:
  • Transition metal-based catalysts;
  • Intermetallic compound catalysts;
  • Catalysts formed during in situ reactions;
  • Synergistic multiple catalysts;
  • Catalysts with special morphology.
At the end of the text, a brief discussion on the future prospects will be highlighted (Section 5).

2. A Short Overview of MgH2 and Its Hydrogen Storage Properties

As was cited above in this review paper, solid-state hydrogen storage is still a significant technological challenge; nevertheless, numerous promising efforts have occurred thus far. Magnesium and magnesium-based alloys, compounds and composites have been extensively studied as hydrogen storage materials for several decades, including the enormous progress in the last 20 years [34]. Among metallic elements, magnesium attracts the highest interest in the field of solid-state hydrogen storage due to its very high theoretical gravimetric (7.6 wt.%) and volumetric (0.11 kg H/L) capacities. Magnesium has several other advantages, i.e., it is lightweight, non-toxic, abundant and relatively safe [26,35]. Moreover, Mg can be produced by well-established technology and the raw material cost is relatively small [27]. The reaction of Mg with H2 can be inferred from the corresponding binary phase diagram (Figure 1), which has been calculated using the Thermo-Calc program PARROT [36]. At moderate pressures, the only hydride phase that co-exists with Mg in equilibrium is MgH2, which has a tetragonal TiO2 (rutile type) unit cell with a space group of P42/mnm. This low-pressure hydride phase is commonly designated as β-MgH2 [26]. The Mg–H bonding interaction in β-MgH2 is rather complex, mainly ionic, but exhibits some covalent character as well [37]. As was stated by the authors, this weak but significant covalent bonding can be a huge advantage on the hydrogenation–dehydrogenation performance of elemental magnesium. Upon increasing the pressure above 8 GPa, a polymorphic structural transformation occurs in the Mg–H system, leading to the formation of a high-pressure orthorhombic γ-MgH2 phase (unit cell structure PbO2) [26].
The thermodynamic aspects of metallic hydride formation can be understood from Figure 2. At the very early stages of H-absorption (when the hydrogen concentration CH < 0.1 H/M), the hydrogen atoms can occupy interstitial positions in the metallic lattice, resulting in the formation of a solid solution (α-phase), see also the left-hand side image. By continuing the rapid pressure increase, only a limited amount of extra hydrogen atoms can be absorbed in the α-phase; instead, the nucleation of a stoichiometric hydride phase (β-phase) commences at a critical H-concentration, which is then followed by a wide plateau. In this wide concentration range, the α-phase and β-phase co-exist, and they are in thermodynamic equilibrium. This concentration range determines the amount of the reversible absorbed hydrogen. When the total volume of the material is transformed into the hydride phase (see the right-hand side image), a second steep pressure rise can be observed in the pressure-composition isotherms, relating to the capture of new H-atoms in interstitial sites of the β-phase. It is also envisaged from Figure 2 that the width of plateau has a strong temperature dependence, which is totally eliminated at a critical temperature, TC. Above this temperature, the transition between the two phases is continuous. peq at a given temperature is the pressure at which the hydride phase is in equilibrium with the gaseous H2. For pressures p < peq, the hydride phase is thermodynamically unstable, which leads to the dehydrogenation of the β-phase. When the external pressure is larger than peq, hydrogen absorption becomes favorable.
The temperature dependence of peq corresponding to the α↔β reaction can be given using the following Van’t Hoff equation [38]:
l n p e q p e q 0 = Δ H R · 1 T Δ S R ,
where Δ H and Δ S are the formation enthalpy and entropy of the hydride phase and p e q 0 is a reference pressure, usually taken as the normal atmospheric pressure. From the slope and the intercept of the l n p e q p e q 0   v s .   1 T function, Δ H and Δ S can be determined, see Figure 2. As Δ S mainly corresponds to the formation of molecular H2 gas dissolved in the solid state and vice versa, it can be approximated using the standard entropy of hydrogen for all metal-hydride systems: Δ S = −0.135 kJ mol−1K−1H2 [38]. On the other hand, the Δ H enthalpy term, covering a wide range for different metals, corresponds to the stability of the metal–hydrogen bond. Based on Equation (1), to reach peq = 1 bar at room temperature, Δ H should be taken as −39 kJ mol−1H2.
Since the formation enthalpy of tetragonal MgH2 phase is relatively high ( Δ H = −74.5 kJ mol−1 H2 [39]), one can conclude that to desorb H2 under the applied pressure of p = 1 bar, magnesium hydride should be heated to 550 K, see the corresponding Figure 3. It is also recognized that for peq = 10 bar, an applied temperature of ~640 K is required. Thus, it is evident that due to the high stability of the Mg–H bonds, the hydrogenation/dehydrogenation process of elemental magnesium can only operate at elevated temperatures.
When a hydrogen molecule approaches the magnesium surface (at a distance ~0.2 nm), a physisorbed state is formed by a weak Van der Waals force [38]. At shorter distances, the H2 molecule has to overcome an energy barrier (Eact) for dissociation and the formation of a new Mg–H bond. The magnitude of Eact strongly depends on the type of the metal involved in the hydrogenation process. In the next step, these chemisorbed hydrogen atoms penetrate into the subsurface layer and then diffuse towards the interstitial sites of metal lattice, forming the intermetallic hydride phase. Since Eact for the hydrogenation and dehydrogenation of elemental Mg/MgH2 is rather high: E a c t a b s = 90 kJ/molH and E a c t d e s = 152–180 kJ/molH [40,41,42,43,44,45,46,47], respectively, the hydrogen absorption/desorption kinetics are poor at ambient temperatures. The above described thermodynamical limitations of hydrogenation in magnesium and the sluggish sorption kinetics are the common drawbacks of the on-board commercialization of MgH2 [25]. In order to overcome these difficulties, the key issue is the simultaneous development of both the kinetic and thermodynamic performance of Mg-based systems.
One possible solution to alter the thermodynamics in the favorable direction without a significant capacity loss is reducing the crystallite size to nanometric regimes. Thus far, a huge amount of research has been targeting the nanostructuring process [48,49,50]; however, it was shown that the reduction in the Δ H hydride formation enthalpy is significant only for clusters containing ~10 Mg atoms [51]. Anyway, nanocrystallization has a more enhanced effect on the sorption kinetics, since it increases the surface area and promotes the formation of lattice defects (dislocations, vacancies), which positively influences the diffusion of hydrogen along the grain boundaries into the bulk of the Mg particles [52,53]. The different coordination number of the Mg and H atoms in the grain boundary regions, coupled with a unique surface morphology, are responsible for the improved kinetic performance of Mg-nanoparticles [54]. In some cases, especially when the crystallite size reduction is manifested by high-energy ball milling, the formation of a metastable orthorhombic γ-MgH2 phase also takes place, which results in the weakening of the Mg–H bonds and a decrease in the hydrogen sorption temperature [55].
Alternatively, Δ H can also be reduced if Mg is alloyed with another element. In the classical case, transition metals, such as Ni, were applied [56,57,58], leading to the formation of the intermetallic Mg2Ni compound, where the formation enthalpy of the hydride phase is reduced to Δ H = −64.5 kJ mol−1H2. Unfortunately, the maximum storage capacity is also significantly reduced (3.6 wt.%) at the same time. The destabilization of magnesium hydride without a significant capacity loss is the most important challenge by the selection of the proper additive. High-energy ball milling of Mg or MgH2 with a small number of catalysts (up to 5–10 mol%) such as metal oxides, metal halides and many others has been used to improve the kinetic performance. During this nanocrystallization process, a homogeneous dispersion of the catalyst particles occurs throughout the whole composite material, increasing the number of contact sites with the hydride phase. Nevertheless, ball milling still exhibits some limitations for certain Mg-based systems, i.e., hydrogenation cycling can destroy the defect structure generated during the deformation process [32].

3. High-Energy Ball Milling

High-energy ball milling (HEBM) or mechanical attrition produces nanostructured materials by structural decomposition of coarse-grained structures as the result of severe plastic deformation (SPD). This has become a widely used technique to synthesize nanocrystalline powders because of its simplicity, the relatively inexpensive equipment and the applicability to essentially all classes of materials [59]. During milling, only crystallite size reduction is required; therefore, reasonably short milling times are sufficient. Nevertheless, in some cases, the following serious problems are usually cited: contamination from milling media and/or milling atmosphere, potential fire risk and the need to consolidate the powder product with maintaining its original nanostructural feature.

3.1. Ball Milling Devices

A variety of different types of ball mills have been applied extensively in the last couple of decades for the mechanical processing of powders, including vibratory mills/shaker mills, planetary mills and attritors. The most common ball mills used for experimental studies are vibratory mills, such as the SPEX 8000 mill. In these mills, a cylindrical vial of about 0.1-liter capacity undergoes a vibratory motion in the horizontal direction (see Figure 4, left). The vial contains typically 1 g of powder and 5–20 g of grinding ball(s), resulting in a 5:1–20:1 ball to powder mass ratio and vibrates at a frequency and amplitude of approximately 20 Hz and 10–50 mm, respectively [60]. Vibratory mills are convenient for laboratory experiments; however, they have not been scaled up to a larger size yet. A significantly larger amount of powder material can be processed in planetary mills, which can exhibit ~0.25–1.0 L of vial capacity. In some cases, 2–4 such containers are rotated about two separated parallel axes, similarly to the rotation of the Earth around the Sun (see Figure 4, right). The diameter of the grinding balls is larger than in vibratory mills, and consequently, the ball to powder mass ratio can also be much higher. Attritor mills are widely used for ultrafine grinding of ceramics and industrial minerals. Milling occurs in a stationary container filled up with a huge number of grinding balls, which are mixed by impellers attached to a vertical drive axis. A major problem experienced using vertical-axis attritor mills for dry grinding is the gravity-driven segregation of the powder to the bottom of the mill [59]. Nevertheless, these devices are especially capable of the large-scale production of up to tons of milled powders.

3.2. Formation of Nanoparticles during Ball Milling

The process of HEBM is characterized by repeated collisions between the grinding balls, the vial and the powder particles. Until now, several extensive models describing the dynamics of ball milling and its influence on microstructural changes of milled powders have been carried out [60,61,62,63]. These studies have pointed out that the products are dependent on the physical and mechanical properties of the milled powder and the type of the grinding media. Parameters associated with the milling process, such as impact velocity, impact force, collision energy, ball mass and density, ball to powder mass ratio and vial temperature were examined [64,65] and studied using computer simulations [66]. Using a free falling experiment, a close analogy to the collision events occurring in vibratory mills indicated the magnitude of the impact force increases with an increasing ball size for any impact velocities [64]. It was also concluded that the impact between the ball and the plate is significantly influenced by the thickness of the powder layer.
The deformation process of a powder agglomerate during an impact with the milling balls can be inferred from Figure 5. When the ball impacts the powder first, the powder behaves as a porous material, as illustrated in Figure 5 left. During the impact, powder particles slide and rotate, rearranging their positions due to particle interactions [68]. As a result, the porosity of the powder decreases, while some particles may escape from the contact area. The porous structure of the powder will be altered until it is compressed to a critical height of h’ (Figure 5, center). In this stage, the fraction of the kinetic energy of the ball is dissipated. As the impact progresses further, the elastic and plastic deformation of the individual powder particles trapped in the contact area commences (Figure 5, right).

4. Recent Research Trends in the Catalysis of Mg-Based Hydrogen Storage Materials

4.1. Transition Metal-Based Additives for Catalyzing MgH2 and Other Mg-Based Materials

As was demonstrated in Section 2, nanocrystalline Ni as an additive to Mg/MgH2 positively affects the hydrogen storage properties due to the reduction in Δ H [69]. When the intermetallic compound Mg2Ni is achieved via the HEBM method from a Mg–Ni powder mixture followed by compression, the dehydrogenation occurs at a lower temperature [70]. The hydrogenated state is mainly characterized by a mixture of hexagonal Mg2NiH0.3 solid solution and monoclinic Mg2NiH4 phase [71]. It was also obtained that the nanocrystalline Mg2Ni to Mg2NiH4 phase transformation exhibits an enthalpy of Δ H = −57.47 kJ/molH2 [69]. When ultra-high pressures (2.5 to 4 GPa) were applied for several days to a hydrogenated Mg-Ni alloy during compression, some fraction of the stable tetragonal β-MgH2 was converted into the metastable orthorhombic γ-MgH2 phase, yielding the decrease in the onset hydrogen desorption temperature to 60 °C [72]. By further optimizing the synthesis conditions, it is believed that—if all the stable MgH2 transforms into γ-MgH2—desorption of 4.5 wt.% H2 can take place at 200 °C. In a thorough study, it was demonstrated that the chemical nature of Ni-based additives to nanocrystalline Mg is critical to their H-storage performance [73]. Ni catalyst nanopowders prepared by HEBM with MgH2 exhibited a broad particle size distribution, while those produced from anhydrous NiCl2 were 10–100 times finer and much more uniform, both in shape and size. It is evident that the selection of the initial catalyst form and their controlled dispersion on MgH2 can be a tool for tailoring the hydrogen storage properties of transition metal-doped Mg-based materials. Subsequent nanoscale Ni-based additives, such as Ni3C, Ni3N, NiO and Ni2P, revealed a strong reduction in the onset temperature of dehydrogenation of ball-milled MgH2, reaching a value as low as 160 °C for the MgH2–Ni3C composite [45]. The overall dehydrogenation performance of the series of these selected compounds directly correlates with the electron-negativity of the dopants. Low energy wet milling of Mg with Ni+V additives under different organic solvents indicated that the diffusion of hydrogen is facilitated in these systems, since the surfactant effect on particle surface is improved [74]. It turned out that among the three organic solvents tested (C6H6, C6H12 and C4H8O), benzene provides the maximum hydrogen absorption capacity.
The catalytic effect of nanocrystalline Fe on ball-milled Mg is outstanding, i.e., this composite could be hydrogenated even at 0 °C up to 45% of the theoretical capacity [75]. The significantly improved absorption–desorption behavior with respect to pristine Mg is attributed to the nano-engineered surface of the magnesium particles. These homogeneously distributed nano-Fe particles on the Mg surface indicate the homogeneous activation of the material. At the same time, a negligible capacity loss is observed up to 50 full sorption cycles, while dehydrogenation can be ascribed by a three-dimensional controlled diffusion. A nanocomposite mixture of MgH2, Mg2FeH6 and Fe has been achieved by planetary milling of Mg and Fe under 30 bar H2 pressure [76]. It was demonstrated that the amount of the Mg2FeH6 complex hydride increased with the increasing milling time. In situ synchrotron XRD experiments upon heating showed that MgH2 is the first hydride to decompose. These nanocomposites present very fast sorption kinetics, i.e., desorption is completed within 2 min at 350 °C. By analyzing the hydrogen uptake measurements of Mg catalyzed by 10 wt.% Fe2O3 taken at different temperatures, it was established that a rate limiting step of the initial absorption is the dissociative chemisorption of the H2 molecules on the particle surfaces [77]. On the other hand, results obtained from the absorption rate function revealed that the initial hydrogenation stage is followed by the diffusion of H-atoms through the MgH2 layer.
TiO2 is considered as one of the best metal-oxide catalysts to enhance the hydrogenation/dehydrogenation properties of magnesium, including the reduction in the onset of the desorption temperature. Very recently, aside from the type and crystallite size of the catalyst additives, it was also confirmed that the shape and morphology of these particles may show a strong correlation with the intrinsic H-storage properties of the host material [78]. In the research of Z. Ma and coworkers, TiO2 nanosheets exposed with different amounts of {001} and {101} facets have been synthesized and ball milled with MgH2. The authors have found that a huge number of H-absorbing active sites and defects formed on the surface of the high surface energy {001} facets dominated TiO2, which can significantly accelerate the hydrogen dissociation and recombination, see the schematic image in Figure 6 [79]. In addition, the hydrogen desorption activation energy of MgH2 is also remarkably reduced to −67.6 kJ/mol when TiO2 nanosheets exposed with {001} facets doped into MgH2 [80].
When MgH2 was milled together with a porous carbon-supported TiO2 nanocomposite catalyst, the onset of the dehydrogenation temperature is reduced by 95 °C, while the cyclic stability has improved significantly [42]. In the corresponding density functional calculations, the main absorption structures of an MgH2 molecule on low free surface energy rutile (110) TiO2 and anatase (101) TiO2 surfaces were constructed, and the results indicated that the Mg–H bonds are extended and weakened, in agreement with the experimental observations. These findings confirmed that TiO2 acts as an important catalytic additive to lower the kinetic barrier of the dehydrogenation of MgH2. When Mg is catalyzed by Ti3AlC2 MAX-phase, the kinetic mechanism of hydrogen desorption can be ascribed using the Johnson–Mehl–Avrami process with an exponent n = 3.09, corresponding to a constant nucleation rate and three-dimensional growth [81]. On the contrary, the contracting volume model describes the dehydrogenation performance of the pristine nanocrystalline MgH2, indicating that the MAX-phase promotes a change in the hydrogen sorption mechanism.
A couple of years ago, titanium isopropoxide (TTIP), as an organic alkoxide, was used as an additive to MgH2 and its effect was compared to the benchmark Nb2O5 catalyst [82]. The authors found that only a very small amount (0.5 mol%) of TTIP milled with MgH2 was sufficient to enhance the kinetics, producing equally good results as Nb2O5 and it has a superior hydrogen capacity. Since TTIP is liquid at room temperature, it may disperse more readily among the MgH2 crystalline particles than most powder additives. In addition, a simple hand mixing of TTIP with the hydride powder demonstrated excellent desorption behavior, most probably due to complete and continuous coverage of the MgH2 particles by the liquid additive. A study by the same authors on the effect of the HEBM gas environment (Ar, H2) was conducted on MgH2 milled together with different additives, such as TTIP, Nb2O5 and fullerene C60 [83]. It was found that the milling environment has very little effect on the dehydrogenation performance of the hydride phase; however, a significant difference takes place in the absorption kinetics when Nb2O5 is milled with MgH2 under a H2 or Ar atmosphere. It is interesting that an Ar milling environment has a superior effect on the hydrogenation of MgH2 catalyzed by Nb2O5. In a parallel investigation, another alkoxide liquid, i.e., Nb(V) ethoxide [Nb(OCH2–CH3)5], was hand mixed with pre-milled MgH2 in order to combine the positive influence of Nb-based catalysts on the hydrogen storage properties with the highly dispersed state achieved by liquid additives [84]. Upon heating, Nb(OCH2–CH3)5 reacts with MgH2, releasing C2H6. The absorption and desorption activation energies are significantly reduced, and the dehydrogenation occurs 80 °C below the decomposition temperature of HEBM MgH2. Irrespective of the amount of the Nb(V) ethoxide additive (0.1–1 mol%), the decomposition of MgH2 occurs 90 °C below the value of pure hydride [85]. An excellent reversibility is obtained (5.3 and 5.1 wt.% for absorption and desorption) at 300 °C in 3 min.
A notable improvement in the reduction in the H-desorption temperature can be achieved when MgH2 is cryo-milled together with fluoride additives [86]. NbF5 and FeF3 have a stronger mechanical effect on the formation of a very fine microstructure than the corresponding oxides (Nb2O5 and Fe2O3), since they act as a lubricant agent during the milling process.
In a detailed study on the hydrogen storage performance of nanocrystalline MgH2 catalyzed by different amounts of TiH2, it was pointed out that an optimal TiH2 content exists (0.025 mol%) [87]. As seen in Figure 7, increasing the TiH2 content decreases the H-capacity of Mg/MgH2 due to the higher atomic weight of Ti compared to Mg and the formation of irreversible TiH2 takes place. On the contrary, TiH2 addition enhances the sorption kinetics due to the faster hydrogen diffusion in TiH2 than in MgH2 and improves the reversible cycling stability, which leads to a well-determined optimum TiH2 content. Using different processing devices (planetary and oscillating ball mills) for the mechanical grinding of the MgH2–TiH2 system, it was found that the kinetics of hydrogen absorption are more dependent on the raw starting components than the manufacturing processing parameters [88].
A different amount of TiO2 addition (x = 0–10 wt.%) to a rare-earth containing a Mg80Ni10La7Ce3 alloy by mechanical milling yields the improvement on the thermodynamics and hydride stability, reaching the minimum dehydrogenation enthalpy and the minimum onset desorption temperature for x = 5 wt.% [89]. XRD and complimentary HRTEM analysis of the hydride state revealed a multi-step reaction pathway, including the formation of several hydrides, i.e., MgH2, Mg2NiH4, LaH3 and CeH3. It was also established that the phase interfaces of the latter two hydrides provide low activation energy diffusion channels for the hydrogen atoms. A novel VNbO5 ternary oxide was synthesized by annealing of pre-milled mixture of V2O5 and Nb2O5 in a 1:1 molar ratio and then was HEBMed to commercial MgH2 [90]. The catalytic effect of VNbO5 is remarkable, since the desorption temperature of the doped MgH2 decreased to 200 °C after the first hydrogenation cycle, which is in correlation with the good dispersion of the catalyst particles in the whole system.

4.2. Intermetallic Compounds as Catalysts for the H-Sorption of MgH2

Very recently, various intermetallic compounds were also added to MgH2 by ball milling to improve its H-sorption properties. Mg–40 wt.% TiFe nanocomposites with different milling parameters have been synthesized by adding TiFe intermetallic compound to Mg by HEBM. As a first step, the TiFe intermetallic compound was produced by pre-milling of TiH2 and Fe powders [91]. When the TiFe catalyst is synthesized under ethanol to improve its refinement level, no traces of untransformed Mg were detected after hydrogenation at room temperature. This can be explained by the thermal activation of the Mg/TiFe interface that enhances the initial diffusion of decomposed hydrogen during the absorption process. The best initial hydrogenation kinetics can be attained when the Mg–TiFe nanocomposite was prepared in a planetary mill (3 wt.% after 60 min of absorption); nevertheless, the maximum overall capacity (4.0 wt.%) can be reached after HEBM in a shaker mill [92].
Very recently, El-Eskandarany et al. successfully synthesized a new synthetic binary Mg–5 wt.% TiMn2 nanocomposite system with an advanced hydrogenation/dehydrogenation performance [93]. The beneficial effect of TiMn2 addition on the cyclability is outstanding, with respect to pure milled nanocrystalline Mg, the nanocomposite powder shows almost zero degradation of the achieved H-storage capacity after 1000 sorption cycles. Even at this stage, the TiMn2 nanoparticles are homogeneously adhered onto the surface of the MgH2 powders. Based on additional microstructural experiments, the authors concluded that this system neglects any undesired grain-coarsening, which is crucial from the point of view of industrial applications. To attain this purpose, the as-synthesized nanocomposite powder was employed as a source in a complete hydrogen storage system for fuel-cell applications. A mixture of γ-MgH2 and β-MgH2 with a nanosized crystallite structure was stabilized by the addition of 10 wt.% ZrNi5 powders by reactive milling under a hydrogen gas atmosphere [94]. The corresponding HRTEM study presented in Figure 8 confirms the formation of intimate fine crystals (5–15 nm). The as-milled MgH2–10 ZrNi5 nanocomposite exhibits an outstanding H-absorption performance, i.e., full absorption can be reached within 1 min at 275 °C. As a continuation of this research, Ti2Ni metallic glassy nanopowders were added to MgH2 via cryo-milling [95]. Structural investigations revealed that the short-range order atomic structure was preserved after the severe plastic deformation during the HEBM procedure. In addition, this disordered structure is stable up to 400 °C. Similar to the prior results, the Mg–10 wt.% Ti2Ni system also possesses excellent life-time cyclic stability, achieving more than 80 full hydrogenation/dehydrogenation cycles without any degradation, which can directly be associated with the stable nano-morphology of the MgH2 nanoparticles. Similar results have been obtained when 7 wt.% amorphous LaNi3 is milled to MgH2, i.e., increased milling time promotes the uniform distribution of the amorphous nanoparticles in the hydride matrix [96]. In addition, these hard LaNi3 nanoparticles are capable of penetrating through the MgO layer, which facilitates excellent H-absorption (6 wt.% at 200 °C) and extraordinary long cyclic lifetime.
Large scale applications of Mg-based nanocomposites have been targeted by the HEBM synthesis of Mg-(Fe20V80) material [97]. The increased addition of FeV to Mg facilitates the hydrogenation performance of MgH2, including faster kinetics and lower activation energy of hydride decomposition, while the H-absorption mechanism changes from random nucleation and 3D growth (undoped MgH2, Avrami exponent n = 4) to diffusion controlled (Mg–10 wt.% (Fe20V80), Avrami exponent n = 2.5). HEBM of 5 at.% VTiCr with MgH2 was carried out under high pressure (100 bar) hydrogen gas. The excellent hydrogenation kinetics and cyclic stability of MgH2 doped by a bcc VTiCr alloy is attributed to the perfectly distributed nano-sized catalyst particles [98]. In this research, it was demonstrated that this VTiCr–MgH2 alloy can absorb hydrogen at pressures as low as 0.04 bar, due to the thermodynamically feasible reactions. Pre-milled Zr0.67Ni0.33 amorphous powder was added as a precursor to nanocrystalline MgH2 via shaker mill [99]. Structural and morphological experiments showed that the addition of 10 wt.% amorphous powder led to the formation of very fine MgH2 nanoparticles and resulted in enhanced sorption kinetics. This composite can absorb 4 wt.% H2 within a minute at 250 °C. The nucleation of a small amount of ZrO2 can further catalyze the hydrogenation process by reducing the activation energy of hydrogen dissociation. Furthermore, a synergistic effect between the amorphous Zr-Ni particles and the possible dissolution of Ni can destabilize the MgH2 phase. Combined ball milling and hydride combustion was successfully applied to synthesize a novel MgH2–MgC0.5Co3 composite in order to ameliorate the hydrogenation performance of MgH2 [100]. Compared to the nanocrystalline MgH2, the onset temperature of dehydrogenation of the composite material is as low as 237 °C. At the same time, the MgC0.5Co3 additive was stable; no phase transformation occurred during several sorption cycles.

4.3. Formation of Catalysts during In Situ Reactions

In certain cases, an in situ reaction can take place between several additives during the HEBM process, resulting in a compositional change. In other cases, a reduction of the catalyst by MgH2 can also occur either during the ball milling procedure or during the dehydrogenation/hydrogenation cycling.
It was demonstrated in a systematic study that, under special circumstances, the high-energy ball milling can induce a reaction between MgH2 and a stable metal oxide additive Nb2O5 [101]. Reduction of the metal oxide by MgH2 occurs over a wide composition range, and as a result MgxNbyOx+y rock salt phase forms. However, 10 h of milling was necessary for this transformation to occur, and by further extending the milling time (30 h), an increased yield of the rock salt phase could be achieved. The newly formed MgxNbyOx+y promotes the dehydrogenation of MgH2, as was indicated by the decreased desorption peak temperature (~35 °C reduction). Changing the material of the grinding medium and the vial (from steel to zirconia) enables the modification of the in situ reaction, as it results in a blend of several rock salt products with different stoichiometries [102]. A similar observation was made in case of TiO2 additive mixed to MgH2 via a planetary mill using a relatively high ball to powder weight ratio (70:1), i.e., reduction of TiO2 and in situ formation of Ti dissolved MgO (MgxTiyOx+y phase) occurred during prolonged (30 h) milling [103]. Such a product was shown to enhance the desorption kinetics of MgH2 and to lower the desorption temperature [104]. In case of both Nb2O5 and TiO2, it was suggested that the metal dissolved MgO is a catalytically active phase [102,104]; additionally, the dissolution of Nb leads to the expansion of the MgO lattice, which can cause cracks in the MgO layer, enhancing the diffusion of hydrogen [102].
Another study investigated the catalytic effect of TiVO3.5 oxide on the hydrogen sorption of MgH2 [41]. It was found that during ball milling, TiVO3.5 is reduced by MgH2, and very fine metallic Ti and V nanoparticles are created. These nanoparticles can improve the dissociation and recombination of H2 molecules and act as a hydrogen pump for the magnesium hydride. It was also pointed out that the catalytic activity of the in situ formed Ti and V nanoparticles are superior to those added directly to MgH2, as it is difficult to reduce the particle size down to several nanometers and disperse them homogenously by ball milling only. In some cases, the reduction of the additive and the formation of in situ species do not take place during the HEBM treatment, but rather during the following desorption–absorption cycling, as was demonstrated in ref. [105]. To be specific, after 20 cycles, the 5 wt.% of Fe2O3 and Fe3O4 catalysts are completely reduced to metallic Fe.
Complex metal oxides were recently investigated as additives for MgH2 due to their ability to react with the hydride during the desorption process and produce in situ formed phases. A partial reduction of MnFe2O4 occurred during the dehydrogenation (T = 450 °C) of 10 wt.% MnFe2O4 catalyzed MgH2 [106]. According to the authors, the produced Fe particles and Mg0.9Mn0.1O oxide phase may be the real active species, enhancing the sorption properties of the hydride. A similar study was conducted on CuFe2O4 doped MgH2, in which case, CuFe2O4 is partially reduced to Mg2Cu, Fe and MgO during the desorption process (see the XRD patterns in Figure 9) [107]. It was suggested that the formation of these catalytically active phases may be reason of the decreased desorption temperature and improved kinetics. N.A. Ali and co-workers synthesized MgNiO2 nanoflakes using hydrothermal method and doped to MgH2 [108]. The MgNiO2 reacts with MgH2 during the H-sorption and as a result MgO and NiO form. It was demonstrated that the absorption/desorption rate increased, and the desorption activation energy decreased significantly, compared to the as-milled MgH2. It is interesting to note that samples catalyzed with only MgO or NiO do not show significant improvement in the desorption performance over the as-milled MgH2, which indicates that these in situ created phases have a remarkable synergistic catalytic effect as well.
Several works have demonstrated that a reaction can take place between Ni additive and MgH2 during the dehydrogenation process, namely the additive is reduced and an Mg2Ni intermetallic phase forms [109,110,111,112,113,114,115]. In general, this transformation occurs in the course of the first desorption, while in case of the subsequent cycles, a reversible conversion of Mg2Ni↔Mg2NiH4 takes place alongside the Mg↔MgH2 reaction. The kinetic measurements of such composites show excellent absorption/desorption behavior; for example, MgH2 doped by NiMoO4 nanorods is able to absorb 5.5 wt.% hydrogen in 10 min at 150 °C and 3.2 MPa [114]. Similarly, MgH2 catalyzed by carbon encapsulated iron-nickel nanoparticles (Fe0.64Ni0.36@C) has reached an absorption capacity of 5.18 wt.% hydrogen in 20 min at 150°C and 3 MPa [44]. P. Yao et al. investigated the catalytic effect of Ni@rGO (Ni nanoparticles anchored on reduced graphene oxide) on the H-storage properties of MgH2, they measured a good absorption rate (3.7 wt.% in 10 min) even at a temperature as low as 100 °C (p = 3 MPa) [112]. The desorption kinetics was also fast, i.e., 6.1 wt.% hydrogen could be desorbed in 15 min (T = 300 °C, p = 0.0004 MPa). It was indicated that the in situ formed Mg2Ni/Mg2NiH4 can improve the catalytic activity over unreacted Ni [112]. Even better dehydrogenation properties were attained for Ni/TiO2 nanocomposite-doped MgH2, namely, the amount of desorbed hydrogen can reach 6.5 wt.% in 7 min at 265 °C and 0.002 MPa [109]. In such composites, the remarkable H-storage performance is generally attributed to the in situ formed phase (i.e., Mg2Ni/Mg2NiH4). The catalytic mechanism was well explained in ref [116]; accordingly, the Mg2NiH4 phase decomposes earlier than MgH2 during the desorption process; the newly created Mg2Ni can assist the dehydrogenation of MgH2 in a way that H atoms first diffuse to the Mg2Ni and then to the solid–gas interface. This mechanism is often referred to as a “hydrogen pump” [44,117,118], since the intermetallic phase acts as a diffusion channel for hydrogen and also serves as a heterogeneous nucleation site. It was also pointed out that the micro-strain associated with the volume change occurring during the Mg2Ni↔Mg2NiH4 transformation is beneficial for the diffusion of H and, thus, the kinetic performance of the composite [110,119]. When using Ni-containing additives, besides the formation of Mg2Ni, other in situ formed phases can also appear, such as MgS in MgH2 catalyzed by carbon supported Ni3S2 [120] and α-Fe in MgH2 doped by a transition metal-based MOF (metal-organic framework) [121]. These catalytic species play an important role in the hydrogenation/dehydrogenation process and usually provide active sites for the nucleation reaction. Doping nanocrystalline Mg2FeH6 by Ni promotes the reaction of Ni with MgH2 and Fe, resulting in the formation of Mg2NiH4 and FeNi3 phases, respectively [122]. The Mg2FeH6–20Ni alloy possesses a two-step dehydrogenation sequence, in contrast to the single-step reaction of the as-prepared material. The reaction between Mg2Ni (from the decomposition of Mg2NiH4) and Mg2FeH6 forms a new Mg-Fe-Ni-H quaternary hydride with a significantly lower onset dehydrogenation temperature.
Other material classes have also shown the ability to form in situ phases in combination with MgH2; for example, a recent paper studied different transition metal sulfides (such as TiS2, NbS2, MoS2, MnS, CoS2 and CuS) as additives [123]. In situ formed phases such as MgS, TiH2, NbH, Mo, Mn, Mg2CoH5 and MgCu2 can be viewed as an effective catalyst of the (de)hydrogenation reaction, as the sorption kinetics significantly improved, mainly at lower temperatures. MgS and TiH2 are suggested to serve as nucleation sites for Mg/MgH2 and diffusion channels for hydrogen, Nb/NbH and Mg2CoH5/Mg2Co mainly act as hydrogen pump, while Mo, Mn and MgCu2 can facilitate the decomposition of MgH2 by weakening the Mg-H bond. There are also examples of desorption-induced phases in cases of fluoride (K2SiF6 [124], K2NbF7 [125]), and chloride (HfCl4) [126] materials, where KH, MgF2, Mg2Si, Nb and MgCl2 phases appear, respectively. Investigations on MgH2 catalyzed by FeB [127], CoFeB [128] and CoB/CNT [129] indicated that in situ generated phases also appear in case of transition metal borides. The formed Fe [127,128], B [127,128], CoFe [128] and Co3MgC [128,129] phases are considered to provide additional nucleation sites for the Mg↔MgH2 phase transformation, while Fe can also enhance the dissociation of H2 molecules.
Based on the above presented reports, it is clear that in situ formed (either during ball milling or H-sorption) catalysts have a remarkable ability to improve the hydrogen sorption performance of Mg-based materials. In terms of catalytic activity, they often surpass those additives that are not formed in situ, but only mixed with MgH2 through ball milling. The reason behind this observation can have multiple origins: (i) the in situ formed phases have altered the electronic structure compared to the original additive [41,102,109,128], which could be beneficial to the interaction with hydrogen atoms/molecules; (ii) the small size and uniform distribution (which were observed in multiple cases [41,109,110,117,128]) are also an important feature for in situ generated catalysts, since more interphase regions and a larger surface area covered by these products lead to an enhanced overall catalytic activity.

4.4. Utilization of Synergy between Multiple Catalysts

As was also presented in Section 4.1, additives such as transition metals, transition metal oxides and carbon-based materials are effective catalysts for the dehydrogenation of MgH2. However, an increasing number of publications demonstrated that greater improvements can be achieved in the sorption properties by combining different additives than using a single catalyst. These observations were attributed to a synergistic catalytic mechanism occurring during the (de)hydrogenation process.
In a recent paper, Ni/TiO2 core-shell particles were introduced to MgH2 by ball milling and the kinetics of the composite were analyzed [130]. In the presence of the Ni/TiO2 co-catalyst, a significantly faster hydrogen desorption could be observed than for single Ni or TiO2 additives and the difference is more striking at lower temperatures (T = 250 °C). It was suggested that the Ni/TiO2 core-shell structure can assist electron transport and, thus, enhance the recombination of hydrogen; additionally, the TiO2 coating prevents grain growth and stabilizes the morphology. Another example of a synergistic catalysis is given by Chen et al., who dispersed Co nanoparticles on TiO2 and this nanocomposite (Co/TiO2) was added to MgH2 through ball milling [131]. It was possible to achieve a lower desorption peak temperature (Tdes = 235.2 °C compared to 329.4 and 288.4 °C for Co and TiO2, respectively) and faster isothermal desorption kinetics with Co/TiO2 than either with Co or TiO2. These improvements can be attributed to the synergistic effect of Co and TiO2, i.e., Co can destabilize the MgH2 owing to the higher strength of the Co–H bonds than that of Mg–H; meanwhile, the electrons from the conduction band of TiO2 migrate to the Co, and as a result of the interaction of H with the resulting holes, facilitate the recombination reaction. Another investigation was conducted on the catalytic performance of sandwich-like Ni/Ti3C2 in MgH2, where an electron transfer was observed between Ni and Ti3C2 [132]. The numerous interfaces presented by the catalyst and its altered electronic state can effectively facilitate the hydrogen absorption/desorption processes. It was also suggested that Ni has a positive effect on the dissociation of hydrogen and on the weakening of the Mg–H bond, while Ti3C2 can prevent the agglomeration of Ni or MgH2 particles. The multiple valence states of Ti (Ti0, Ti3+), appearing during ball milling, can assist the conversion between Mg2+ and Mg. The simultaneous addition of TiF3 and Nb2O5 to MgH2 via a vibratory type high-energy ball mill was studied in a recent piece of research [46]. A lower desorption temperature and dehydrogenation activation energy were found than when using only one type of catalyst. The additives are suggested to enhance the diffusion and recombination processes of hydrogen and also serve as nucleation sites for the Mg phase.
In some cases, in situ formed phases (in detail: Section 4.3) can also have a synergistic catalytic effect on the hydrogen sorption of magnesium hydride. In NiTiO3-doped MgH2, Mg2Ni/Mg2NiH4 forms during the desorption/absorption process and these phases can enhance the H-sorption reactions in combination with multiple valence titanium compounds (Ti4+, Ti3+ and Ti2+), which are generated during the ball milling and also during dehydrogenation [116]. Similar findings were obtained for MgH2 catalyzed by Ni/TiO2 [109], for which the synergistic effect was demonstrated through DSC and TPD measurements (see Figure 10), i.e., lower desorption temperatures can be found for the composite additive than for single catalysts. In both of the above cases, the underlying mechanism was proposed to be a multi-step process; firstly, the multivalent titanium facilitates the electron transfer associated with the decomposition of MgH2 (H transfers an electron to Ti4+, while Mg2+ obtains an electron from Ti2+). The formed H atom can now diffuse to Mg2Ni, which serves as a fast diffusion channel, also called as a “hydrogen pump” [116]. In case of transition metal-based MOF (metal-organic framework)-doped MgH2, a synergistic improvement of the desorption was demonstrated for the in situ formed Mg2Ni/Mg2NiH4 and α-Fe phases [121]. Iron primarily enhances the nucleation and growth of Mg on the interface of α-Fe and MgH2, while Mg2Ni, working as a “hydrogen pump”, provides channels for the migration of H atoms (see Figure 11).
A synergistic mechanism was also observed when transition metal carbides (TiC, ZrC, WC) were milled to a Mg–Ni alloy [133]. Accordingly, the desorption starts with the decomposition of Mg2NiH4, the stress accompanied by this phase transformation assists the dehydrogenation of MgH2 together with the carbide phase, as the latter can facilitate the charge transfer between Mg2+ and H. The carbide and Mg2Ni can also act as nucleation sites for Mg/MgH2.
A combination of carbon-based additives with other types of catalysts is also proven to be an effective way to achieve improved hydrogen storage properties for Mg-based materials. For example, a simultaneous application of graphene oxide-based porous carbon and TiCl3 yields better desorption kinetics for MgH2 than the individual use of these catalysts [134]. According to the authors, the carbon additive mainly serves as a scaffold material, thus it prevents the agglomeration of MgH2 and ensures the homogeneous distribution of TiCl3 particles. Similar findings were concluded when various materials supported on graphene were added to MgH2 or an Mg-based alloy through ball milling. In case of TiH2 templated over graphene, the graphene support effectively inhibited the agglomeration of TiH2; additionally, the milling procedure introduced defects to the graphene, which is suggested to play a co-catalytic role with the TiH2 [135]. Another study demonstrated that an electronic interaction exists between FeCoNi nanoparticles and their graphene support, which leads to an enhanced catalytic effect [136]. In addition to the important role of preventing agglomeration and improving the dispersion of additives, graphene can also provide numerous phase boundaries owing to its high surface area, which serves as diffusion channels for hydrogen. This feature also enables a synergistic catalytic effect with different materials that mainly improve the dissociation reaction, such as Ni [137] or TiF3 [138], to improve the sorption properties of a Mg–Al alloy. Graphene can also assist the dissociation of hydrogen in combination with NbN0.9O0.1 when ball milled to MgH2, as was reported in ref. [139]. The NbN0.9O0.1 phase, which was formed in situ from N-doped Nb2O5 nanorods, was shown to weaken the Mg–H and H–H bonds considerably, and the degree of this effect is ameliorated if graphene was also present. Synergistic improvement of the hydrogen storage properties was also reported for composites of carbon and Ni [118] or Co [118,140] nanoparticles, for which carbon layers form around the metallic catalyst. In this arrangement, the carbon can effectively prevent aggregation of the nanoparticles while maintaining their catalytic activity.
Reduced graphene oxide (rGO) was applied in several cases to support different transition metal-based catalysts, such as Ni [112], Ni3Fe (see Figure 12) [119], NiS [110] and V2O3 [141] in Mg-based hydrogen storage materials. A decrease in the hydrogen desorption temperature and acceleration of kinetics were observed upon introduction of the co-catalysts, owing to the synergistic effect between rGO and transition metal compounds (either directly added or in situ formed). The reduced graphene oxide mainly improves the dispersion of catalyst particles and stabilizes their morphology, thus ensuring better overall catalytic activity and cyclic stability. It can also provide a diffusion channel for hydrogen; hence, H atoms dissociated on the surface of transition metal-based additives can spill over to the Mg [110].
Besides graphene, carbon nanotubes (CNT) were also combined with other additives to improve the hydrogen sorption properties of MgH2. Gao et al. reported a lower dehydrogenation temperature for MgH2 doped with FeB [127] or CoFeB [128] and decorated with CNTs, compared to the case where only FeB or CoFeB was used, respectively. Another paper has shown an increased absorption rate upon simultaneous application of TiF3 and multiwalled carbon nanotubes in a Mg-Ni alloy [142]. For Mg–Nb2O5–CNT composites prepared by high-energy milling, improved kinetics was observed [143] and the use of multiple catalysts enabled a desorption capacity close to the theoretical value to be achieved (6.43 wt.%) [144]. In most of these systems, the different catalysts are believed to play different roles in the hydrogen absorption/desorption process, namely the transition metal-based compounds provide nucleation sites for MgH2/Mg and can also accelerate dissociation/recombination of H2 molecules at the surface. At the same time, nanotubes, owing to their morphology, are suggested to promote the diffusion of hydrogen through the interior of the particles. It was indicated that CNT may also have a positive effect on the dissociation of hydrogen [142], and can suppress extensive grain growth during cycling [144,145].
The above cited examples demonstrate the significant advantages of applying multiple catalysts to improve the hydrogen storage performance of Mg-based materials. The common key feature of these systems is the combination of different catalytic mechanisms through introducing different types of additives. It is also worth mentioning that the extent of the synergistic effect depends on the preparation conditions of the catalysts and the composite, as was shown in refs [109,112,132,134].

4.5. Catalysts with Special Morphology

Apart from the chemical composition, morphological features of the applied additives can also be an important factor to consider when one designs a novel hydrogen storage material. In the last few years, numerous works have been dedicated to catalysts with unique morphology to enhance the hydrogen ab-/desorption performance of Mg-based materials.
In the previous section we already mentioned some systems in which the morphology of the additives plays an important role; these are mainly catalysts with carbon support, which prevents the excessive agglomeration of particles. Core-shell structures are particularly interesting, as they can effectively separate particles that would otherwise be prone to aggregation. For example, an amorphous carbon layer is formed around transition metal nanoparticle cores such as Ni [118] and Co [118,140], which resulted in remarkable catalytic activity, as indicated by fast absorption/desorption reactions (see Section 4.4). In ref. [44], carbon was coated on an iron-nickel alloy (Fe0.64Ni0.36@C) and this was added to Mg by ball milling. While the milling process did not destroy the core-shell structure, it ensured the distribution of the catalyst in the Mg-based composite. A slightly different approach was presented by S. Wang and coworkers, i.e., Ni nanoparticles were dispersed on porous hollow carbon nanospheres, which were synthesized separately [115]. In this case, the Ni particles (with an average size of ~10 nm) could be found on the surface as well as in the inner cavities of the hollow carbon nanospheres; this enabled the high loading (90 wt.%) of Ni with no agglomeration. Adding this catalyst to MgH2 in a planetary ball mill yielded notable hydrogen sorption rates (6.3 wt.% hydrogen absorbed in 100 s at T = 200 °C and p = 5 MPa). In spite of the carbon shell covering the metallic particles, an in situ reaction between Mg and the transition metal could occur for all the above examples. This means that the carbon either does not form a continuous layer or can be broken, so that the core can interact with the Mg matrix as well as with hydrogen; at the same time, agglomeration of the particles is still suppressed. This is an important attribute to note for future research in this field.
M. Zhang et al. synthesized flower-like carbon wrapped TiO2 (TiO2@C) using a combination of the solvothermal process and low temperature pyrolysis and subsequently added this to MgH2 thorough ball milling [146]. The flower-like structure of the catalyst is composed of numerous short rods (see Figure 13, left); however, the ball milling process destroys this morphology, and many small pieces of the catalyst are created instead, which, in turn, cover the surface of MgH2 particles (Figure 13, right). In this way, the initial loose structure of the additive promotes the uniform distribution of the catalyst and enables the observed fast absorption rates.
Catalysts with layered structures were also applied to improve the sorption properties of MgH2, for example, TiVO3.5, with a multilayered structure, was shown to considerably decrease the absorption/desorption temperature (e.g., the dehydrogenation temperature reduced by 75 °C) and the activation energy (59% reduction compared to pristine MgH2) [41]. Another investigation presented a Ni/Ti3C2 additive with a sandwich-like structure fabricated by modified wet chemical synthesis [132]. Fast hydrogen sorption and good cyclic stability were observed for MgH2 catalyzed by the sandwich-like Ni/Ti3C2. According to the authors, the layered structure of the catalyst presents plenty of interfaces that play a substantial role in the hydrogenation reaction, as these offer hydrogen diffusion paths and assist the electron transfer process (see also Section 4.4).
Catalysts with a reduced dimension have caught some attention recently, as the interaction with hydrogen can be different in these materials than in their bulk counterparts. ZrCo nanosheets prepared via a wet chemical method were ball milled to MgH2, and as a result the hydrogen sorption properties of MgH2 were improved, as indicated by the considerable reduction in the dehydrogenation temperature [147]. The ZrCo nanosheets covering the surface of MgH2 were assumed to work as a hydrogen pump. Nanosheets can be also interesting, because it is possible to expose a specific high-surface-energy crystallographic plane in large quantity, as was shown in case of TiO2 nanosheets (see also Section 4.1) [80]. Additionally, the nanosheets are also able to restrain the excessive grain growth of MgH2 during the absorption/desorption cycling.
Carbon nanotubes are often applied as an additive to Mg-based hydrogen storage materials owing to their special tubular morphology. In many cases, they are used in combination with other, mainly transition metal-based catalysts, in order to achieve a synergistic catalytic effect (see Section 4.4) [128,142,144]. In general, it is assumed that the nanotubes can enhance the diffusion of hydrogen by serving as diffusion channels. However, a satisfactory catalytic effect can only be achieved by the breakdown of the large agglomerations of CNTs, as insufficient dispersion can lead to slow absorption/desorption kinetics [148]. On the other hand, as the dispersion process is usually ball milling, it is important to limit the milling time and intensity, because the nanotube structure can be damaged by the high-energy impacts, and amorphous carbon can form [144]. Hence, in case of such catalysts with special morphology, optimal milling parameters should be applied; the use of lower energy ball milling may be advantageous.
Not only carbon can form a nanotube structure, other materials, such as Ti-based oxides, can also have such morphology. Recently, titanate nanotubes prepared using the hydrothermal method from anatase TiO2 was applied to improve the sorption properties of magnesium using high-energy ball milling and high-pressure torsion. The morphological differences due to the variation of the milling time were found to have a significant impact on the absorption kinetics of Mg [149]. Another group synthesized Na2Ti3O7 nanotubes and nanorods using the hydrothermal and solid-state method, respectively (see Figure 14), and mixed them to MgH2 in a planetary ball mill. It was shown that the nanotubes have a superior catalytic effect over the nanorods, as indicated by the 52.6 °C difference in the dehydrogenation peak temperature and also by the higher desorption rate [43]. The Na2Ti3O7 nanotubes were more effective in catalyzing the desorption of MgH2 than bulk Na2Ti3O7, which demonstrates the advantages of the nanotube morphology. According to the authors, the Na2Ti3O7 nanotubes serve as diffusion channels along Mg/MgH2 interfaces.
Several transition metal-based nanorods were prepared and tested to improve the storage properties of MgH2. Porous rod-like NiTiO3 and CoTiO3 can decrease the dehydrogenation temperature, although NiTiO3 provides better catalytic effect (Tdes = 261.5 °C and 298 °C, respectively, compared to the Tdes = 322 °C of ball milled MgH2) [116]. NiMoO4 and CoMoO4 nanorods similarly enhance the non-isothermal and also the isothermal desorption performance of magnesium hydride [114]. The catalytic effect in both of the above papers was mainly attributed to the in situ formed intermetallic phase; it can be assumed that the nanorod morphology, owing to its high active surface area, can promote this transformation. N-doped Nb2O5 nanorods were synthesized on graphene support and were milled to MgH2 [139]. A significant decrease was obtained for the dehydrogenation barrier, i.e., the apparent activation energy was changed from 139 kJ/mol (as-milled MgH2) to 81 kJ/mol after the addition of the nanorod catalyst.
Carbon nanofibers, which supported Ni nanoparticles, were produced using the electrospinning method and then ball milled to MgH2 [150]. The nanofibers ensured the homogeneous distribution of Ni particles around the hydride phase and also played the important role of preventing the aggregation of Ni. This effect led to fast absorption/desorption reactions and excellent cyclic stability (capacity retention of 99.8% in the first 20 cycles).
S. Hu et al. investigated the catalytic effect of ultrathin K2Ti8O17 nanobelts on the H-sorption kinetics of MgH2 [47]. Excellent absorption/desorption rates were observed even at lower temperatures and a marked decrease was found in the dehydrogenation activation energy, compared to the pristine MgH2 (from 175.3 kJ/mol to 116.3 kJ/mol). The catalytic effect was attributed to the oxygen vacancies in the nanobelts formed during the ball milling preparation process. It was also proposed that the special morphology of the nanobelts is beneficial to expose these oxygen vacancies.
Based on the cited investigations, there is a certain potential in the catalysts with unique morphology to improve the H-storage properties of Mg-based materials. The high surface area, which is a common feature of these additives, can not only enhance the interaction with hydrogen (such as dissociation), but the increased number of interfaces can also improve the diffusion properties of the material. Hence, proper dispersion of the catalyst over the Mg matrix is of great importance; on the other hand, keeping the original morphology during the dispersion process (which is usually ball milling) is equally essential in most cases.
At the end of this review paper, we summarize the hydrogen sorption properties of some selected ball-milled Mg-based systems in Table 1. As one can recognize, reasonable hydrogenation capacities (5–6.2 wt.%) can be measured for most of the materials at absorption temperatures as low as Tabs = 100–200 °C. The lowest Tabs values correspond to nanocrystalline Mg doped by very different types of additives, including K2Ti8O17 nanobelts, CoFeB plus CNT catalysts and in situ formed N-Nb2O5 with graphene. On the contrary, similar capacity values (5–6.8 wt.%) for dehydrogenation can only be achieved at elevated temperatures, typically at Tdes = 250–300 °C. The best performance includes catalysts, such as TiVO3.5, sandwich-like Ni/Ti3C2 and K2Ti8O17 nanobelts. It is also realized from Table 1 that all of the cited catalysts promote a significant reduction in the E a c t d e s value when it is compared to pure Mg. The best performance is realized when magnesium is milled together with TiVO3.5, NiS with reduced graphene oxide or flower-like TiO2/C.

5. Future Perspectives

Magnesium-based composites hold a significant potential for future applications in the field of hydrogen storage. However, substantial improvements still have to be accomplished before they can be readily utilized in practice, in spite of the remarkable research activity and progress carried out, especially in the last 5 years. Notably, the vast majority of the investigations reported here have focused on the enhancement of the kinetic properties of absorption/desorption reactions. Nevertheless, due to the low equilibrium pressure of the Mg-MgH2 system, the reduction in the dehydrogenation temperature is strongly limited. Hence, changing the thermodynamic properties of this system and increasing the equilibrium pressure are highly desirable to improve the hydrogen release process.
Based on the reports discussed in this review, some important aspects can be highlighted that could support future works. Taking the advantage of the reducing ability of MgH2 and the formation of in situ phases holds a significant potential for fabricating Mg-based hydrogen storage materials with improved properties. Hence, it is recommended to explore such phenomenon for other material combinations as well. A combination of multiple catalysts is also proved to be a promising way to enhance to hydrogen sorption properties of MgH2. In order to find the best combinations, the effect of additives on the different rate limiting steps of the sorption reactions and the mechanism of the synergistic effect should be taken into account, thus it would be worthwhile to explore these aspects in greater detail. Additionally, the sample preparation process may also affect the synergy of the catalysts; a systematic study is needed to clarify the underlying mechanism.

6. Summary

In this paper, we have reviewed the most promising investigations, carried out in the last 5 years, on Mg-based hydrogen storage materials prepared by high-energy ball milling. There were a large number of works focusing on nanocomposites containing Mg or MgH2 and a small amount (typically 5–10 wt.%) of additives. The effect of these additives has been discussed, focusing on their catalytic mechanisms. A wide range of materials has been covered, including different transition metals and transition metal-based compounds. Although these material-classes are used for a long time in this field, some novel interesting advances were attained, such as experiments on liquid additives and investigations on catalysts with specific crystallographic orientations. Additionally, various intermetallic compounds have been shown to accelerate the absorption/desorption rate. Numerous recent reports presented evidence of reactions between the host magnesium-based material and the additive, either during the milling procedure or during hydrogen cycling. It was demonstrated that various catalytically active species can be formed through these in situ reactions, which resulted in a superior hydrogen sorption performance. The utilization of multiple catalysts is another increasingly popular strategy in Mg-based systems. By carefully choosing the combination of additives and the preparation conditions, remarkable improvements can be achieved in the hydrogenation/dehydrogenation kinetics and cycling stability through the synergistic effect of catalysts. Apart from the chemical composition, morphology can also be a relevant aspect of an additive, as was demonstrated through various examples. Different types of nanorods, nanotubes, nanosheets and core-shell structures have shown the ability to enhance the H-sorption properties of MgH2 and increase the cycling stability.

Author Contributions

Á.R. and M.G., writing—original draft preparation; Á.R., supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was completed in the ELTE Institutional Excellence Program (TKP2020-IKA-05) financed by the Hungarian Ministry of Human Capacities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. International Energy Agency. Key World Energy Statistics 2020. Available online: https://www.iea.org/reports/key-world-energy-statistics-2020 (accessed on 13 September 2021).
  2. Directorate-General for Energy (European Commission). EU Energy in Figures: Statistical Pocketbook 2020; Publications Office of the European Union: Luxembourg, 2020; ISBN 978-92-76-19443-9. [Google Scholar]
  3. Abe, J.O.; Popoola, A.P.I.; Ajenifuja, E.; Popoola, O.M. Hydrogen energy, economy and storage: Review and recommendation. Int. J. Hydrogen Energy 2019, 44, 15072–15086. [Google Scholar] [CrossRef]
  4. Rusman, N.A.A.; Dahari, M. A review on the current progress of metal hydrides material for solid-state hydrogen storage applications. Int. J. Hydrogen Energy 2016, 41, 12108–12126. [Google Scholar] [CrossRef]
  5. Abdalla, A.M.; Hossain, S.; Nisfindy, O.B.; Azad, A.T.; Dawood, M.; Azad, A.K. Hydrogen production, storage, transportation and key challenges with applications: A review. Energy Convers. Manag. 2018, 165, 602–627. [Google Scholar] [CrossRef]
  6. Yang, J.; Sudik, A.; Wolverton, C.; Siegel, D.J. High capacity hydrogen storage materials: Attributes for automotive applications and techniques for materials discovery. Chem. Soc. Rev. 2010, 39, 656–675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Niaz, S.; Manzoor, T.; Pandith, A.H. Hydrogen storage: Materials, methods and perspectives. Renew. Sustain. Energy Rev. 2015, 50, 457–469. [Google Scholar] [CrossRef]
  8. Martino, M.; Ruocco, C.; Meloni, E.; Pullumbi, P.; Palma, V. Main hydrogen production processes: An overview. Catalysts 2021, 11, 547. [Google Scholar] [CrossRef]
  9. Nazir, H.; Louis, C.; Jose, S.; Prakash, J.; Muthuswamy, N.; Buan, M.E.M.; Flox, C.; Chavan, S.; Shi, X.; Kauranen, P.; et al. Is the H2 economy realizable in the foreseeable future? Part I: H2 production methods. Int. J. Hydrogen Energy 2020, 45, 13777–13788. [Google Scholar] [CrossRef]
  10. Sharma, S.; Ghoshal, S.K. Hydrogen the future transportation fuel: From production to applications. Renew. Sustain. Energy Rev. 2015, 43, 1151–1158. [Google Scholar] [CrossRef]
  11. Nazir, H.; Muthuswamy, N.; Louis, C.; Jose, S.; Prakash, J.; Buan, M.E.M.; Flox, C.; Chavan, S.; Shi, X.; Kauranen, P.; et al. Is the H2 economy realizable in the foreseeable future? Part III: H2 usage technologies, applications, and challenges and opportunities. Int. J. Hydrogen Energy 2020, 45, 28217–28239. [Google Scholar] [CrossRef]
  12. Hosseini, S.E.; Butler, B.; Abdul Wahid, M. Hydrogen as a battery for a rooftop household solar power generation unit. Int. J. Hydrogen Energy 2020, 45, 25811–25826. [Google Scholar] [CrossRef]
  13. Mehrjerdi, H. Off-grid solar powered charging station for electric and hydrogen vehicles including fuel cell and hydrogen storage. Int. J. Hydrogen Energy 2019, 44, 11574–11583. [Google Scholar] [CrossRef]
  14. Lin, R.-H.; Zhao, Y.-Y.; Wu, B.-D. Toward a hydrogen society: Hydrogen and smart grid integration. Int. J. Hydrogen Energy 2020, 45, 20164–20175. [Google Scholar] [CrossRef]
  15. Brey, J.J. Use of hydrogen as a seasonal energy storage system to manage renewable power deployment in Spain by 2030. Int. J. Hydrogen Energy 2021, 46, 17447–17457. [Google Scholar] [CrossRef]
  16. Tarasov, B.P.; Fursikov, P.V.; Volodin, A.A.; Bocharnikov, M.S.; Shimkus, Y.Y.; Kashin, A.M.; Yartys, V.A.; Chidziva, S.; Pasupathi, S.; Lototskyy, M.V. Metal hydride hydrogen storage and compression systems for energy storage technologies. Int. J. Hydrogen Energy 2021, 46, 13647–13657. [Google Scholar] [CrossRef]
  17. Ren, J.; Musyoka, N.M.; Langmi, H.W.; Mathe, M.; Liao, S. Current research trends and perspectives on materials-based hydrogen storage solutions: A critical review. Int. J. Hydrogen Energy 2017, 42, 289–311. [Google Scholar] [CrossRef]
  18. US DOE Target Explanation Document: Onboard Hydrogen Storage for Light-Duty Fuel Cell Vehicles. Available online: https://www.energy.gov/eere/fuelcells/doe-technical-targets-onboard-hydrogen-storage-light-duty-vehicles (accessed on 13 September 2021).
  19. 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]
  20. Barthelemy, H.; Weber, M.; Barbier, F. Hydrogen storage: Recent improvements and industrial perspectives. Int. J. Hydrogen Energy 2017, 42, 7254–7262. [Google Scholar] [CrossRef]
  21. Shao, H.; He, L.; Lin, H.; Li, H.-W. Progress and trends in magnesium-based materials for energy-storage research: A review. Energy Technol. 2018, 6, 445–458. [Google Scholar] [CrossRef] [Green Version]
  22. Ströbel, R.; Garche, J.; Moseley, P.T.; Jörissen, L.; Wolf, G. Hydrogen storage by carbon materials. J. Power Sources 2006, 159, 781–801. [Google Scholar] [CrossRef]
  23. Broom, D.P.; Webb, C.J.; Fanourgakis, G.S.; Froudakis, G.E.; Trikalitis, P.N.; Hirscher, M. Concepts for improving hydrogen storage in nanoporous materials. Int. J. Hydrogen Energy 2019, 44, 7768–7779. [Google Scholar] [CrossRef]
  24. Wang, L.; Yang, R.T. Hydrogen storage on carbon-based adsorbents and storage at ambient temperature by hydrogen spillover. Catal. Rev. 2010, 52, 411–461. [Google Scholar] [CrossRef]
  25. Wang, H.; Lin, H.J.; Cai, W.T.; Ouyang, L.Z.; 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]
  26. Varin, R.A.; Czujko, T.; Wronski, Z.S. Nanomaterials for Solid State Hydrogen Storage; Springer Science: New York, NY, USA, 2009. [Google Scholar]
  27. 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]
  28. Zhou, C.; Zhang, J.; Bowman, R.C.; Fang, Z.Z. Roles of Ti-based catalysts on magnesium hydride and its hydrogen storage properties. Inorganics 2021, 9, 36. [Google Scholar] [CrossRef]
  29. Milanese, C.; Jensen, T.R.; Hauback, B.C.; Pistidda, C.; Dornheim, M.; Yang, H.; Lombardo, L.; Zuettel, A.; Filinchuk, Y.; Ngene, P.; et al. Complex hydrides for energy storage. Int. J. Hydrogen Energy 2019, 44, 7860–7874. [Google Scholar] [CrossRef] [Green Version]
  30. Mohtadi, R.; Orimo, S. The renaissance of hydrides as energy materials. Nat. Rev. Mater. 2016, 2, 16091. [Google Scholar] [CrossRef]
  31. Jain, I.P.; Jain, P.; Jain, A. Novel hydrogen storage materials: A review of lightweight complex hydrides. J. Alloys Compd. 2010, 503, 303–339. [Google Scholar] [CrossRef]
  32. Lyu, J.; Lider, A.; Kudiiarov, V. Using ball milling for modification of the hydrogenation/dehydrogenation process in magnesium-based hydrogen storage materials: An overview. Metals 2019, 9, 768. [Google Scholar] [CrossRef] [Green Version]
  33. Révész, Á.; Kánya, Z.; Verebélyi, T.; Szabó, P.J.; Zhilyaev, A.P.; Spassov, T. The effect of high-pressure torsion on the microstructure and hydrogen absorption kinetics of ball-milled Mg70Ni30. J. Alloys Compd. 2010, 504, 83–88. [Google Scholar] [CrossRef]
  34. Aguey-Zinsou, K.-F.; Ares-Fernández, J.-R. Hydrogen in magnesium: New perspectives toward functional stores. Energy Environ. Sci. 2010, 3, 526–543. [Google Scholar] [CrossRef]
  35. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef]
  36. Zeng, K.; Klassen, T.; Oelerich, W.; Bormann, R. Critical assessment and thermodynamic modeling of the Mg–H system. Int. J. Hydrogen Energy 1999, 24, 989–1004. [Google Scholar] [CrossRef]
  37. Noritake, T.; Aoki, M.; Towata, S.; Seno, Y.; Hirose, Y.; Nishibori, E.; Takata, M.; Sakata, M. Chemical bonding of hydrogen in MgH2. Appl. Phys. Lett. 2002, 81, 2008–2010. [Google Scholar] [CrossRef]
  38. Züttel, A. Materials for hydrogen storage. Mater. Today 2003, 6, 24–33. [Google Scholar] [CrossRef]
  39. 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 2016, 122, 97. [Google Scholar] [CrossRef] [Green Version]
  40. Fernández, J.F.; Sánchez, C.R. Rate determining step in the absorption and desorption of hydrogen by magnesium. J. Alloys Compd. 2002, 340, 189–198. [Google Scholar] [CrossRef]
  41. Zhang, X.; Shen, Z.; Jian, N.; Hu, J.; Du, F.; Yao, J.; Gao, M.; Liu, Y.; Pan, H. A novel complex oxide TiVO3.5 as a highly active catalytic precursor for improving the hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2018, 43, 23327–23335. [Google Scholar] [CrossRef]
  42. Zhang, X.; Leng, Z.; Gao, M.; Hu, J.; Du, F.; Yao, J.; Pan, H.; Liu, Y. Enhanced hydrogen storage properties of MgH2 catalyzed with carbon-supported nanocrystalline TiO2. J. Power Sources 2018, 398, 183–192. [Google Scholar] [CrossRef]
  43. Zhang, L.; Chen, L.; Fan, X.; Xiao, X.; Zheng, J.; Huang, X. Enhanced hydrogen storage properties of MgH2 with numerous hydrogen diffusion channels provided by Na2Ti3O7 nanotubes. J. Mater. Chem. A 2017, 5, 6178–6185. [Google Scholar] [CrossRef]
  44. Ding, Z.; Fu, Y.; Zhang, L.; Rodríguez-Pérez, I.A.; Zhang, H.; Wang, W.; Li, Y.; Han, S. Improve hydrogen sorption kinetics of MgH2 by doping carbon-encapsulated iron-nickel nanoparticles. J. Alloys Compd. 2020, 843, 156035. [Google Scholar] [CrossRef]
  45. Zhang, Q.; Zang, L.; Huang, Y.; Gao, P.; Jiao, L.; Yuan, H.; Wang, Y. Improved hydrogen storage properties of MgH2 with Ni-based compounds. Int. J. Hydrogen Energy 2017, 42, 24247–24255. [Google Scholar] [CrossRef]
  46. Zhang, T.; Hou, X.; Hu, R.; Kou, H.; Li, J. Non-isothermal synergetic catalytic effect of TiF3 and Nb2O5 on dehydrogenation high-energy ball milled MgH2. Mater. Chem. Phys. 2016, 183, 65–75. [Google Scholar] [CrossRef]
  47. Hu, S.; Zhang, H.; Yuan, Z.; Wang, Y.; Fan, G.; Fan, Y.; Liu, B. Ultrathin K2Ti8O17 nanobelts for improving the hydrogen storage kinetics of MgH2. J. Alloys Compd. 2021, 881, 160571. [Google Scholar] [CrossRef]
  48. Zaluski, L.; Zaluska, A.; Tessier, P.; Stroem-Olsen, J.O.; Schulz, R. Nanocrystalline hydrogen absorbing alloys. Mater. Sci. Forum 1996, 225–227, 853–858. [Google Scholar] [CrossRef]
  49. Oelerich, W.; Klassen, T.; Bormann, R. Metal oxides as catalysts for improved hydrogen sorption in nanocrystalline Mg-based materials. J. Alloys Compd. 2001, 315, 237–242. [Google Scholar] [CrossRef]
  50. Fátay, D.; Révész, Á.; Spassov, T. Particle size and catalytic effect on the dehydriding of MgH2. J. Alloys Compd. 2005, 399, 237–241. [Google Scholar] [CrossRef]
  51. Pasquini, L. The effects of nanostructure on the hydrogen sorption properties of magnesium-based metallic compounds: A review. Crystals 2018, 8, 106. [Google Scholar] [CrossRef] [Green Version]
  52. Fátay, D.; Spassov, T.; Delchev, P.; Ribárik, G.; Révész, A. Microstructural development in nanocrystalline MgH2 during H-absorption/desorption cycling. Int. J. Hydrogen Energy 2007, 32, 2914–2919. [Google Scholar] [CrossRef]
  53. Révész, Á.; Fátay, D. Microstructural evolution of ball-milled MgH2 during a complete dehydrogenation–hydrogenation cycle. J. Power Sources 2010, 195, 6997–7002. [Google Scholar] [CrossRef]
  54. Paidar, V. Magnesium hydrides and their phase transitions. Int. J. Hydrogen Energy 2016, 41, 9769–9773. [Google Scholar] [CrossRef]
  55. Zhou, S.; Zhang, Q.; Chen, H.; Zang, X.; Zhou, X.; Wang, R.; Jiang, X.; Yang, B.; Jiang, R. Crystalline structure, energy calculation and dehydriding thermodynamics of magnesium hydride from reactive milling. Int. J. Hydrogen Energy 2015, 40, 11484–11490. [Google Scholar] [CrossRef]
  56. Urretavizcaya, G.; García, G.; Serafini, D.; Meyer, G. Mg-Ni alloys for hydrogen storage obtained by ball milling. Lat. Am. Appl. Res. 2002, 32, 289–294. [Google Scholar]
  57. Révész, Á.; Gajdics, M.; Spassov, T. Microstructural evolution of ball-milled Mg–Ni powder during hydrogen sorption. Int. J. Hydrogen Energy 2013, 38, 8342–8349. [Google Scholar] [CrossRef]
  58. Orimo, S.; Ikeda, K.; Fujii, H.; Fujikawa, Y.; Kitano, Y.; Yamamoto, K. Structural and hydriding properties of the Mg-Ni-H system with nano- and/or amorphous structures. Acta Mater. 1997, 45, 2271–2278. [Google Scholar] [CrossRef]
  59. Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  60. Urakaev, F.K.H. 2—Mechanism and kinetics of mechanochemical processes. In High-Energy Ball Milling; Sopicka-Lizer, M., Ed.; Woodhead Publishing: Sawston, UK, 2010; pp. 9–44. ISBN 978-1-84569-531-6. [Google Scholar]
  61. Abdellaoui, M.; Gaffet, E. The physics of mechanical alloying in a planetary ball mill: Mathematical treatment. Acta Metall. Mater. 1995, 43, 1087–1098. [Google Scholar] [CrossRef]
  62. Abdellaoui, M.; Gaffet, E. The physics of mechanical alloying in a modified horizontal rod mill: Mathematical treatment. Acta Mater. 1996, 44, 725–734. [Google Scholar] [CrossRef]
  63. Magini, M.; Iasonna, A.; Padella, F. Ball milling: An experimental support to the energy transfer evaluated by the collision model. Scr. Mater. 1996, 34, 13–19. [Google Scholar] [CrossRef]
  64. Huang, H.; Dallimore, M.P.; Pan, J.; McCormick, P.G. An investigation of the effect of powder on the impact characteristics between a ball and a plate using free falling experiments. Mater. Sci. Eng. A 1998, 241, 38–47. [Google Scholar] [CrossRef]
  65. Schaffer, G.B.; Forrester, J.S. The influence of collision energy and strain accumulation on the kinetics of mechanical alloying. J. Mater. Sci. 1997, 32, 3157–3162. [Google Scholar] [CrossRef]
  66. Huang, H.; Pan, J.; McCormick, P.G. On the dynamics of mechanical milling in a vibratory mill. Mater. Sci. Eng. A 1997, 232, 55–62. [Google Scholar] [CrossRef]
  67. Zolriasatein, A.; Shokuhfar, A.; Safari, F.; Abdi, N. Comparative study of SPEX and planetary milling methods for the fabrication of complex metallic alloy nanoparticles. Micro Nano Lett. 2018, 13, 448–451. [Google Scholar] [CrossRef]
  68. Révész, Á. Microstructure and Thermal Properties of Nanocrystalline Materials. Ph.D. Thesis, Eötvös University, Budapest, Hungary, 2000. [Google Scholar]
  69. Révész, Á.; Gajdics, M.; Schafler, E.; Calizzi, M.; Pasquini, L. Dehydrogenation-hydrogenation characteristics of nanocrystalline Mg2Ni powders compacted by high-pressure torsion. J. Alloys Compd. 2017, 702, 84–91. [Google Scholar] [CrossRef]
  70. Khan, D.; Zou, J.; Zeng, X.; Ding, W. Hydrogen storage properties of nanocrystalline Mg2Ni prepared from compressed 2MgH2–Ni powder. Int. J. Hydrogen Energy 2018, 43, 22391–22400. [Google Scholar] [CrossRef]
  71. Gajdics, M.; Calizzi, M.; Pasquini, L.; Schafler, E.; Révész, Á. Characterization of a nanocrystalline Mg–Ni alloy processed by high-pressure torsion during hydrogenation and dehydrogenation. Int. J. Hydrogen Energy 2016, 41, 9803–9809. [Google Scholar] [CrossRef]
  72. Weng, Z.; Retita, I.; Tseng, Y.-S.; Berry, A.J.; Scott, D.R.; Leung, D.; Wang, Y.; Chan, S.L.I. γ-MgH2 induced by high pressure for low temperature dehydrogenation. Int. J. Hydrogen Energy 2021, 46, 5441–5448. [Google Scholar] [CrossRef]
  73. House, S.D.; Vajo, J.J.; Ren, C.; Zaluzec, N.J.; Rockett, A.A.; Robertson, I.M. Impact of initial catalyst form on the 3D structure and performance of ball-milled Ni-catalyzed MgH2 for hydrogen storage. Int. J. Hydrogen Energy 2017, 42, 5177–5187. [Google Scholar] [CrossRef] [Green Version]
  74. Sakamoto, T.; Atsumi, H. Hydrogen absorption/desorption characteristics of Mg-V-Ni hydrogen storage alloys. Fusion Eng. Des. 2019, 138, 6–9. [Google Scholar] [CrossRef]
  75. Kumar, S.; Singh, A.; Tiwari, G.P.; Kojima, Y.; Kain, V. Thermodynamics and kinetics of nano-engineered Mg-MgH2 system for reversible hydrogen storage application. Thermochim. Acta 2017, 652, 103–108. [Google Scholar] [CrossRef]
  76. Antiqueira, F.J.; Leiva, D.R.; Zepon, G.; de Cunha, B.F.R.F.; Figueroa, S.J.A.; Botta, W.J. Fast hydrogen absorption/desorption kinetics in reactive milled Mg-8 Mol% Fe nanocomposites. Int. J. Hydrogen Energy 2020, 45, 12408–12418. [Google Scholar] [CrossRef]
  77. Song, M.Y.; Kwak, Y.J. Hydrogen charging kinetics of Mg-10wt% Fe2O3 prepared via MgH2-forming mechanical milling. Mater. Res. Bull. 2021, 140, 111304. [Google Scholar] [CrossRef]
  78. 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]
  79. 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]
  80. Zhang, M.; Xiao, X.; Wang, X.; Chen, M.; Lu, Y.; Liu, M.; Chen, L. Excellent catalysis of TiO2 nanosheets with high-surface-energy {001} facets on the hydrogen storage properties of MgH2. Nanoscale 2019, 11, 7465–7473. [Google Scholar] [CrossRef]
  81. Lakhnik, A.M.; Kirian, I.M.; Rud, A.D. The Mg/MAX-phase composite for hydrogen storage. Int. J. Hydrogen Energy 2021. [Google Scholar] [CrossRef]
  82. Alsabawi, K.; Webb, T.A.; Gray, E. MacA.; Webb, C.J. Kinetic enhancement of the sorption properties of MgH2 with the additive titanium isopropoxide. Int. J. Hydrogen Energy 2017, 42, 5227–5234. [Google Scholar] [CrossRef]
  83. Alsabawi, K.; Gray, E.M.A.; Webb, C.J. The effect of ball-milling gas environment on the sorption kinetics of MgH2 with/without additives for hydrogen storage. Int. J. Hydrogen Energy 2019, 44, 2976–2980. [Google Scholar] [CrossRef]
  84. Cortez, J.J.; Castro, F.J.; Troiani, H.E.; Pighin, S.A.; Urretavizcaya, G. Kinetic improvement of H2 absorption and desorption properties in Mg/MgH2 by using niobium ethoxide as additive. Int. J. Hydrogen Energy 2019, 44, 11961–11969. [Google Scholar] [CrossRef]
  85. Ojeda, X.A.; Castro, F.J.; Pighin, S.A.; Troiani, H.E.; Moreno, M.S.; Urretavizcaya, G. Hydrogen absorption and desorption properties of Mg/MgH2 with nanometric dispersion of small amounts of Nb(V) ethoxide. Int. J. Hydrogen Energy 2021, 46, 4126–4136. [Google Scholar] [CrossRef]
  86. Floriano, R.; Deledda, S.; Hauback, B.C.; Leiva, D.R.; Botta, W.J. Iron and niobium based additives in magnesium hydride: Microstructure and hydrogen storage properties. Int. J. Hydrogen Energy 2017, 42, 6810–6819. [Google Scholar] [CrossRef]
  87. Rizo-Acosta, P.; Cuevas, F.; Latroche, M. Optimization of TiH2 content for fast and efficient hydrogen cycling of MgH2-TiH2 nanocomposites. Int. J. Hydrogen Energy 2018, 43, 16774–16781. [Google Scholar] [CrossRef]
  88. Andrés, T.B.; Zélis Luis, M.; Marcos, M. Differences in the heterogeneous nature of hydriding/dehydriding kinetics of MgH2-TiH2 nanocomposites. Int. J. Hydrogen Energy 2020, 45, 27421–27433. [Google Scholar] [CrossRef]
  89. Zhang, Y.; Zhang, W.; Wei, X.; Yuan, Z.; Gao, J.; Guo, S.; Ren, H. Catalytic effects of TiO2 on hydrogen storage thermodynamics and kinetics of the as-milled Mg-based alloy. Mater. Charact. 2021, 176, 111118. [Google Scholar] [CrossRef]
  90. Valentoni, A.; Mulas, G.; Enzo, S.; Garroni, S. Remarkable hydrogen storage properties of MgH2 doped with VNbO5. Phys. Chem. Chem. Phys. 2018, 20, 4100–4108. [Google Scholar] [CrossRef]
  91. Silva, R.A.; Leal Neto, R.M.; Leiva, D.R.; Ishikawa, T.T.; Kiminami, C.S.; Jorge, A.M.; Botta, W.J. Room temperature hydrogen absorption by Mg and MgTiFe nanocomposites processed by high-energy ball milling. Int. J. Hydrogen Energy 2018, 43, 12251–12259. [Google Scholar] [CrossRef]
  92. Neto, R.M.L.; de Araújo Silva, R.; Floriano, R.; Coutinho, G.C.S.; Falcão, R.B.; Leiva, D.R.; Botta Filho, W.J. Synthesis by high-energy ball milling of MgH2-TiFe composites for hydrogen storage. Mater. Sci. Forum 2017, 899, 13–18. [Google Scholar] [CrossRef]
  93. El-Eskandarany, M.S.; Shaban, E.; Aldakheel, F.; Alkandary, A.; Behbehani, M.; Al-Saidi, M. Synthetic nanocomposite MgH2/5wt.% TiMn2 powders for solid-hydrogen storage tank integrated with PEM fuel cell. Sci. Rep. 2017, 7, 13296. [Google Scholar] [CrossRef] [Green Version]
  94. El-Eskandarany, M.S.; Shaban, E.; Al-Matrouk, H.; Behbehani, M.; Alkandary, A.; Aldakheel, F.; Ali, N.; Ahmed, S.A. Structure, morphology and hydrogen storage kinetics of nanocomposite MgH2/10wt% ZrNi5 powders. Mater. Today Energy 2017, 3, 60–71. [Google Scholar] [CrossRef]
  95. El-Eskandarany, M.S. Metallic glassy Ti2Ni grain-growth inhibitor powder for enhancing the hydrogenation/dehydrogenation kinetics of MgH2. RSC Adv. 2019, 9, 1036–1046. [Google Scholar] [CrossRef] [Green Version]
  96. El-Eskandarany, M.S.; Saeed, M.; Al-Nasrallah, E.; Al-Ajmi, F.; Banyan, M. Effect of LaNi3 amorphous alloy nanopowders on the performance and hydrogen storage properties of MgH2. Energies 2019, 12, 1005. [Google Scholar] [CrossRef] [Green Version]
  97. Lototskyy, M.; Goh, J.; Davids, M.W.; Linkov, V.; Khotseng, L.; Ntsendwana, B.; Denys, R.; Yartys, V.A. Nanostructured hydrogen storage materials prepared by high-energy reactive ball milling of magnesium and ferrovanadium. Int. J. Hydrogen Energy 2019, 44, 6687–6701. [Google Scholar] [CrossRef]
  98. Zhou, C.; Fang, Z.Z.; Sun, P.; Xu, L.; Liu, Y. Capturing low-pressure hydrogen using V-Ti-Cr catalyzed magnesium hydride. J. Power Sources 2019, 413, 139–147. [Google Scholar] [CrossRef]
  99. Dong, J.; Panda, S.; Zhu, W.; Zou, J.; Ding, W. Enhanced hydrogen sorption properties of MgH2 when doped with mechanically alloyed amorphous Zr0.67Ni0.33 particles. Int. J. Hydrogen Energy 2020, 45, 28144–28153. [Google Scholar] [CrossRef]
  100. Fu, Y.; Ding, Z.; Zhang, L.; Zhang, H.; Wang, W.; Li, Y.; Han, S. Catalytic effect of a novel MgC0.5Co3 compound on the dehydrogenation of MgH2. Prog. Nat. Sci. Mater. Int. 2021, 31, 264–269. [Google Scholar] [CrossRef]
  101. Pukazhselvan, D.; Bdikin, I.; Perez, J.; Carbó-Argibay, E.; Antunes, I.; Stroppa, D.G.; Fagg, D.P. Formation of Mg–Nb–O rock salt structures in a series of mechanochemically activated MgH2 + nNb2O5 (n = 0.083–1.50) mixtures. Int. J. Hydrogen Energy 2016, 41, 2677–2688. [Google Scholar] [CrossRef]
  102. Pukazhselvan, D.; Otero-Irurueta, G.; Pérez, J.; Singh, B.; Bdikin, I.; Singh, M.K.; Fagg, D.P. Crystal structure, phase stoichiometry and chemical environment of MgxNbyOx+y nanoparticles and their impact on hydrogen storage in MgH2. Int. J. Hydrogen Energy 2016, 41, 11709–11715. [Google Scholar] [CrossRef]
  103. Pukazhselvan, D.; Nasani, N.; Sandhya, K.S.; Singh, B.; Bdikin, I.; Koga, N.; Fagg, D.P. Role of chemical interaction between MgH2 and TiO2 additive on the hydrogen storage behavior of MgH2. Appl. Surf. Sci. 2017, 420, 740–745. [Google Scholar] [CrossRef]
  104. Pukazhselvan, D.; Nasani, N.; Correia, P.; Carbó-Argibay, E.; Otero-Irurueta, G.; Stroppa, D.G.; Fagg, D.P. Evolution of reduced Ti containing phase(s) in MgH2/TiO2 system and its effect on the hydrogen storage behavior of MgH2. J. Power Sources 2017, 362, 174–183. [Google Scholar] [CrossRef]
  105. Gattia, D.M.; Jangir, M.; Jain, I.P. Study on nanostructured MgH2 with Fe and its oxides for hydrogen storage applications. J. Alloys Compd. 2019, 801, 188–191. [Google Scholar] [CrossRef]
  106. Idris, N.H.; Mustafa, N.S.; Ismail, M. MnFe2O4 nanopowder synthesised via a simple hydrothermal method for promoting hydrogen sorption from MgH2. Int. J. Hydrogen Energy 2017, 42, 21114–21120. [Google Scholar] [CrossRef]
  107. Ismail, M.; Mustafa, N.S.; Ali, N.A.; Sazelee, N.A.; Yahya, M.S. The hydrogen storage properties and catalytic mechanism of the CuFe2O4-doped MgH2 composite system. Int. J. Hydrogen Energy 2019, 44, 318–324. [Google Scholar] [CrossRef]
  108. Ali, N.A.; Idris, N.H.; Din, M.F.M.; Yahya, M.S.; Ismail, M. Nanoflakes MgNiO2 synthesised via a simple hydrothermal method and its catalytic roles on the hydrogen sorption performance of MgH2. J. Alloys Compd. 2019, 796, 279–286. [Google Scholar] [CrossRef]
  109. Chen, M.; Xiao, X.; Zhang, M.; Liu, M.; Huang, X.; Zheng, J.; Zhang, Y.; Jiang, L.; Chen, L. Excellent synergistic catalytic mechanism of in-situ formed nanosized Mg2Ni and multiple valence titanium for improved hydrogen desorption properties of magnesium hydride. Int. J. Hydrogen Energy 2019, 44, 1750–1759. [Google Scholar] [CrossRef]
  110. Xie, X.; Chen, M.; Liu, P.; Shang, J.; Liu, T. High hydrogen desorption properties of Mg-based nanocomposite at moderate temperatures: The effects of multiple catalysts in situ formed by adding nickel sulfides/graphene. J. Power Sources 2017, 371, 112–118. [Google Scholar] [CrossRef]
  111. Yong, H.; Wang, S.; Ma, J.; Zhang, K.; Zhao, D.; Hu, J.; Zhang, Y. Dual-tuning of de/hydrogenation kinetic properties of Mg-based hydrogen storage alloy by building a Ni-/Co-multi-platform collaborative system. Int. J. Hydrogen Energy 2021, 46, 24202–24213. [Google Scholar] [CrossRef]
  112. Yao, P.; Jiang, Y.; Liu, Y.; Wu, C.; Chou, K.-C.; Lyu, T.; Li, Q. Catalytic effect of Ni@rGO on the hydrogen storage properties of MgH2. J. Magnes. Alloys 2020, 8, 461–471. [Google Scholar] [CrossRef]
  113. Fu, Y.; Ding, Z.; Ren, S.; Li, X.; Zhou, S.; Zhang, L.; Wang, W.; Wu, L.; Li, Y.; Han, S. Effect of in-situ formed Mg2Ni/Mg2NiH4 compounds on hydrogen storage performance of MgH2. Int. J. Hydrogen Energy 2020, 45, 28154–28162. [Google Scholar] [CrossRef]
  114. Huang, T.; Huang, X.; Hu, C.; Wang, J.; Liu, H.; Ma, Z.; Zou, J.; Ding, W. Enhancing hydrogen storage properties of MgH2 through addition of Ni/CoMoO4 nanorods. Mater. Today Energy 2021, 19, 100613. [Google Scholar] [CrossRef]
  115. Wang, S.; Gao, M.; Yao, Z.; Xian, K.; Wu, M.; Liu, Y.; Sun, W.; Pan, H. High-loading, ultrafine Ni nanoparticles dispersed on porous hollow carbon nanospheres for fast (de)hydrogenation kinetics of MgH2. J. Magnes. Alloys 2021. [Google Scholar] [CrossRef]
  116. Huang, X.; Xiao, X.; Wang, X.; Wang, C.; Fan, X.; Tang, Z.; Wang, C.; Wang, Q.; Chen, L. Synergistic catalytic activity of porous rod-like TMTiO3 (TM = Ni and Co) for reversible hydrogen storage of magnesium hydride. J. Phys. Chem. C 2018, 122, 27973–27982. [Google Scholar] [CrossRef]
  117. Zhao, Y.; Zhu, Y.; Liu, J.; Ma, Z.; Zhang, J.; Liu, Y.; Li, Y.; Li, L. Enhancing hydrogen storage properties of MgH2 by core-shell CoNi@C. J. Alloys Compd. 2021, 862, 158004. [Google Scholar] [CrossRef]
  118. Huang, X.; Xiao, X.; Zhang, W.; Fan, X.; Zhang, L.; Cheng, C.; Li, S.; Ge, H.; Wang, Q.; Chen, L. Transition metal (Co, Ni) nanoparticles wrapped with carbon and their superior catalytic activities for the reversible hydrogen storage of magnesium hydride. Phys. Chem. Chem. Phys. 2017, 19, 4019–4029. [Google Scholar] [CrossRef] [PubMed]
  119. Liu, J.; Ma, Z.; Liu, Z.; Tang, Q.; Zhu, Y.; Lin, H.; Zhang, Y.; Zhang, J.; Liu, Y.; Li, L. Synergistic effect of RGO supported Ni3Fe on hydrogen storage performance of MgH2. Int. J. Hydrogen Energy 2020, 45, 16622–16633. [Google Scholar] [CrossRef]
  120. Zeng, L.; Lan, Z.; Li, B.; Liang, H.; Wen, X.; Huang, X.; Tan, J.; Liu, H.; Zhou, W.; Guo, J. Facile synthesis of a Ni3S2@C composite using cation exchange resin as an efficient catalyst to improve the kinetic properties of MgH2. J. Magnes. Alloys 2021. [Google Scholar] [CrossRef]
  121. 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]
  122. Thiangviriya, S.; Plerdsranoy, P.; Hagenah, A.; Le, T.T.; Kidkhunthod, P.; Utke, O.; Dornheim, M.; Klassen, T.; Pistidda, C.; Utke, R. Effects of Ni-loading contents on dehydrogenation kinetics and reversibility of Mg2FeH6. Int. J. Hydrogen Energy 2021, 46, 32099–32109. [Google Scholar] [CrossRef]
  123. Wang, P.; Tian, Z.; Wang, Z.; Xia, C.; Yang, T.; Ou, X. Improved hydrogen storage properties of MgH2 using transition metal sulfides as catalyst. Int. J. Hydrogen Energy 2021, 46, 27107–27118. [Google Scholar] [CrossRef]
  124. Ismail, M.; Yahya, M.S.; Sazelee, N.A.; Ali, N.A.; Yap, F.A.H.; Mustafa, N.S. The effect of K2SiF6 on the MgH2 hydrogen storage properties. J. Magnes. Alloys 2020, 8, 832–840. [Google Scholar] [CrossRef]
  125. Yahya, M.S.; Sulaiman, N.N.; Mustafa, N.S.; Halim Yap, F.A.; Ismail, M. Improvement of hydrogen storage properties in MgH2 catalysed by K2NbF7. Int. J. Hydrogen Energy 2018, 43, 14532–14540. [Google Scholar] [CrossRef]
  126. Ismail, M. Effect of adding different percentages of HfCl4 on the hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2021, 46, 8621–8628. [Google Scholar] [CrossRef]
  127. Gao, S.; Wang, X.; Liu, H.; He, T.; Wang, Y.; Li, S.; Yan, M. Effects of nano-composites (FeB, FeB/CNTs) on hydrogen storage properties of MgH2. J. Power Sources 2019, 438, 227006. [Google Scholar] [CrossRef]
  128. Gao, S.; Wang, X.; Liu, H.; He, T.; Wang, Y.; Li, S.; Yan, M. CNTs decorated with CoFeB as a dopant to remarkably Improve the dehydrogenation/rehydrogenation performance and cyclic Stability of MgH2. Int. J. Hydrogen Energy 2020, 45, 28964–28973. [Google Scholar] [CrossRef]
  129. Gao, S.; Liu, H.; Xu, L.; Li, S.; Wang, X.; Yan, M. Hydrogen storage properties of nano-CoB/CNTs catalyzed MgH2. J. Alloys Compd. 2018, 735, 635–642. [Google Scholar] [CrossRef]
  130. Zhang, J.; Shi, R.; Zhu, Y.; Liu, Y.; Zhang, Y.; Li, S.; Li, L. Remarkable synergistic catalysis of Ni-doped ultrafine TiO2 on hydrogen sorption kinetics of MgH2. ACS Appl. Mater. Interfaces 2018, 10, 24975–24980. [Google Scholar] [CrossRef] [PubMed]
  131. Chen, M.; Xiao, X.; Zhang, M.; Zheng, J.; Liu, M.; Wang, X.; Jiang, L.; Chen, L. Highly dispersed metal nanoparticles on TiO2 acted as nano redox reactor and its synergistic catalysis on the hydrogen storage properties of magnesium hydride. Int. J. Hydrogen Energy 2019, 44, 15100–15109. [Google Scholar] [CrossRef]
  132. Gao, H.; Shi, R.; Zhu, J.; Liu, Y.; Shao, Y.; Zhu, Y.; Zhang, J.; Li, L.; Hu, X. Interface effect in sandwich like Ni/Ti3C2 catalysts on hydrogen storage performance of MgH2. Appl. Surf. Sci. 2021, 564, 150302. [Google Scholar] [CrossRef]
  133. Yao, L.; Lyu, X.; Zhang, J.; Liu, Y.; Zhu, Y.; Lin, H.; Zhang, Y.; Li, L. Remarkable synergistic effects of Mg2NiH4 and transition metal carbides (TiC, ZrC, WC) on enhancing the hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2020, 45, 6765–6779. [Google Scholar] [CrossRef]
  134. Wang, K.; Wu, G.; Cao, H.; Li, H.; Zhao, X. Improved reversible dehydrogenation properties of MgH2 by the synergetic effects of graphene oxide-based porous carbon and TiCl3. Int. J. Hydrogen Energy 2018, 43, 7440–7446. [Google Scholar] [CrossRef]
  135. Verma, S.K.; Bhatnagar, A.; Shukla, V.; Soni, P.K.; Pandey, A.P.; Yadav, T.P.; Srivastava, O.N. Multiple improvements of hydrogen sorption and their mechanism for MgH2 catalyzed through TiH2@Gr. Int. J. Hydrogen Energy 2020, 45, 19516–19530. [Google Scholar] [CrossRef]
  136. Singh, S.; Bhatnagar, A.; Shukla, V.; Vishwakarma, A.K.; Soni, P.K.; Verma, S.K.; Shaz, M.A.; Sinha, A.S.K.; Srivastava, O.N. Ternary transition metal alloy FeCoNi nanoparticles on graphene as new catalyst for hydrogen sorption in MgH2. Int. J. Hydrogen Energy 2020, 45, 774–786. [Google Scholar] [CrossRef]
  137. Huang, H.X.; Yuan, J.G.; Zhang, B.; Zhang, J.G.; Zhu, Y.F.; Li, L.Q.; Wu, Y.; Zhou, S.X. Improvement in the hydrogenation-dehydrogenation performance of a Mg–Al alloy by graphene supported Ni. Int. J. Hydrogen Energy 2020, 45, 798–808. [Google Scholar] [CrossRef]
  138. Huang, X.; Tao, A.; Guo, J.; Wei, W.; Guo, J.; Lan, Z. Synergistic effect of TiF3@ graphene on the hydrogen storage properties of Mg–Al alloy. Int. J. Hydrogen Energy 2018, 43, 1651–1657. [Google Scholar] [CrossRef]
  139. Wang, K.; Zhang, X.; Liu, Y.; Ren, Z.; Zhang, X.; Hu, J.; Gao, M.; Pan, H. Graphene-induced growth of N-doped niobium pentaoxide nanorods with high catalytic activity for hydrogen storage in MgH2. Chem. Eng. J. 2021, 406, 126831. [Google Scholar] [CrossRef]
  140. Yong, H.; Wei, X.; Hu, J.; Yuan, Z.; Guo, S.; Zhao, D.; Zhang, Y. Hydrogen storage behavior of Mg-based alloy catalyzed by carbon-cobalt composites. J. Magnes. Alloys 2021. [Google Scholar] [CrossRef]
  141. Du, J.; Lan, Z.; Zhang, H.; Lü, S.; Liu, H.; Guo, J. Catalytic enhanced hydrogen storage properties of Mg-based alloy by the addition of reduced graphene oxide supported V2O3 nanocomposite. J. Alloys Compd. 2019, 802, 660–667. [Google Scholar] [CrossRef]
  142. Hou, X.; Hu, R.; Zhang, T.; Kou, H.; Li, J. Hydrogenation thermodynamics of melt-spun magnesium rich Mg–Ni nanocrystalline alloys with the addition of multiwalled carbon nanotubes and TiF3. J. Power Sources 2016, 306, 437–447. [Google Scholar] [CrossRef]
  143. Chuang, Y.-S.; Hwang, S.-J. Synthesis and hydrogen absorption/desorption properties of Mg–Nb2O5–SWCNT/MWCNT nanocomposite prepared by reactive milling. J. Alloys Compd. 2016, 656, 835–842. [Google Scholar] [CrossRef]
  144. Gajdics, M.; Spassov, T.; Kis, V.K.; Schafler, E.; Révész, Á. Microstructural and morphological investigations on Mg-Nb2O5-CNT nanocomposites processed by high-pressure torsion for hydrogen storage applications. Int. J. Hydrogen Energy 2020, 45, 7917–7928. [Google Scholar] [CrossRef]
  145. Lu, X.; Zhang, L.; Yu, H.; Lu, Z.; He, J.; Zheng, J.; Wu, F.; Chen, L. Achieving superior hydrogen storage properties of MgH2 by the effect of TiFe and carbon nanotubes. Chem. Eng. J. 2021, 422, 130101. [Google Scholar] [CrossRef]
  146. Zhang, M.; Xiao, X.; Luo, B.; Liu, M.; Chen, M.; Chen, L. Superior de/hydrogenation performances of MgH2 catalyzed by 3D flower-like TiO2@C nanostructures. J. Energy Chem. 2020, 46, 191–198. [Google Scholar] [CrossRef]
  147. Zhang, L.; Cai, Z.; Zhu, X.; Yao, Z.; Sun, Z.; Ji, L.; Yan, N.; Xiao, B.; Chen, L. Two-dimensional ZrCo nanosheets as highly effective catalyst for hydrogen storage in MgH2. J. Alloys Compd. 2019, 805, 295–302. [Google Scholar] [CrossRef]
  148. Révész, Á.; Spassov, T.; Kis, V.K.; Schafler, E.; Gajdics, M. The influence of preparation conditions on the hydrogen sorption of Mg-Nb2 O5-CNT produced by ball milling and subsequent high-pressure torsion. J. Nanosci. Nanotechnol. 2020, 20, 4587–4590. [Google Scholar] [CrossRef] [PubMed]
  149. Gajdics, M.; Spassov, T.; Kovács Kis, V.; Béke, F.; Novák, Z.; Schafler, E.; Révész, Á. Microstructural Investigation of nanocrystalline hydrogen-storing Mg-titanate nanotube composites processed by high-pressure torsion. Energies 2020, 13, 563. [Google Scholar] [CrossRef] [Green Version]
  150. Meng, Q.; Huang, Y.; Ye, J.; Xia, G.; Wang, G.; Dong, L.; Yang, Z.; Yu, X. Electrospun carbon nanofibers with in-situ encapsulated Ni nanoparticles as catalyst for enhanced hydrogen storage of MgH2. J. Alloys Compd. 2021, 851, 156874. [Google Scholar] [CrossRef]
Figure 1. Mg–H binary phase diagram [36].
Figure 1. Mg–H binary phase diagram [36].
Energies 14 06400 g001
Figure 2. Hydrogen absorption pressure-composition isotherms for a typical metal–hydrogen system, together with the Van’t Hoff plot [38].
Figure 2. Hydrogen absorption pressure-composition isotherms for a typical metal–hydrogen system, together with the Van’t Hoff plot [38].
Energies 14 06400 g002
Figure 3. Temperature dependence of the equilibrium pressure for the Mg↔MgH2 reaction [39].
Figure 3. Temperature dependence of the equilibrium pressure for the Mg↔MgH2 reaction [39].
Energies 14 06400 g003
Figure 4. Schematic illustration and working principle of a (left) SPEX 8000 vibratory mill and (right) Fritsch planetary mill [67].
Figure 4. Schematic illustration and working principle of a (left) SPEX 8000 vibratory mill and (right) Fritsch planetary mill [67].
Energies 14 06400 g004
Figure 5. Illustration of the deformation of powder agglomerate during an impact process [68] (left) porous powder particles before the impact with the milling ball, (center) powder particles are compressed by the ball, (right) elastic and plastic deformation takes place.
Figure 5. Illustration of the deformation of powder agglomerate during an impact process [68] (left) porous powder particles before the impact with the milling ball, (center) powder particles are compressed by the ball, (right) elastic and plastic deformation takes place.
Energies 14 06400 g005
Figure 6. Schematic illustration of hydrogen dissociation and recombination on the TiO2 sheet [79].
Figure 6. Schematic illustration of hydrogen dissociation and recombination on the TiO2 sheet [79].
Energies 14 06400 g006
Figure 7. Reversible and irreversible contributions to the hydrogen storage of the (1–y) MgH2 + y TiH2 nanocomposite [87].
Figure 7. Reversible and irreversible contributions to the hydrogen storage of the (1–y) MgH2 + y TiH2 nanocomposite [87].
Energies 14 06400 g007
Figure 8. (a) HRTEM image of MgH2/10 wt.% ZrNi5 nanocomposite powders obtained after 50 h of ball milling time. The FFT lattice images for zones I, II, III, IV, V and VI shown in (a) are displayed in (bg), respectively [94].
Figure 8. (a) HRTEM image of MgH2/10 wt.% ZrNi5 nanocomposite powders obtained after 50 h of ball milling time. The FFT lattice images for zones I, II, III, IV, V and VI shown in (a) are displayed in (bg), respectively [94].
Energies 14 06400 g008
Figure 9. XRD profiles of 10 wt.% CuFe2O4-doped MgH2 sample (a) after ball milling, (b) after desorption at 450 °C and (c) after re-absorption at 250 °C [107].
Figure 9. XRD profiles of 10 wt.% CuFe2O4-doped MgH2 sample (a) after ball milling, (b) after desorption at 450 °C and (c) after re-absorption at 250 °C [107].
Energies 14 06400 g009
Figure 10. (a) DSC and (b) TPD curves of MgH2 doped by different catalysts. Note the MgH2–Ni–TiO2 is a simple co-milling of the additives, while Ni/TiO2 was prepared using the solvothermal method [109].
Figure 10. (a) DSC and (b) TPD curves of MgH2 doped by different catalysts. Note the MgH2–Ni–TiO2 is a simple co-milling of the additives, while Ni/TiO2 was prepared using the solvothermal method [109].
Energies 14 06400 g010
Figure 11. Schematic illustration of the synergistic catalytic effect of in situ formed phases in TM-MOF (TM = Fe, Ni) doped MgH2 [121].
Figure 11. Schematic illustration of the synergistic catalytic effect of in situ formed phases in TM-MOF (TM = Fe, Ni) doped MgH2 [121].
Energies 14 06400 g011
Figure 12. Schematic representation of synergistic effect of rGO and in situ formed phases during (de)hydrogenation of MgH2 [119].
Figure 12. Schematic representation of synergistic effect of rGO and in situ formed phases during (de)hydrogenation of MgH2 [119].
Energies 14 06400 g012
Figure 13. (left) SEM image of flower-like TiO2@C, and (right) schematic diagram of the morphological changes during the milling process [146].
Figure 13. (left) SEM image of flower-like TiO2@C, and (right) schematic diagram of the morphological changes during the milling process [146].
Energies 14 06400 g013
Figure 14. (a) Transmission electron micrographs of Na2Ti3O7 nanorods and (b) nanotubes with the corresponding selected area electron diffraction patterns [43].
Figure 14. (a) Transmission electron micrographs of Na2Ti3O7 nanorods and (b) nanotubes with the corresponding selected area electron diffraction patterns [43].
Energies 14 06400 g014
Table 1. Absorption/desorption properties and desorption activation energy of a few Mg-based hydrogen storage systems.
Table 1. Absorption/desorption properties and desorption activation energy of a few Mg-based hydrogen storage systems.
MaterialAbsorption PerformanceDesorption PerformanceDesorption Activation Energy (kJ/mol)Ref.
MgH2 + 10 wt.% TiVO3.56.3 wt.%/200 s 6.8 wt.%/30 min 62.4[41]
200 °C/1 MPa300 °C
MgH2 + 5 wt.% Na2Ti3O7 nanotube6 wt.%/60 s 6.5 wt.%/6 min 70.4[43]
275 °C/3.2 MPa300 °C/3 kPa
MgH2 + 5 wt.% K2Ti8O17 nanobelts6.2 wt.%/12.6 min 6.6 wt.%/2.4 min 116.3[47]
100 °C/2.24 MPa280 °C
MgH2 + 5 wt.% TiO26.4 w.t%/250 s6.5 wt.%/1000 s 76.1[79]
200 °C/3 MPa300 °C/0.005 MPa
MgH2 + 1mol.% Nb(V) ethoxide5.5 wt.%/20 min 5.2 wt.%/10 min 75[84]
300 °C/1 MPa300 °C/20 kPa
La7Ce3Mg80Ni10 + 5 wt.% TiO24 wt.%/45 s 3 wt.%/172 s 57.4[89]
200 °C/3 MPa300 °C/0.0001 MPa
MgH2 + 10 wt.% ZrNi55.3 wt.%/30 min 5.3 wt.%/15 min 110[94]
250 °C/0.8 MPa275 °C/0.01 MPa
MgH2 + 10 wt.% Ti2Ni5.1 wt.%/100 s 5.7 wt.%/333 s 87.3[95]
225 °C/1 MPa225 °C/0.02 MPa
MgH2 + 10 wt.% MnFe2O45.6 wt.%/10 min 4.6 wt.%/10 min 108.4[106]
200 °C/3 MPa320 °C/0.1 MPa
Mg + 5 wt.% NiS/rGO5 wt.%/10 min 4.5 wt.%/60 min 63[110]
300 °C/4 MPa300 °C/100 Pa
MgH2 + 10 wt.% NiMoO4 nanorod5.5 wt.%/10 min 6 wt.%/10 min 85.9[114]
150 °C/3.2 MPa300 °C
MgH2 + 8 wt.% CoNi@C6 wt.%/200 s 6.17 wt.%/1800 s 78.5[117]
150 °C/3 MPa300 °C/0.005 MPa
MgH2 + 5 wt.% K2SiF64.5 wt.%/2 min 5.1 wt.%/30 min 114[124]
250 °C/3 MPa320 °C/0.1 MPa
MgH2 + 10 wt.% CoFeB/CNT6.2 wt.%/10 min 6.5 wt.%/30 min 83.2[128]
150 °C/5 MPa300 °C
MgH2 + 5 wt.% Ni/Ti3C25.6 wt.%/50 s 6.73 wt.%/2400 s 91.6[132]
200 °C/3 MPa300 °C/0.005 MPa
Mg95Ni5 + 5 wt.% TiC4.44 wt.%/1800 s 4.73 wt.%/1800 s 74.1[133]
150 °C/3 MPa250 °C/0.005 MPa
MgH2 + 5 wt.% FeCoNi@GS6.01 wt.%/1.65 min 6.14 wt.%/8.5 min 85.4[136]
290 °C/1.5 MPa290 °C/0.1 MPa
MgH2 + 10 wt.% N-Nb2O5@C6.2 wt.%/60 min 6.2 wt.%/60 min 81[139]
100 °C/5 MPa225 °C
Mg90Ce5Y5 + 10 wt.% C@Co4.5 wt.%/100 min 4.5 wt.%/11 min 81.9[140]
200 °C/3 MPa300 °C/0.005 MPa
MgH2 + 5 wt.% flower-like TiO2/C6 wt.%/40 min 6 wt.%/7 min 67.1[146]
150 °C/5 MPa250 °C/1 kPa
MgH2 + 10 wt.% ZrCo nanosheet4.4 wt.%/10 min 5.3 wt.%/5 min 90.4[147]
120 °C/3 MPa300 °C
MgH2 + 10 wt.% Ni@C6 wt.%/800 s 6 wt.%/3500 s 93.1[150]
300 °C/3 MPa300 °C
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Révész, Á.; Gajdics, M. Improved H-Storage Performance of Novel Mg-Based Nanocomposites Prepared by High-Energy Ball Milling: A Review. Energies 2021, 14, 6400. https://doi.org/10.3390/en14196400

AMA Style

Révész Á, Gajdics M. Improved H-Storage Performance of Novel Mg-Based Nanocomposites Prepared by High-Energy Ball Milling: A Review. Energies. 2021; 14(19):6400. https://doi.org/10.3390/en14196400

Chicago/Turabian Style

Révész, Ádám, and Marcell Gajdics. 2021. "Improved H-Storage Performance of Novel Mg-Based Nanocomposites Prepared by High-Energy Ball Milling: A Review" Energies 14, no. 19: 6400. https://doi.org/10.3390/en14196400

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