Next Article in Journal
Advanced Process Simulation of Low Pressure Die Cast A356 Aluminum Automotive Wheels—Part I, Process Characterization
Next Article in Special Issue
A Novel Non-Equiatomic (W35Ta35Mo15Nb15)95Ni5 Refractory High Entropy Alloy with High Density Fabricated by Powder Metallurgical Process
Previous Article in Journal
Prediction of Mechanical Properties for High Strength Low Alloyed Steels in a Commercial Hot Dip Galvanizing Line without Soaking Section
Previous Article in Special Issue
Mechanical Amorphization and Recrystallization of Mn-Co(Fe)-Ge(Si) Compositions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Review on Hydrogen Absorption Behaviour of Metal Alloys Prepared through Mechanical Alloying

by
Thabang Ronny Somo
1,2,
Thabiso Carol Maponya
1,
Moegamat Wafeeq Davids
2,
Mpitloane Joseph Hato
1,*,
Mykhaylo Volodymyrovich Lototskyy
2 and
Kwena Desmond Modibane
1,*
1
Nanotechnology Research Lab, Department of Chemistry, School of Physical and Mineral Sciences, University of Limpopo (Turfloop), Polokwane 0727, South Africa
2
Hydrogen South Africa (HySA), South African Institute of Advanced Chemistry, University of the Western Cape, Cape Town 7535, South Africa
*
Authors to whom correspondence should be addressed.
Metals 2020, 10(5), 562; https://doi.org/10.3390/met10050562
Submission received: 21 March 2020 / Revised: 14 April 2020 / Accepted: 18 April 2020 / Published: 26 April 2020
(This article belongs to the Special Issue Mechanical Alloying: Processing and Materials)

Abstract

:
Hydride-forming alloys are currently considered reliable and suitable hydrogen storage materials because of their relatively high volumetric densities, and reversible H2 absorption/desorption kinetics, with high storage capacity. Nonetheless, their practical use is obstructed by several factors, including deterioration and slow hydrogen absorption/desorption kinetics resulting from the surface chemical action of gas impurities. Lately, common strategies, such as spark plasma sintering, mechanical alloying, melt spinning, surface modification and alloying with other elements have been exploited, in order to overcome kinetic barriers. Through these techniques, improvements in hydriding kinetics has been achieved, however, it is still far from that required in practical application. In this review, we provide a critical overview on the effect of mechanical alloying of various metal hydrides (MHs), ranging from binary hydrides (CaH2, MgH2, etc) to ternary hydrides (examples being Ti-Mn-N and Ca-La-Mg-based systems), that are used in solid-state hydrogen storage, while we also deliver comparative study on how the aforementioned alloy preparation techniques affect H2 absorption/desorption kinetics of different MHs. Comparisons have been made on the resultant material phases attained by mechanical alloying with those of melt spinning and spark plasma sintering techniques. The reaction mechanism, surface modification techniques and hydrogen storage properties of these various MHs were discussed in detail. We also discussed the remaining challenges and proposed some suggestions to the emerging research of MHs. Based on the findings obtained in this review, the combination of two or more compatible techniques, e.g., synthesis of metal alloy materials through mechanical alloying followed by surface modification (metal deposition, metal-metal co-deposition or fluorination), may provide better hydriding kinetics.

1. Introduction

The physical storage of hydrogen through compressed H2 tanks, liquid H2 tanks and cryo-compressed tanks has been condemned, due to several limitations, including the high pressure required for compression of the hydrogen gas and a large amount of energy for liquefaction of hydrogen [1]. A convenient way to store hydrogen is through the absorption of hydrogen onto solid nanostructured materials such as metal alloys, chemical sorbents (metal organic frameworks (MOFs)) and carbon-based materials) or chemical hydrides [2]. Out of the above-mentioned chemical storage materials, metal hydrides offer a safe, volume-efficient technology aimed at hydrogen storage for reversible on-board applications due to their large hydrogen uptake capacity (7.6 wt. % by Mg-type alloys) [3,4]. There are excellent reviews, regarding MHs, prepared by equilibrium procedure for hydrogen storage application that have been published [4,5,6]. However, there are not a lot of reviews on the MHs prepared by non-equilibrium procedure.
It is well-documented that metallic hydrides are chemically formed from the reaction of hydrogen gas with metal or alloy according to Equation (1) [1]:
Metal (M) + H2 (g) → metal hydride (MH).
The formation of metal hydride undergoes six distinct states of hydrogen absorption in metals: Hydrogen to metal; physisorption, which involves physical attachment of hydrogen to a metal through van der Waals interaction; chemisorption, the chemical attachment of hydrogen to metal by forming a chemical bond; subsurface hydrogen; solid solution (α-phase) and hydride (β-phase) [7]. The formation of α- and β-phases may be described by Equations (2) and (3), respectively [8]. An α-phase is formed when small amount of hydrogen occupies the interstitial sites of the host metal M, producing an H-poor solid solution, while β-phase involves saturation of the solid solution to generate a second phase:
M + 1 2 y H 2     MH y
MH y + 1 2   ( x y )   H 2     MH x .
Where M represents metal and H represents hydrogen atoms. During the formation of β-phase, the equilibrium pressure is fixed at any given temperature. According to Gibbs’ phase rule (F = C – P + 2, in which F is degrees of freedom, C represent the components of a reaction and P is the number of phases), the number of degrees of freedom, F, is one during the interaction of metal (M) and hydrogen (H) atoms since three phases P (two solid and one gaseous) and two components C (H and M) are available [8]. The constant pressure level is referred to as the plateau pressure. During the formation, the metal atoms stretch apart to accommodate the hydrogen atoms and the physical structure of the metal atoms may also change to form a metal hydride. For example, Table 1 shows different types of metal alloys ranging from AB to AB5, demonstrating that hydrogenation of metal is accompanied by increase in lattice parameters of the parent alloy [9,10,11,12,13]. The A2B alloys were reported to possess higher hydrogen storage capacity of 3.75 and 3.04 wt. % for P6222, and Fddd phase, respectively [11]. On the other hand, AB5 showed to have low absorption capacity of 1.43 wt. % [10]. Moreover, peaks from X-ray diffractogram may broaden and shift towards smaller 2θ values as a result of larger lattice parameters [9].
Conversely, metal hydrides release H2 gas when heat is applied on the hydride. This process is called desorption. Hydrogen atoms (H) move from the metal to the surface of the material, combine into hydrogen molecules (H2) and flow away as hydrogen gas. The metal atoms contract to form its original metal crystal structure [14]. The metal alloys in Table 1 can easily be synthesized by a well-known equilibrium procedure called arc or induction melting [15,16]. However, alloys prepared through this method suffer from poor hydrogenation performances and poisoning intolerance [17]. Recent findings reported that a convenient way to overcome some of these challenges is to prepare metal alloys using mechanical alloying (MA) through high-energy ball milling technique [15,18]. In this review, we provide the recent developments on hydrogenation performances of metal alloys prepared by mechanical alloying technique, comparison of MA with other non-equilibrium techniques. Moreover, we also consider the effect of combining MA with some surface modification technologies. We anticipate that this work can provide an inspiring perspective for further research on MHs with good hydrogen storage properties.

2. Non-Equilibrium Preparation Techniques

Mechanical, kinetic and thermodynamic performances related to the formation of metal hydride are affected by structure and phases of metal or metal alloy of interest. It is currently well-known that the structure and phase of alloy materials can be controlled by synthesizing them under non-equilibrium conditions [19]. There are several processes/techniques, which are in commercial use for the synthesis of alloy materials under such conditions. Amongst such techniques, much attention has been given to mechanical alloying, melt spinning and spark plasma sintering techniques [20]. According to Scudino et al. [21], the general underlying idea in all these techniques is to synthesize alloy materials in a non-equilibrium state by “energizing and quenching”. The energizing process involves the use of external dynamical force, such as melting; evaporation; irradiation; application of pressure; or storing of mechanical energy by plastic deformation on solid-state materials to put them onto a non-equilibrium state [22]. During this process, the solid-state materials change in phase to gas or liquid. In “quenching”, a subsequent step after “energization”, the material of interest is extinguished into a frozen state, and it is at this point where the desired structure or phase can be easily controlled [23,24]. The alloys synthesized in this sequence possess particular structural characteristics that are difficult or rare to attain by equilibrium preparation methods: Nano-crystalline grain size with a high density of grain boundaries and even lacking of long-range order (amorphous state). These structures alloy quick hydrogen uptake and excellent cycle-life behaviour [25]. Among all preparation methods that follow “energizing and quenching’’ mechanism, MA has been given much attention due to its user-friendly, comparatively less costly machinery and applicability to various metal alloys [19]. In this regard, mechanical alloying experimental technique is discussed with the point of interest being its effect on hydrogenation performances of metal alloy materials.

3. Mechanical Alloying

3.1. Background

Mechanical alloying, also known as ball milling, is a solid-state powder processing technique involving repeated welding, fracturing, and rewelding of powder particles in a high-energy ball mill [19]. Repetitions of welding, as well as fracturing of particles, generate a lamellar structure within the powders. MA method is usually applied when synthesizing highly immiscible phases in a non-equilibrium state [26]. The immiscible phases that are prepared, include super-saturated solid solutions, non-equilibrium alloys that are usually amorphous in nature, metastable crystalline and quasicrystalline phases. The procedure of MA begins with blending of the powders in the correct extent and stacking the powder blend into the mill along with the grinding medium. This blend is then milled for the ideal time allotment until a consistent state is achieved when the arrangement of each powder particle is equivalent to the extent of the elements in the initial powder blend. The milled powder is then united into a mass shape and exposed to heat, in order to get the ideal phase and structure. Thus, the vital components of MA procedure are the raw materials, mill and procedure parameters. This procedure was described by Shang et al. [27]. Ares et al. [25] observed that diffraction peaks from the X-ray diffractogram (XRD) of Mm (Ni, Mn, Al, Co)5.2 became broader with milling time. An example of MA process is depicted in Figure 1 (synthesis of Al–5wt. %Fe–4wt. %Mn blend), with scanning electron microscopic (SEM) images provided for all the steps. The double mechanical milling, as observed in the figure is necessary to refine the intermetallics and decrease the particle size of the blend [19]. MA has also been utilized to synthesize AB (FeTi), A2B (Mg2Ni) and AB5 (binary LaNi5) alloys [28]. In most cases, the rate of hydrogenation has been greatly enhanced when compared with that of materials prepared through conventional equilibrium methods, but due to the amorphous nature of milled alloys, a loss of hydrogen content was observed [25]. MA provides various advantages, including smaller restrictions with respect to composition and works much better when applied to A2B alloys [29]. Moreover, through the use of MA method, nano-crystalline alloys are obtained, starting either with a pre-prepared alloy or with a mixture of pure elemental metals [28]. Nano-crystalline alloys that are prepared through MA technique possess better activation compared to their polycrystalline counterparts [28]. For example, Polycrystalline FeTi is activated by heating up to temperatures above 673 K for one to two hours followed by annealing at 7 bar pressure and its hydrogenation requires a high pressure of 40–70 bar [30]. However, with a mechanically-alloyed nano-crystalline FeTi alloy, a single vacuum heating at 673 K for 30 min is sufficient to obtain full hydrogen absorption cycles [30].

3.2. Mg-Based Alloys

3.2.1. Structural Characteristics of Mechanically Milled Mg-Based Alloys

The crystal structure of a metallic Mg is hexagonal with lattice parameters a = 3.2094 Å and c = 5.2108 Å and a space group of P63/mmc [31]. Formation of MgH2 hydride consists of two steps, i.e., preparation of Mg nanoparticles, followed by hydrogen absorption to form the metal hydride. A traditional method for preparing Mg nanoparticles is gas-phase condensation method, which is much better discussed in [32]. As depicted by Figure 2a, hydrogenation of Mg nanoparticles tempers with XRD diffraction pattern of the parent nanoparticles [33]. Diffraction peaks at 2θ ≈ 48°, 34° and 32° disappear, while several new peaks develop, forming tetragonal magnesium hydride (β-MgH2) phase [33]. To illustrate the difference between traditionally synthesized MgH2 hydride and those synthesized through mechanical/ball milling, XRD diffraction patterns of MgH2 and as-milled hydrided Mg-Al alloy are compared in Figure 2a,b, respectively. For as-milled Mg-Al alloy, Li et al. [31] reported four stable phases: the fcc solid solution of magnesium in aluminum, the hcp solid solution of aluminum in magnesium, the β-phase (Mg2Al3) and the γ-phase (Mg17Al12). Such rewarding characteristic gives rise to excellent hydriding behavior of as-milled Mg-based alloys.
In general, the morphology and particle size distribution are also important structural parameters, carrying a huge responsibility for surface physisorption of molecular hydrogen (H2), surface dissociative chemisorption, surface absorption, hydrogen diffusion and phase transformation [34]. These two parameters are well-interpreted using scanning electron microscopic (SEM) images. For Mg-based alloys, mechanically milled alloys show faster hydrogen absorption kinetics and larger uptake capacity, compared to their unmilled counterparts and one of the effective factors contributing to this, is the reduction of particle sizes of the parent materials during mechanical alloying [35]. According to the literature, small particle sizes allow easy mobility of molecular hydrogen and quick dissociation into atomic hydrogen on the surface of a material [33]. As shown in Figure 3, all the five SEM images of milled Mg-based alloys constitute of smaller particles than that of traditionally sysnthesized MgH2, [33]. Therefore, their hydrogen uptake performance is expected to be superior to that of MgH2. This is further discussed in the next section.

3.2.2. Hydrogen Absorption Behaviour of Mechanically Alloyed Mg-Based Alloys

Mg-based alloys are reported to possess energy density of 9 MJ/kg and the highest maximum hydrogen storage capacity (7.6 and 3.6 wt. % for MgH2, and Mg2NiH4, respectively) compared to all hydride-forming materials [40,41]. The hydride stability of this alloy is very high, and therefore, a high dissociation temperature of around 561 K is required to transform MgH2 hydride into Mg and H2 gas [42]. Furthermore, the hydriding kinetics of MgH2 hydride is relatively slow because the dissociation of molecular hydrogen into hydrogen atoms is difficult on the surface of Mg metal [43]. Several reports are found in the literature concerning synthesis and hydrogen properties of Mg-based alloys synthesized by mechanical alloying [3,44,45,46]. Another interesting study is Andreasen [31]. The authors’work was based on the comparison of hydrogen kinetics, presented in Figure 4a, of un-milled Mg and those of ball milled Mg-Al compounds with varying Al content [31]. The kinetics increased gradually with an increase in Al content due to easy dissociation of hydrogen molecules on the surface of Al, as a result of negligible activation barrier for hydrogen dissociation for this metal. It is, therefore, understood that hydrogen molecules split into atomic hydrogen on the surface of Al before diffusing and attaching on the active sites of Mg, which has higher hydrogen sticking ability compared to Al [31]. Lototskyy et al. [37] reported the effect of synthesizing different Mg-based composites varying the amounts of multi-walled carbon nanotubes, activated carbon, FeTiO3 and graphite using mechanical alloying at a rotation speed of 500 rpm for a period of 120 min on hydrogenation behaviour of Mg, Figure 4b. The authors discovered that hydrogenation behaviour (both the uptake capacity and kinetics) of Mg mechanically alloyed with carbon-based materials are exceptional. Ref. [46] reveals that the better hydrogenation behaviour possessed by milled alloys can be attributed to their combination of tetragonal (a = 4.52 Å, c = 3.02 Å) and orthorhombic (a = 4.48 Å, b = 5.40 Å, c = 4.90 Å) phases as major phases, since these two phases are associated with small particle size and a presence of strain in the particles. Similar trends of hydrogen capacity and kinetics were observed by Lototskyy et al. [47] and Williams et al. [48] when they assessed Mg-10(FeV)-5MWCNT and Mg-5Pd composites through mechanical alloying, respectively. Table 2 and Table 3 represent a compiled data of literature based on hydrogenation behaviour of Mg-based alloy materials synthesized through ball milling and mechanical alloying.

3.2.3. Ball-Milled MgH2 Doped with Catalytic Additives

In the previous section, much attention is given to mechanically milled Mg-based alloys where a pure milled/un-milled Mg metal is admixed with other several materials, such as FeTiO3 before their interaction with hydrogen is subsequently explored. Here, we discuss the influence of catalytic effect of doping ball milled MgH2 with different catalystic additives. Previous experiments indicate clearly that admixing, deposition or doping of catalysts and additives could greatly enhance thermodynamic property and de/rehydrogenation kinetic parameters of MgH2. In these studies, metal [49], metal oxide [50,51,52], metal halides [36,53], metal hydride [49], carbon materials [54] and nanosized alloys [55,56] were used as the catalysts and additives. Table 2 summarizes thermodynamic and kinetic effects of these catalysts. The materials in the table were prepared through ball milling technique.
We can conclude from the table that doped MgH2 materials, prepared through mechanical milling, enjoy superior hydrogenation kinetics, compared to undoped materials. However, this particular excellent improvement significantly reduces the maximum absorption capacity of the system. For metal halide family, particularly MgH2+10wt. %CeCl3, the loss of capacity can be related to in situ formation of CeMg alloy and MgCl2, which seem to be good catalysts for faster absorption but poor gas absorbers [53]. On the other hand, drastic decrease of absorption capacity revealed by MgH2+SWNT composite is suggested to come from structure destruction of the SWNT during milling [54]. Table 2 also shows that composites which resulted in major decrease in hydrogen capacity are those with a dopant of 10 wt. % or more. We therefore suggest preparation of MgH2 composites with dopants that are less than a concentration of 10 wt. %. Moreover, milling period should also be taken into consideration when these materials are formed, more especially when carbon materials are in abundance as they seem to be sensitive to milling conditions. The difference between enthalpy of desorption (ΔHdes) of the materials is negligible and as thus thermodynamic property of MgH2 remain unchanged upon doping.

3.3. AB5-Type Alloys

3.3.1. Structural Characteristics of Mechanically Milled AB5-Type Alloys

Another interesting class of metal alloys is AB5. The A site in AB5 represents one or more strongly hydride-forming elements while B site represents one or more weakly hydride-forming elements, and usually helps to dissociate the H2 molecules on the surface of alloy material [58]. The A site is usually occupied by La, Ca or rare earth element while Ni, Cu, Co, Pt or Fe fill the B site [59]. AB5 alloys are known to have a CuCa5-type hexagonal crystal structure which normally belongs to space group P6/mmm (#191) [60,61]. Singh et al. [62] evaluated the effect of milling MmNi4:6Fe0:4 alloy, where Mm refers to mischmetal, varying the milling period from 10, 20 and 30 min, at a milling speed of 200 rpm. As shown by Figure 5, mechanical milling of this alloy results in slight shift of diffraction peaks towards lower 2θ values, reflecting increase in lattice parameters, and peak width broadening. Peak width broadening is a reflection of decrease in particle size. In general, the lattice structure of MmNi5-type alloy consists of two stacked planes at ɀ = 0 and ɀ = ½, and the first plane contains 1a and 2c, while the latter plane contains 3g crystallographic sites. Both Mm and Ni atoms occupy the 1a and 2c sites, while the 3g sites are occupied by Ni atoms only. Therefore in the presence of Fe to form MmNi4:6Fe0:4 alloy, Fe atoms replace Ni on the 3g sites [63]. Electronic configuration of Fe and that of Ni are very close (3d64s2 for Fe and 3d84s2 for Ni) and as a result covalent character, which is stronger than ionic character, is expected in the Ni/Fe-bearing planes. Thus, during mechanical milling of an alloy material with such character, strained platelet shaped particles are expected to form [64]. Such particles allow smooth movement of hydrogen during sorption better than fine powdered particles. This statement is supported by Table 3 as it is observed from the table that materials that were milled for longer period of time are associated with lower hydrogen uptake capacity compared to their counterparts that were milled for shorter period. This is due to the mechanical energy that is being pumped into the alloy matrix during milling, which breaks the particles, while forming strain in the matrix, and thus, for longer milling periods fractured and much finer powders, which are not recommended may result [63,64].
To be more accurate, not only particle size gives precise measurement of hydrogen uptake capacity by a material, but it also provides the nature of agglomeration, as well as homogeneity of the particles [62]. Short milling process yields non-homogeneous microstructural (wide distribution of nano-crystallite sizes and amorphous phase) and morphological alloy powders [25]. On the other hand, Parker et al. [35] recommends defects, fractures and porosity as faster absorption kinetics agents. However, it is very challenging to fully release hydrogen from a porous and fractured material. Figure 6 depicts SEM images of un-milled MmNi5 alloy and those of MmNi5 milled with different transition metals at concentration of 2.0 wt. %. for un-milled alloy, its structure lacks strain and consists of very large particles, and thus, hydrogen diffusion is difficult for such material [67].
In contrast to MmNi4:6Fe0:4 alloy, neither covalent nor ionic bonding is expected between the bulk alloy MmNi5 and the added transition metals (Co, Ni, Mn and Fe), since there was no substitution or melting during mechanical milling, thus, a physical attachment was involved [67]. As a result, the extent to which the particles of each alloy crush and become reduced in size during milling depends largely on ductility and hardness of the transition metal used. According to Mohs scale of metal hardness, Mn has Mohs hardness of 6, followed by Co with hardness of 5 and the last two are Fe and Ni both with hardness of 4.5 and 4, respectively [68]. Therefore, during ball milling, it is difficult to fracture and form finer powders with Mn particles than with Ni particles.

3.3.2. Hydrogen Absorption Behaviour of Mechanically Alloyed AB5-Type Alloys

These metal hydride alloys are typified by LaNi5 and they can be easily activated for hydrogen absorption [69]. A reaction between these alloys and hydrogen at a room temperature and pressure above equilibrium plateau pressure results in a ternary metal hydride [69]. Of 37 voids available within LaNi5 lattice structure, only 6 voids are occupied by hydrogen atoms to form LaNi5H6 ternary hydride [70]. It is notable that the hysteresis effect in AB5-type alloy is relatively small as compared to other low temperature systems with a large and distinct miscibility gap [59]. Mechanical alloying has also been employed to synthesize AB5 alloys [25,62,67,71]. For example, MmNi5 material mechanically alloyed at 200 rpm for milling time of 180 min exhibits faster hydrogenation kinetics as compared to arc-melted MmNi5 material [67]. Furthermore, Figure 7a,b present the absorption and desorption kinetic curves of MmNi5 alloy ball-milled with transition metals at concentration of 2.0 wt. %, respectively, showing reversible hydrogen absorption and desorption between 1.0 to 1.6 wt. % [67]. Singh and co-workers [62] observed that longer milling of MnNi4.6Fe0.4 alloy material improved hydriding kinetics, but this elevated kinetics behaviour comes at a cost of maximum uptake capacity as it deteriorated with increasing milling time. On the same note, an increase in rotation speed from 200 rpm to 300 rpm at constant milling time resulted in a reduction of hydrogen capacity. Similar trend was observed by Ares and co-workers [25] and they summarized that the trend is a result of extra hydrogen-trapping sites in amorphous domains, exhibited by short-milled alloys. For LaNi5 alloy, it is reported that ball-milling convert a part of LaNi5 to a non-absorbing state and as a result poor hydrogen uptake capacity is observed [71]. It was seen that there is quite a number of published work found in the literature for mechanical alloying of AB5-type alloys [62,67]. However, understanding the general extent to which mechanical alloying enhance the ability of these type of alloys to absorb hydrogen, as compared to their traditionally or arc-melted counterparts is still a gap. Firstly, to our knowledge no proper measurements of the hydrogen sticking or diffusion probabilities on either un-milled or milled AB5-type particles have been documented. The second issue is that mechanical alloying tempers with several physical properties that are vital for hydrogen sorption performance; these include expansion of the unit cell volume, strain and homogeneity of particles, particle size distribution, distance between adjacent voids and several more.

3.4. AB2-Type Alloys

3.4.1. Structural Characteristics of Mechanically Milled AB2-Type Alloys

In the AB2 Laves phase alloy, metal A is usually Ti or Zr, and metal B is another transition metal including V, Cr or Mn. These alloys crystallize in Lave phase structures [72,73]. Three types of laves phases; namely: hexagonal C14 phase (MgZn2), hexagonal C36 phase (MgNi2) and cubic C15 phase (MgCu2) represent AB2- type alloys [72]. In comparison, hydrogenation performance of hexagonal C36 phase has been reported to be poor [73], and it is not reviewed here. C15 and C14 phases (corresponding to atomic ratio of RA/RB = 1.225) have shown better performance as hydrogen absorbents [73]. Based on their structural stability, which is dependent on the atomic ratio, electronegativity and valence of the elements used [74,75], hexagonal C14 ZrCr2 Laves phases are more stable compared to cubic C15 [76].
Figure 8 represents example of powder diffraction patterns for C14 Laves phase type alloys. As seen with Mg-based and AB5-type alloys discussed earlier in this review, longer milling periods results in reduction of crystallite size and particle size. Furthermore, several other factors such as disappearance of peaks (between 2θ ≈ 48 and 100 in Figure 8), as well as peak broadening occur. To some extent, intense change of these factors may result in deterioration of hydrogen performance.

3.4.2. Hydrogen Absorption Behaviour of Mechanically Alloyed AB2-Type Alloys

The reported hydrogen storage of AB2-type alloys is nearly 2 wt. %, with low hysteresis [75]. But binary AB2-type alloys prepared by equilibrium techniques suffer from poor surface activity and poisoning intolerance [77]. As a result, efforts through synthesizing AB2 alloys utilizing non-equilibrium techniques such as mechanical alloying have been exploited [66]. For example, Ti0.5Zr0.5MnCr material prepared through mechanical alloying for a milling period of 5 min was reported to exhibit larger hydrogen absorption capacity compared to as-cast Ti0.5Zr0.5MnCr and materials milled much longer (Table 3 and Figure 9) [65].
However, those that are milled longer than 5 min exhibit faster hydriding kinetics and poor hydrogen absorption capacity compared to as-cast alloy [65]. Based on the literature data contained in this document, we understand that fractured fine, smaller particle sizes and amorphous nature, which are obtained through milling time above 25 min favour faster hydrogenation kinetics. Whereas combination of large to moderate particles and nano-crystalline structure, obtained during short milling time of less than 10 min favours large hydrogen absorption capacity.

3.5. Comparison between Hydriding Kinetics of Materials Prepared by MA with Other Techniques

Apart from mechanical alloying, melt spinning and spark plasma sintering techniques are also prevailing synthesis methods and thus we opted to make some comparisons between these three methods. Melt-spinning technique is a useful method that is used to change structures of as-cast alloys to amorphous structure. This method improves the amorphous nature of materials by reducing their grain size [44]. The only problem related to this method is the poor cycle stability of the melt-spun alloys, due to the disappearing of the metastable structures generated by melt-spinning [78]. Another attracting technique is spark plasma sintering (SPS), due to its ability to easily synthesize metal powders, with different melting temperatures within just 5–10 min [79]. This technique involves generation of spark plasma through direct pulse current and working concurrently with a uniaxial pressure. Sintering takes place at low temperatures in a small period of time and this makes the alloys to have low porosity [80]. SPS offers numerous favourable advantages over customary sintering procedures such as atmospheric furnace. For instance, the heating rate around 1273 K per min can be accomplished [79]. Further focal points are the lower sintering temperature, short holding time, no need of pre-compaction and shorter sintering time results in the likelihood to sinter nanometric powder to near theoretical density [81]. Although this technique produces alloys with high purity surfaces, it is known that the hydriding kinetics of composites, prepared by SPS, is greatly improved only when hydrogen absorption is conducted at high temperatures (573 K) [79].
Huang et al. [82] reported that alloys with amorphous surfaces absorb hydrogen faster compared to their crystalline counterparts. Schwarz and Johnson [83] recommended that two conditions should be fulfilled for an amorphous phase to form from a mixed elemental powder blend. The first condition is that the metals involved must have a vast negative heat of blending in the liquid state. According to a study conducted by Termsuksawad et al. [84], metals that have large negative heat of mixing exhibited a complete amorphous phase, while those with smaller negative heat displayed partial amorphous nature. Second, the condition is that diffusion rates of such metal systems must differ greatly. Metal alloy materials with nano-crystalline and/or amorphous phases are easily obtainable through several preparation techniques, including mechanical milling, spark plasma sintering and melt spinning [85,86]. There are numerous published papers supporting the above statement, and we combined some of those findings in Table 4. All the as-milled, as-spun and sintered alloy materials in Table 4 are reported to exhibit amorphous phases, and they are compared with their as-cast crystalline counterparts. According to the table, mechanical alloying of as-cast materials usually has a negative impact on the maximum hydrogen absorption capacity as all materials showed drastic decrease in their respective hydrogen content, except for Ti0.5Zr0.5MnCr alloy. Similar to other authors, Khajavi et al. [65] rationalized a loss in hydrogen capacity by deducing that milling process decreases the crystallite size, resulting in peak broadening. Nonetheless, milling significantly enhanced the absorption kinetics, and this is indicated by exceptional rate constants associated with all the as-milled counterparts in Table 4. Almost all the rate constants presented in the table were estimated from kinetic curves using chemisorption model [65] as shown by Equation (4),
α = kt
where α is the reacted fraction, t is hydrogenation time and k is rate constant. Metastable alloys (as amorphous, nano-crystalline, quasicrystalline, polycrystalline) can be prepared through melt spinning technique, since it allows quick quenching of melted metals. Most as-spun materials reported in this review have amorphous (examples being Mg11Y2Ni2 and Mg3LaNi0.1) and nano-crystalline phases (examples being Mg-10Ni-2Mn and Mg2Ni0.9Co0.1). On a contrary to mechanical alloying, melt spinning technique enhances both the hydrogen storage capacity and the kinetics (Table 4). The speedy hydrogenation kinetics for as-spun materials can be related to fine particle size and the crystal defects shaped as a result of rapid solidification, which facilitates swift nucleation and diffusion of hydrogen [23]. The enormous quantity of interfaces and grain boundaries created within the as-spun alloy offers easy access for hydrogen diffusion and facilitate easy absorption of hydrogen [23]. Another interesting technique featured in this review is spark plasma sintering. In the same way as MA and melt spinning, SPS-synthesized alloy materials exhibit an amorphous/nanocrystallite phase [87,88]. The absorption kinetics and capacity of sintered materials are affected by the amount (percentage), the type of material, and sintering conditions (temperature and pressure) used for SPS process. For example, hydrogen content of metallic Mg was reduced extremely from 7.26 wt. % to 2.52 wt. % upon blending with 77% V77.8Zr7.4Ti7.4Ni7.4 through SPS while hydrogen content of V35(Ti,Cr)65 sintered with ZrMn2 remained almost the same [87,88]. Interestingly, the rate constants of sintered alloy materials in Table 4 are better than those of as-cast materials. This is a clear indication that the SPS method has a positive effect on hydrogen absorption kinetics. Thus, it can be seen that all these three techniques are good candidates for improvement of hydrogen absorption kinetics, but alloys prepared using SPS technique in most cases have low hydrogen capacity. Hence, we take a further comparison between the prevailing techniques–MA and melt spinning.
Zhang et al [44] and Wu et al [92] made some efforts to prepare and compare structures and hydrogenation behaviour of as-milled and as-spun alloy materials. Their findings are summarized in Table 5. From the table, it can be seen that the hydrogen absorption capacity of Mg-10Ni-2Mm alloy is favoured by melt spinning technique as larger amount of hydrogen was stored by the material prepared through this technique than that prepared using mechanical alloying. However, YMg11Ni in Table 5 showed a different trend. The rate constants of materials prepared by mechanical alloying are better than those of materials synthesized by melt spinning technique. This indicates that mechanical alloying plays a significant influence on increasing the hydrogenation kinetics of the alloys as compared to melt spinning.
Hydrogenation of metal alloys is usually governed by three steps, namely (i) dissociation of hydrogen molecules into atomic hydrogen, (ii) the penetration of atomic hydrogen through grain boundaries, and (iii) the binding of atomic hydrogen on the metal, transforming into metal hydride [44]. The rate-determining step of the as-milled alloy is most likely ascribed to the ball milling rendering the second step, i.e., the penetration of atomic hydrogen through grain boundaries. The milling of as-cast alloys during mechanical alloying results in reduction of particle size and increment of high surface-to-volume ratio of alloy particles [93]. Due to these two changes, the distance for atomic hydrogen penetrating through the material is extremely reduced, which facilitates absorption greatly.

3.6. Surface Modification of Ball Milled Metal Alloys

Better and improved hydrogenation kinetics are achieved by nano-crystalline and/or amorphous metal alloys synthesized by mechanical alloying, but such materials still encounter several limitations, including impurity interactions, which include poisoning of the metal by impure gases such as carbon monoxide, sulphur and carbon dioxide; disproportionation during absorption/desorption cycling; retardation; innocuous and difficulty of initial activation [94]. These shortcomings result in rapid loss of hydrogen capacity with cycling, caused by impurities strongly or irreversibly absorbed on the surface active sites, and loss in absorption kinetics due to surface blanketing [95]. To overcome the above-mentioned problem, nano-structuring of ball milled materials and surface modification technology can be employed before hydrogen sorption can take place [96]. Surface modification involves the formation of a protective layer on the surface of metal alloy, thus preventing oxide-containing film to form. The most common surface modification technologies include fluorination, potassium borohydride, Ni-weak acid, ion implantation, hydrochloric (HCl) acid, platinum group metals (PGMs) decomposition and Y3O2 deposition. Yeung et al. [97] reported that the surface modification techniques PGMs deposition is the most attractive technology due to its ability to simultaneously prevent deactivation by guarding the catalytically active sites on the surface of bulk material against and enhance the hydriding kinetics during hydrogen absorption [97,98,99,100]. Moreover, this technique does not only enhance the hydriding rate, but also the absorption capacity of the alloys [99]. In particular, palladium is capable of catalyzing hydrogen sorption reactions, promoting easy activation, and improving poisoning tolerance. It has high hydrogen affinity and it is impermeable to larger molecules such as CO2 and sulphur [82]. Figure 10a–c show different mechanisms of hydrogen dissociation on the surface of Pd-treated AB5-type alloys [100]. During the hydrogen spillover mechanism hydrogen molecules are dissociated into hydrogen atoms on the surface of palladium, which then spillover to the active sites of the bulk alloy [98]. Pd thin film on the surface of AB5 alloy dissociates hydrogen molecules and allows the resulting hydrogen atoms to pass through the film as they diffuse into the bulk alloy. However, this technique is costly as huge amount of Pd is utilized [99].
Moreover, PGM-treated materials also exhibit the power to absorb hydrogen at low temperature, without the activation step prior to absorption, and enhanced kinetics of hydrogen sorption processes [101]. This method is extremely active towards dissociative hydrogen chemisorption and permeable for hydrogen atoms; enabling hydrogen to rapidly absorb into the bulk material while still maintaining the hydrogenation activity even after exposure to the impurities inhibiting hydrogen sorption [102]. The properties being improved include surface catalytic activity, specific surface area layer, activation characteristics, as well as protective nature of the complexes against poisoning materials, such as sulphur, water vapour and carbon oxides [82].
Davids et al. [103] exploited hydrogenation behaviour of ball-milled TiFe alloy surface modified using a metal-organic chemical vapour deposition technique (MOCVD), by the thermal decomposition of palladium (II) acetylacetonate (Pd[acac]2), mixed with the powder of the parent alloy. The authors reported that alloy materials containing 0.5% Pd have better hydrogenation behaviour than materials containing 1% Pd. The same trend in deterioration of hydrogenation performance, with an increase in Pd content, was observed by Pasquini and co-workers [104] who varied Pd content (corresponding to 4, 8 and 13 wt. %) on the surface of Mg metal. The observations are due to the theory that large amount of metal deposits on the surface of TiFe alloy and Mg metal results in slow penetration of hydrogen over a given period [104], as a result of fracture and defect blockage of the parent alloy material, which are usually formed during ball milling. Doyle et al. [105] recommended that the total weight of PGMs during surface modification of metal hydride-forming alloy should be in trace amounts (≤0.1wt. %) to avoid losses in hydrogenation behaviour. Another interesting study in literature was conducted by Modibane et al. [106] in which the authors carried out ball milling of AB5-type materials out using a FRITSCH Pulverisette planetary mill and subsequently deposited Pd nanoparticles using autocatalytic Pd deposition technique. Upon ball milling of the parent alloy (Figure 11a), porous agglomerates of micro-size particles (Figure 11b,c) were obtained due to blending and grinding of parent particles. Porous particles are fancied by Parker et al. [35] for enhancement of absorption kinetics. Palladium nanoparticles, which showed discontinuous nature and near-spherical shape, partially cover the AB5 materials (Figure 11d–f) and are responsible for dissociation of hydrogen molecules [106].
Table 6 reveals some of reported literature on surface modified metal alloys synthesized through mechanical alloying. Surface modification techniques herein are based on PGMs deposition and co-deposition. All surface modified alloys exhibit better hydrogenation kinetics as compared to their as-milled counterparts. Variations arise when comparing the hydrogen capacities. For instance, Pt deposition on as-milled La0.9Pr0.05Nd0.05Al0.3Mn0.4Co0.65Ni3.5 alloy results in increase of capacity, while Ru deposition on the same alloy lowers the hydrogen capacity. This is attributed to poor or little hydrogen affinity associated with Ru.

4. Conclusions

Hydrogenation kinetics and storage in solid state materials is a major concern in developing an up-and-coming hydrogen technology and economy. The present work reviews some challenges regarding the hydride-forming alloys, particularly their hydriding kinetics and capacity. In this regard, developments on hydrogenation performances of metal alloys, prepared by the mechanical alloying technique, were discussed. Moreover, this technique was compared with other preparation techniques, such as spark plasma sintering and melt spinning. The work is carried out in an attempt to facilitate suitable alloy preparation technique that can extensively improve the hydrogenation kinetics of different alloy materials. Although, much attention has been given to the techniques that are discussed in this work, with the aim of minimizing dehydrogenation temperature, enhancing the kinetics and poisoning-tolerance of different alloys, further research is needed. The collected literature data reveal encouraging improvements in terms of hydrogen storage kinetics through mechanical alloying. In the light of the achievements, through the combination of mechanical alloying and melt spinning techniques there is high potential in developing better metal hydride materials with high hydriding kinetics. Mechanically alloyed materials still suffer from cycling stability; they are prone to oxide layers formation, which hinders rapid absorption of hydrogen; and they also exhibit poisonous-intolerance. As a result, we further reviewed different surface modification techniques that can overcome these drawbacks. Surface modification of mechanically alloyed materials was overlooked at and it was seen that it promotes hydrogen storage kinetics by forming a protective layer on the surface of bulk alloy, thereby preventing the formation of oxide layer. Moreover, surface modification through deposition of PGMs on the surface of alloy materials stimulates poisonous tolerance of parent alloys.

Author Contributions

Conceptualization, K.D.M., and M.V.L.; funding acquisition, K.D.M. and M.J.H.; project administration, K.D.M.; Supervision, K.D.M., and M.V.L.; writing—original draft, T.R.S., T.C.M., M.W.D., K.D.M., M.J.H. and M.V.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation under Thuthuka program (UIDs. 117727 and 118113), University of Limpopo (Research Development Grants R202 and R232) and Sasol Inzalo Foundation, South Africa.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Andreas, R. Hydrogen storage methods. Naturwissenschaften 2004, 91, 157–172. [Google Scholar]
  2. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Mater. Sustain. Energy 2011, 414, 265–270. [Google Scholar]
  3. Berlouis, L.E.A.; Cabrera, E.; Hall, P.J. Thermal analysis investigation of hydriding properties of nanocrystalline Mg–Ni–and Mg–Fe-based alloys prepared by high-energy ball milling. J. Alloys Compd. 2000, 305, 82–89. [Google Scholar] [CrossRef]
  4. Sakintuna, B.; Lamari-darkrim, F.; Hirscher, M. Metal hydride materials for solid hydrogen storage: A review. Int. J. Hydrog. Energy 2007, 32, 1121–1140. [Google Scholar] [CrossRef]
  5. Crivello, J.C.; Dam, B.; Yartys, V.A. Review of magnesium hydride-based materials: Development and optimisation. Appl. Phys. A Mater. Sci. Process. 2016, 122, 1–20. [Google Scholar] [CrossRef] [Green Version]
  6. Liu, T.; Chen, C.; Qin, C.; Li, X. Improved hydrogen storage properties of Mg-based nanocomposite by addition of LaNi5 nanoparticles. Int. J. Hydrogen Energy 2014, 39, 18273–18279. [Google Scholar] [CrossRef]
  7. Yartys, V.A.; Lototskyy, M.V. An overview of hydrogen storage methods. In Hydrogen Materials Science and Chemistry of Carbon Nanomaterials; Springer: Dordrecht, The Netherland, 2004; pp. 75–104. [Google Scholar]
  8. Berube, V.; Chen, G.; Dresselhaus, M.S. Impact of nanostructuring on the enthalpy of formation of metal hydrides. Int. J. Hydrog. Energy 2008, 33, 4122–4131. [Google Scholar] [CrossRef]
  9. Pandey, S.K.; Srivastava, A.; Srivastava, O.N. Improvement in hydrogen storage capacity in LaNi5 through substitution of Ni by Fe. Int. J. Hydrog. Energy 2007, 32, 2461–2465. [Google Scholar] [CrossRef]
  10. Kazakov, A.N.; Dunikov, D.O.; Mitrokhin, S.V. AB5-type intermetallic compounds for biohydrogen purification and storage. Int. J. Hydrog. Energy 2016, 41, 21774–21779. [Google Scholar] [CrossRef]
  11. Takeda, H.; Kabutomori, T.; Ohnishi, K. Metal hydride air-conditioning. Encycl. Life Support Syst. 2009, 2, 249–263. [Google Scholar]
  12. Davids, W.; Lototskyy, M.V.; Linkov, V. Chemical surface modification for the improvement of the hydrogenation kinetics and poisoning resistance of TiFe. J. Alloys Compd. 2011, 509S, S770–S774. [Google Scholar]
  13. Semboshi, S.; Masahashi, N.; Hanada, S. Degradation of hydrogen absorbing capacity in cyclically hydrogenated TiMn2. Acta Metall. Sin. 2001, 49, 927–935. [Google Scholar] [CrossRef]
  14. Andreasen, A. Predicting formation enthalpies of metal hydrides. Nature 2001, 414, 353–358. [Google Scholar]
  15. Chebab, S.; Abdellaoui, M.; Latroche, M.; Paul-boncour, V. LaCaMgNi9 synthesized by mechanical alloying: Structural and electrochemical characterization. J. Tunis. Chem. Soc. 2016, 18, 52–59. [Google Scholar]
  16. Young, K. Increase in the Surface Catalytic Ability by Addition of Palladium in C14 Metal Hydride Alloy. Batteries. 2017, 3, 26. [Google Scholar] [CrossRef] [Green Version]
  17. Zhang, Y.; Li, B.; Ren, H.; Ding, X.; Liu, X.; Chen, L. An investigation on hydrogen storage kinetics of nanocrystalline and amorphous Mg2Ni1−xCox(x = 0 – 0.4) alloy prepared by melt spinning. J. Alloys Compd. 2011, 509, 2808–2814. [Google Scholar] [CrossRef]
  18. Simchi, H.; Kaflou, A.; Simchi, A. Synergetic effect of Ni and Nb2O5 on dehydrogenation properties of nanostructured MgH2 synthesized by high-energy mechanical alloying. Int. J. Hydrog. Energy 2009, 34, 7724–7730. [Google Scholar] [CrossRef]
  19. Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  20. Suñol, J.J.; Fort, J. Materials developed by mechanical alloying and melt spinning. Int. Rev. Phys. 2008, 2, 31–35. [Google Scholar]
  21. Scudino, S.; Sakaliyska, M.; Surreddi, K.B. Mechanical alloying and milling of Al-Mg alloys. J. Alloys Compd. 2009, 483, 2–7. [Google Scholar] [CrossRef]
  22. Poondi, D.; Singh, J. Synthesis of metastable silver-nickel alloys by a novel laser-liquid-solid interaction technique. J. Mater. Res. 2000, 35, 2467–2476. [Google Scholar]
  23. Xing, N.; Wu, Y.; Zhou, S. Improved hydrogenation-dehydrogenation characteristics of nanostructured melt-spun Mg-10Ni-2Mm alloy processed by rapid solidification. Prog. Nat. Sci. Mater. Int. 2010, 20, 49–53. [Google Scholar] [CrossRef] [Green Version]
  24. Ouyang, L.Z.; Qin, F.X.; Zhu, M. The hydrogen storage behavior of Mg3La and Mg3LaNi0.1. Scr. Mater. 2006, 55, 1075–1078. [Google Scholar] [CrossRef]
  25. Ares, J.R.; Cuevas, F. Influence of thermal annealing on the hydrogenation properties of mechanically milled AB5-type alloys. Mater. Sci. Eng. 2004, 108, 76–80. [Google Scholar] [CrossRef]
  26. Guevara, L.; Welsh, R.; Atwater, M.A. Parametric Effects of Mechanical Alloying on Carbon Nanofiber Catalyst Production in the Ni-Cu System. Metals (Basel) 2018, 8, 286. [Google Scholar] [CrossRef] [Green Version]
  27. Shang, C.X.; Bououdina, M.; Guo, Z.X. Mechanical alloying and electronic simulations of (MgH2 + M) systems (M = Al, Ti, Fe, Ni, Cu and Nb) for hydrogen storage. Int. J. Hydrog. Energy 2004, 29, 73–80. [Google Scholar] [CrossRef]
  28. Zaluski, L.; Zaluska, A.; Strrm-olsen, J.O.; Schulz, R. Catalytic effect of Pd on hydrogen absorption in mechanically alloyed Mg2Ni, LaNi5 and FeTi. J. Alloys Compd. 1995, 217, 295–300. [Google Scholar] [CrossRef]
  29. Iturbe-garcía, J.L.; García-núñez, M.R.; López-muñoz, B.E. Synthesis of the Mg2Ni Alloy Prepared by Mechanical Alloying Using a High Energy Ball Mill. J. Mex. Chem. Soc. 2010, 54, 46–50. [Google Scholar] [CrossRef]
  30. Kumar, S.; Tiwari, G.P.; Krishnamurthy, N. High performance FeTi-3.1 mass% V alloy for on board hydrogen storage solution. Energy 2014, 75, 520–524. [Google Scholar] [CrossRef]
  31. Andreasen, A. Hydrogenation properties of Mg–Al alloys. Int. J. Hydrog. Energy 2008, 33, 7489–7497. [Google Scholar] [CrossRef]
  32. Gleiter, H. Nanostrctured materials: Basic concepts and microstructure. Acta Mater. 2000, 48, 1–29. [Google Scholar] [CrossRef] [Green Version]
  33. Friedrichs, O.; Kolodziejczyk, L.; Fernandez, A. Synthesis of nanocrystalline MgH2 powder by gas-phase condensation and in situ hydridation: TEM, XPS and XRD study. J. Alloys Compd. 2007, 435, 721–724. [Google Scholar] [CrossRef]
  34. Lopez-Suarez, A. Effect of absorption and desorption of hydrogen in Ti and Ti alloys. In New Advances in Hydrogenation Processes: Fundamental and Applications; Iran Polymer and Petrochemical Institute: Teheran, Iran, 2017. [Google Scholar] [CrossRef] [Green Version]
  35. Parker, S.F.; Konrad, M.; Wieland, S.D. The effect of particle size, morphology and support commercial catalysts. Chem. Sci. 2019, 10, 480–489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ismail, M.; Mustafa, N.S.; Yap, A.H. Catalytic effect of CeCl3 on the hydrogen storage properties of MgH2. Mater. Chem. Phys. 2016, 170, 77–82. [Google Scholar] [CrossRef]
  37. Lototskyy, M.; Davids, M.W.; Pollet, B.G. Magnesium-based hydrogen storage nanomaterials prepared by high energy reactive ball milling in hydrogen at the presence of mixed titanium—iron oxide. J. Alloys Compd. 2015, 645, S454–S459. [Google Scholar] [CrossRef]
  38. Bouaricha, S.; Dodelet, J.P.; Schulz, R. Hydriding behavior of Mg–Al and leached Mg–Al compounds prepared by high-energy ball-milling. J. Alloys Compd. 2000, 297, 282–293. [Google Scholar] [CrossRef]
  39. Liang, G.; Huot, J.; Boily, S.; Van Neste, A.; Schulz, R. Hydrogen storage properties of the mechanically milled MgH 2–V nanocomposite. J. Alloys Compd. 1999, 291, 295–299. [Google Scholar] [CrossRef]
  40. Huang, H.X.; Huang, K.L.; Liu, S.Q. Microstructures and electrochemical properties of Mg0.9Ti0.1Ni1-xMx (M = Co, Mn; x = 0, 0.1, 0.2) hydrogen storage alloys. Powder Technol. 2010, 198, 144–148. [Google Scholar] [CrossRef]
  41. Vermeulen, P.; Niessen, R.A.H.; Notten, P.H.L. Hydrogen storage in metastable MgyTi(1-y) thin films. Electrochem. Commun. 2006, 8, 27–32. [Google Scholar] [CrossRef]
  42. Luo, W. (LiNH2-MgH2): A viable hydrogen storage system. J. Alloys Compd. 2004, 381, 284–287. [Google Scholar] [CrossRef]
  43. Castro, F.J.; Fuster, V.; Urretavizcaya, G. Hydrogen sorption properties of a MgH2-10wt. % graphite mixture. J. Alloys Compd. 2011, 509, S595–S598. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Zhang, W.; Qi, W.B.Y. A Comparison Study of Hydrogen Storage Thermodynamics and Kinetics of YMg11Ni Alloy Prepared by Melt Spinning and Ball Milling. Acta Metall. Sin. 2017, 30, 1040–1048. [Google Scholar] [CrossRef] [Green Version]
  45. Imamura, H.; Hashimoto, Y.; Sakata, Y. Preparation and Properties of Ball-Milled MgH2/Al Nanocomposites for Hydrogen Storage. Mater. Trans. 2014, 55, 572–576. [Google Scholar] [CrossRef] [Green Version]
  46. CRYSTMET Database, version 5.0.0; Toth Information Systems Inc.: Ottawa, ONT, Canada, 2013.
  47. Lototskyy, M.; Wafeeq, M.; Yartys, V.A. Nanostructured hydrogen storage materials prepared by high-energy reactive ball milling of magnesium and ferrovanadium. Int. J. Hydrog. Energy 2019, 44, 6687–6701. [Google Scholar] [CrossRef]
  48. Williams, M.; Lototskyy, M.V.; Pollet, B.G. Hydrogen absorption study of high-energy reactive ball milled Mg composites with palladium additives. J. Alloys Compd. 2013, 580, s144–s148. [Google Scholar] [CrossRef] [Green Version]
  49. Bhatnagar, A.; Pandey, S.K.; Srivastava, O.N. Catalytic effect of carbon nanostructures on the hydrogen storage properties of MgH2-NaAlH4 composite. Int. J. Hydrog. Energy 2014, 39, 14240–14246. [Google Scholar] [CrossRef]
  50. Yahya, M.S.; Ismail, M. Catalytic effect of SrTiO3 on the hydrogen storage behaviour of MgH2. J. Energy Chem. 2019, 28, 46–53. [Google Scholar] [CrossRef] [Green Version]
  51. Hanada, N.; Ichikawa, T.; Fujii, H. Remarkable improvement of hydrogen sorption kinetics in magnesium catalyzed with Nb2O5. J. Alloys Compd. 2006, 420, 46–49. [Google Scholar] [CrossRef] [Green Version]
  52. Juahir, N.; Mustafa, N.S.; Sinin, A.M. Improved hydrogen storage properties of MgH2 by addition of Co2NiO nanoparticles. RSC Adv. 2015, 5, 60983–60989. [Google Scholar] [CrossRef]
  53. Malka, I.E.; Pisarek, M.; Bystrzycki, J. A study of the ZrF4, NbF5, TaF5, and TiCl3 influences on the MgH2 sorption properties. Int. J. Hydrog. Energy 2011, 36, 12909–12917. [Google Scholar] [CrossRef]
  54. Wu, C.Z.; Wang, P.; Cheng, H.M. Hydrogen storage properties of MgH2/SWNT composite prepared by ball milling. J. Alloys Compd. 2006, 420, 278–282. [Google Scholar] [CrossRef]
  55. El-eskandarany, M.S.; Shaban, E.; Al-duweesh, A. Superior catalytic effect of nanocrystalline big-cube Zr2 Ni metastable phase for improving the hydrogen sorption / desorption kinetics and cyclability of MgH2 powders. Energy 2015, 91, 274–282. [Google Scholar] [CrossRef]
  56. Santos, S.F.; Ishikawa, T.T.; Huot, J. MgH2 + FeNb nanocomposites for hydrogen storage. Mater. Chem. Phys. 2014, 147, 557–562. [Google Scholar] [CrossRef]
  57. Liang, G.; Huot, J.; Boily, S.; Van Neste, A.; Schulz, R. Hydrogen storage in mechanically milled Mg–LaNi5 and MgH2–LaNi5 composites. J. Alloys Compd. 2000, 297, 261–265. [Google Scholar] [CrossRef]
  58. Chen, X.Q.; Podloucky, R.; Rogl, P. Computational and experimental study of phase stability, cohesive properties, magnetism and electronic structure of TiMn2. Acta Mater. 2003, 51, 1239–1247. [Google Scholar] [CrossRef]
  59. Erika, T.; Sebastian, C.; Fernando, Z. Temperature performance of AB5 hydrogen storage alloy for Ni-MH batteries. Int. J. Hydrog. Energy 2016, 41, 19684–19690. [Google Scholar] [CrossRef]
  60. Hayakawa, H.; Akiba, E.; Kohno, T. Crystal Structures of La–Mg–Nix (x = 3-4) System Hydrogen Storage Alloys. Mater. Trans. 2005, 46, 1393–1401. [Google Scholar] [CrossRef]
  61. Boeije, M.F.J.; Delczeg-czirjak, E.K.; Brück, E. On the phase stability of CaCu 5 -type compounds. J. Alloys Compd. 2017, 722, 549–554. [Google Scholar] [CrossRef]
  62. Singh, A.; Singh, B.K.; Srivastava, O.N. Studies on improvement of hydrogen storage capacity of AB5 type: MmNi4.6Fe0.4 alloy. Int. J. Hydrog. Energy 2004, 29, 1151–1156. [Google Scholar]
  63. Roberts, M.; Smith, R.I.; Hull, S. In situ investigation of commercial Ni(OH)2 and LaNi5-based electrodes by neutron powder diffraction. J. Mater. Res. 2015, 30, 407–416. [Google Scholar]
  64. Cermak, J.; David, B. Catalytic effect of Ni, Mg2Ni and Mg2NiH4 upon hydrogen desorption from MgH2. Int. J. Hydrog. Energy 2011, 6, 3–9. [Google Scholar] [CrossRef]
  65. Khajavi, S.; Rajabi, M.; Huot, J. Effect of cold rolling and ball milling on first hydrogenation of Ti0.5Zr0.5 (Mn1-xFex) Cr1, x = 0, 0.2, 0.4. J. Alloys Compd. 2019, 775, 912–920. [Google Scholar] [CrossRef]
  66. Gkanas, E.I.; Khzouz, M.; Makridis, S.S. Synthesis and Hydrogen Sorption Characteristics of Mechanically Alloyed Mg(NixMn1-x)2 Intermetallics. Mater. Sci. 2017, 12, 257–268. [Google Scholar]
  67. Srivastava, S.; Panwar, K. Effect of transition metals on ball-milled MmNi5 hydrogen storage alloy. Mater. Renew. Sustain. Energy 2015, 4, 19. [Google Scholar] [CrossRef] [Green Version]
  68. Broz, M.E.; Cook, R.F.; Whitney, D.L. Microhardness, toughness, and modulus of Mohs scale minerals. Am. Mineral. 2006, 91, 2006. [Google Scholar] [CrossRef]
  69. Chen, Y.; Sequeira, C.A.C.; Song, X.; Neto, R.; Wang, Q. Polytypism of La–Ni phases in multicomponent AB5 type hydride electrode alloys. Int. J. Hydrog. Energy 2002, 27, 63–68. [Google Scholar] [CrossRef]
  70. Dobrovolsky, V.D. The correlation between ionicity of metal-hydrogen bonds in hydrides and their thermal firmness. In Hydrogen Materials Science and Chemistry of Carbon Nanomaterials; Springer: Dordrecht, The Netherlands, 2007; pp. 421–428. [Google Scholar]
  71. Joseph, B.; Schiavo, B. Effects of ball-milling on the hydrogen sorption properties of LaNi5. J. Alloys Compd. 2009, 480, 912–916. [Google Scholar] [CrossRef]
  72. Koultoukis, E.D.; Christodoulou, C.N.; Fruchart, D. High-Temperature Activated AB2 Nanopowders for Metal Hydride Hydrogen Compression. Int. J. energy Res. 2014, 38, 477–486. [Google Scholar] [CrossRef] [Green Version]
  73. Ouyang, L.; Huang, J.; Zhu, M. Progress of hydrogen storage alloys for Ni-MH rechargeable power batteries in electric vehicles: A review. Mater. Chem. Phys. 2017, 200, 164–178. [Google Scholar] [CrossRef]
  74. Stein, F.; Palm, M.; Sauthoff, G. Structure and stability of Laves phases. Part I. Critical assessment of factors controlling Laves phase stability. Intermetallics 2004, 12, 713–720. [Google Scholar] [CrossRef]
  75. Hammerschmidt, T.; Ladines, A.N.; Drautz, R. Crystal-Structure Analysis with Moments of the Density-of-States: Application to Intermetallic Topologically Close-Packed Phases. Crystals 2016, 6, 18. [Google Scholar] [CrossRef] [Green Version]
  76. Ioannidou, A.; Makridis, S.; Kikkinides, E.S. Structural and Hydrogenation Properties of Zr0.9Ti0.1Cr1.2-xV0.8Nix (x=0, 0.4) Compounds. Mater. Sci. Forum. 2010, 636, 26–31. [Google Scholar]
  77. Liu, P.; Xie, X.; Liu, T. Hydrogen storage properties of (Ti0.85Zr0.15)1.05Mn1.2Cr0.6V0.1M0.1 (M=Ni, Fe, Cu) alloys easily activated at room temperature. Prog. Nat. Sci. Mater. Int. 2017, 27, 652–657. [Google Scholar] [CrossRef]
  78. Zhang, Y.; Xu, S.; Zhao, D. Hydrogen storage kinetics of nanocrystalline and amorphous Cu−Nd-added Mg2Ni-type alloys. Trans. Nonferrous Met. Soc. China 2014, 24, 3524–3533. [Google Scholar] [CrossRef]
  79. Liu, J.; Song, X.P.; Chen, G.L. Hydrogen storage performance of Mg-based composites prepared by spark plasma sintering. J. Alloys Compd. 2009, 486, 338–342. [Google Scholar] [CrossRef]
  80. Becker, H. Processing of bulk Al7075 alloy by spark plasma sintering. Mater. Sci. Eng. 2017, 179, 2050. [Google Scholar]
  81. Pei, P.; Song, X.; Liu, J.; Song, A.; Zhang, P.; Chen, G. Study on the hydrogen desorption mechanism of a Mg-V composite prepared by SPS. Int. J. Hydrog. Energy 2011, 37, 984–989. [Google Scholar] [CrossRef]
  82. Huang, H.; Huang, K. Effect of Fluorination Treatment on Electrochemical Properties of M1Ni3.5 Co0.6Mn0.4Al0.5 Hydrogen Storage Alloy. J. Braz. Chem. Soc. 2012, 23, 951–957. [Google Scholar]
  83. Schwars, R.B.; Johnson, W.L. Formation of an amorphous alloy by solid-state reaction of the pure polycrystalline metals. Phys. Rev. Lett. 1983, 51, 415. [Google Scholar] [CrossRef] [Green Version]
  84. Termsuksawad, P.; Niyomsoan, S.; Gavra, Z. Measurement of hydrogen in alloys by magnetic and electronic techniques. J. Alloys Compd. 2004, 373, 86–95. [Google Scholar] [CrossRef]
  85. Zadorozhnyy, V.Y.; Milovzorov, G.S.; Kaloshkin, S.D. Preparation and hydrogen storage properties of nanocrystalline TiFe synthesized by mechanical alloying. Prog. Nat. Sci. Mater. Int. 2017, 27, 149–155. [Google Scholar] [CrossRef]
  86. Møller, K.T.; Sheppard, D.; Ravnsbæk, D.B. Complex Metal Hydrides for Hydrogen, Thermal and Electrochemical Energy Storage. Energies 2017, 10, 1645. [Google Scholar] [CrossRef] [Green Version]
  87. Song, X.; Zhang, P.; Chen, G. The role of spark plasma sintering on the improvement of hydrogen storage properties of Mg-based composites. Int. J. Hydrog. Energy 2010, 35, 8080–8087. [Google Scholar] [CrossRef]
  88. Pei, P.; Song, X.P.; Chen, G.L. Improving hydrogen storage properties of Laves phase related BCC solid solution alloy by SPS preparation method. Int. J. Hydrog. Energy 2009, 34, 8597–8602. [Google Scholar] [CrossRef]
  89. Lv, P.; Guzik, M.N.; Huot, J. Effect of ball milling and cryomilling on the microstructure and first hydrogenation properties of TiFe + 4 wt.% Zr alloy. J. Mater. Res. Technol. 2019, 10, 1016. [Google Scholar] [CrossRef]
  90. Zhang, Q.A.; Zhang, L.X.; Wang, Q.Q. Crystallization behavior and hydrogen storage kinetics of amorphous Mg11Y2Ni2 alloy. J. Alloys Compd. 2013, 551, 376–381. [Google Scholar] [CrossRef]
  91. Zhang, D.Z.Y.; Li, B.; Ren, H.; Li, X.; Qi, T. Enhanced Hydrogen Storage Kinetics of Nanocrystalline and Amorphous Mg2Ni-type Alloy by Melt Spinning. Materials (Basel) 2011, 4, 274–287. [Google Scholar] [CrossRef] [Green Version]
  92. Wu, Y.; Lototsky, M.V.; Yartys, V.A. Microstructure and hydrogenation behavior of ball-milled and melt-spun Mg–10Ni–2Mm alloys. J. Alloys Compd. 2008, 466, 176–181. [Google Scholar] [CrossRef]
  93. Lass, E.A. Hydrogen storage measurements in novel Mg-based nanostructured alloys produced via rapid solidification and devitrification. Int. J. Hydrog. Energy 2011, 36, 10787–10796. [Google Scholar] [CrossRef]
  94. Modibane, K.D.; Williams, M.; Lototskyy, M. Poisoning-tolerant metal hydride materials and their application for hydrogen separation from CO2/CO containing gas mixtures. Int. J. Hydrog. Energy 2013, 38, 9800–9810. [Google Scholar] [CrossRef] [Green Version]
  95. Aguey-Zinsou, K.F.; Ares Fernandez, J.R.; Klassen, T. Effect of Nb2O5 on MgH2 properties during mechanical milling. Int. J. Hydrog. Energy 2007, 32, 221–233. [Google Scholar]
  96. Williams, M.; Nechaev, A.N.; Lototskyy, M.V. Influence of aminosilane surface functionalization of rare earth hydride-forming alloys on palladium treatment by electroless deposition and hydrogen sorption kinetics of composite materials. Mater. Chem. Phys. 2009, 115, 136–141. [Google Scholar] [CrossRef]
  97. Yeung, K.L.; Christiansen, S.C.; Varma, A. Palladium composite membranes by electroless plating technique Relationships between plating kinetics, film microstructure and membrane performance. J. Memb. Sci. 1999, 159, 107–122. [Google Scholar] [CrossRef]
  98. Parambhath, V.B.; Nagar, R.; Ramaprabhu, S. Effect of Nitrogen Doping on Hydrogen Storage Capacity of Palladium Decorated Graphene. Langmuir 2012, 28, 7826–7833. [Google Scholar] [CrossRef]
  99. Charbonnier, M.; Romand, M.; Goepfert, Y. Palladium reduction: A key step for the electroless Ni metallization of insulating substrates by a tin-free process. Thin Solid Film. 2006, 515, 1623–1633. [Google Scholar] [CrossRef]
  100. Shan, X.; Payer, J.H.; Jennings, W.D. Mechanism of increased performance and durability of Pd-treated metal hydriding alloys. Int. J. Hydrog. Energy 2009, 34, 363–369. [Google Scholar] [CrossRef]
  101. Denys, R.V.; Lototskyy, M.V.; Linkov, V.M.; Williams, M. Palladium mixed-metal surface-modified AB5-type intermetallides enhance hydrogen sorption kinetics. S. Afr. J. Sci. 2010, 106, 1–6. [Google Scholar]
  102. Lototskyy, M.V.; Williams, M.; Yartys, V.A. Surface-modified advanced hydrogen storage alloys for hydrogen separation and purification. J. Alloys Compd. 2011, 509, 555–561. [Google Scholar] [CrossRef]
  103. Davids, W.; Lototskyy, M.V.; Williams, M. Surface modification of TiFe hydrogen storage alloy by metal-organic chemical vapour deposition of palladium. Int. J. Hydrog. Energy 2011, 36, 9743–9750. [Google Scholar] [CrossRef]
  104. Pasquini, L.; Callini, E.; Maurizio, C. Magnesium nanoparticles with transition metal decoration for hydrogen storage. J. Nanoparticle Res. 2011, 13, 5727–5737. [Google Scholar] [CrossRef]
  105. Mark, I.; Doyle, L.; Benjamin, D. Hydrogen storage materials. United States Pat. 2000, 6, 165. [Google Scholar]
  106. Modibane, K.D.; Lototskyy, M.; Davids, M.W.; Williams, M.; Hato, M.J.; Molapo, K.M. Influence of co-milling with palladium black on hydrogen sorption performance and poisoning tolerance of surface modified AB5-type hydrogen storage alloy. J. Alloys Compd. 2018, 750, 523–529. [Google Scholar] [CrossRef]
  107. Ceramics, M.; Skorokhod, V.V.; Klimenko, V.P. Reversable hydriding of LaNi5−xAlx−Pd composite in the presence of carbon monoxide. Powder Metall. Met. Ceram. 2001, 39, 575–583. [Google Scholar]
  108. Willey, D.B.; Pederzolli, D.; Harris, I.R. Low temperature hydrogenation properties of platinum group metal treated, nickel metal hydride electrode alloy. J. Alloys Compd. 2002, 332, 806–809. [Google Scholar] [CrossRef]
Figure 1. Process flowsheet and the microstructures developed during double mechanical alloying of an Al–5wt. %Fe–4wt. %Mn powder mixture, reproduced from [19], with permission from Elsevier, 2001.
Figure 1. Process flowsheet and the microstructures developed during double mechanical alloying of an Al–5wt. %Fe–4wt. %Mn powder mixture, reproduced from [19], with permission from Elsevier, 2001.
Metals 10 00562 g001
Figure 2. XRD diffraction patterns of (a) nanocrystalline Mg and MgH2, and (b) Mg-Al alloy in the hydrogenated state (top) and dehydrogenated state (bottom). Figures reproduced from [33], and [31], respectively, [31] with permission from Elsevier, 2008; [33] with permission from Elsevier, 2007.
Figure 2. XRD diffraction patterns of (a) nanocrystalline Mg and MgH2, and (b) Mg-Al alloy in the hydrogenated state (top) and dehydrogenated state (bottom). Figures reproduced from [33], and [31], respectively, [31] with permission from Elsevier, 2008; [33] with permission from Elsevier, 2007.
Metals 10 00562 g002
Figure 3. The SEM images of the (a) MgH2, (b) as-milled MgH2, (c) as-milled MgH2 + 10wt. % CeCl3, (d) as-milled Mg–10 FeTiO3, (e) as-milled Mg:Al (90:10) and (f) as-milled MgH–5 at. % V. Figures (a–c) were reprinted with permission from [36], with permission from Elsevier, 2016 (d) reproduced from [37], with permission from Elsevier, 2015; (e) reproduced from [38], with permission from Elsevier, 2000 and (f) reproduced from [39], with permission from Elsevier, 1999.
Figure 3. The SEM images of the (a) MgH2, (b) as-milled MgH2, (c) as-milled MgH2 + 10wt. % CeCl3, (d) as-milled Mg–10 FeTiO3, (e) as-milled Mg:Al (90:10) and (f) as-milled MgH–5 at. % V. Figures (a–c) were reprinted with permission from [36], with permission from Elsevier, 2016 (d) reproduced from [37], with permission from Elsevier, 2015; (e) reproduced from [38], with permission from Elsevier, 2000 and (f) reproduced from [39], with permission from Elsevier, 1999.
Metals 10 00562 g003
Figure 4. (a) Hydrogenation kinetics of pure un-milled Mg compared with Mg-Al compounds with varying Al content ball milled for 20 h. Hydrogenation performed at T = 400 °C and a hydrogen pressure of 38 bar. Data extracted from ref. [31], with permission from Elsevier, 2008. (b) Hydrogenation kinetics of Mg and its carbon-based composites (15 bar and H2/250 °C), reproduced from [37], with permission from Elsevier, 2015.
Figure 4. (a) Hydrogenation kinetics of pure un-milled Mg compared with Mg-Al compounds with varying Al content ball milled for 20 h. Hydrogenation performed at T = 400 °C and a hydrogen pressure of 38 bar. Data extracted from ref. [31], with permission from Elsevier, 2008. (b) Hydrogenation kinetics of Mg and its carbon-based composites (15 bar and H2/250 °C), reproduced from [37], with permission from Elsevier, 2015.
Metals 10 00562 g004
Figure 5. X-ray difractograms of the unmilled; (a) and mechanically pulverized; (b) MmNi4:6Fe0:4 alloys and the inset picture focus on broadening of full width at half maximum (FWHM). reproduced from [61], with permission from Elsevier, 2004.
Figure 5. X-ray difractograms of the unmilled; (a) and mechanically pulverized; (b) MmNi4:6Fe0:4 alloys and the inset picture focus on broadening of full width at half maximum (FWHM). reproduced from [61], with permission from Elsevier, 2004.
Metals 10 00562 g005
Figure 6. SEM micrographs of (a) bulk MmNi5 alloy; (b) ball-milled MmNi5 alloy; (c) MmNi5 alloy ball-milled with Co; (d) Ni; (e) Mn and (f) Fe transition metals at concentration of 2.0 wt. %, obtained from [60].
Figure 6. SEM micrographs of (a) bulk MmNi5 alloy; (b) ball-milled MmNi5 alloy; (c) MmNi5 alloy ball-milled with Co; (d) Ni; (e) Mn and (f) Fe transition metals at concentration of 2.0 wt. %, obtained from [60].
Metals 10 00562 g006
Figure 7. (a) Absorption kinetic curves of MmNi5 alloy ball-milled with transition metals at concentration of 2.0 wt. %, and (b) Desorption kinetic curves of MmNi5 alloy ball-milled with transition metals at concentration of 2.0 wt. %. Reprinted from [67].
Figure 7. (a) Absorption kinetic curves of MmNi5 alloy ball-milled with transition metals at concentration of 2.0 wt. %, and (b) Desorption kinetic curves of MmNi5 alloy ball-milled with transition metals at concentration of 2.0 wt. %. Reprinted from [67].
Metals 10 00562 g007
Figure 8. Powder diffraction patterns of alloy Ti0.5Zr0.5MnCr before and after 5, 15, 30 and 60 min of ball milling, reproduced from [65], with permission from Elsevier, 2019.
Figure 8. Powder diffraction patterns of alloy Ti0.5Zr0.5MnCr before and after 5, 15, 30 and 60 min of ball milling, reproduced from [65], with permission from Elsevier, 2019.
Metals 10 00562 g008
Figure 9. Hydrogenation kinetics of Ti0.5Zr0.5MnCr alloys before and after ball milling for 5, 15, 30 and 60 min, reproduced from [58], with permission from Elsevier, 2019.
Figure 9. Hydrogenation kinetics of Ti0.5Zr0.5MnCr alloys before and after ball milling for 5, 15, 30 and 60 min, reproduced from [58], with permission from Elsevier, 2019.
Metals 10 00562 g009
Figure 10. Different mechanisms of hydrogen dissociation on the surface of Pd-treated alloy material; (a) hydrogen diffusion through Pd window, (b) hydrogen spillover with oxide layer being reduced by spilt hydrogen from palladium, and (c) hydrogen spillover with hydrogen atoms diffusing through the oxide layer, reproduced from [100], with permission from Elsevier, 2009.
Figure 10. Different mechanisms of hydrogen dissociation on the surface of Pd-treated alloy material; (a) hydrogen diffusion through Pd window, (b) hydrogen spillover with oxide layer being reduced by spilt hydrogen from palladium, and (c) hydrogen spillover with hydrogen atoms diffusing through the oxide layer, reproduced from [100], with permission from Elsevier, 2009.
Metals 10 00562 g010
Figure 11. Low-((a,b), Hitachi X-650 EM) and high-resolution ((cf), Zeiss Ultra) SEM images of raw AB5 alloy (a), the alloy ball milled without Pd black (b), and the alloy ball milled with Pd black after autocatalytic Pd deposition at different magnifications (cf), reproduced from [106], with permission from Elsevier, 2018.
Figure 11. Low-((a,b), Hitachi X-650 EM) and high-resolution ((cf), Zeiss Ultra) SEM images of raw AB5 alloy (a), the alloy ball milled without Pd black (b), and the alloy ball milled with Pd black after autocatalytic Pd deposition at different magnifications (cf), reproduced from [106], with permission from Elsevier, 2018.
Metals 10 00562 g011
Table 1. Lattice parameter, a, of different types of alloys before and after hydrogenation and their hydrogen storage capacities.
Table 1. Lattice parameter, a, of different types of alloys before and after hydrogenation and their hydrogen storage capacities.
Type of AlloyStructureAlloy Lattice Parameter, a, ÅHydride Lattice Parameter, a, ÅStorage Capacity,
wt. %
References
ABBCC2.9767.0291.75[12]
AB2C144.8664.9021.70[13]
C156.9397.1582.01[11]
AB5CaCu55.0035.3951.43[10]
A2B(P6222)5.2055.4633.75[11]
(Fddd)5.2845.4113.04[11]
Table 2. Thermodynamic and kinetic properties of ball-milled MgH2 doped with catalytic additives.
Table 2. Thermodynamic and kinetic properties of ball-milled MgH2 doped with catalytic additives.
Catalyst FamilyCatalyst AdditiveΔHdes, kJ/mol H2Hydriding Kinetics, min−1Hydrogen Capacity, wt. %References
No catalystUndoped MgH2761.4 × 10−37.66[52,57]
MetalMgH2+10 wt. % Ni756.89[49]
MgH2+10 wt. % Co715.62[49]
Metal halidesMgH2+10 wt. % CeCl375.70.056.32[36]
MgH2+NbF579.73.426.40[53]
MgH2+ZrF476.43.26.44[53]
HydridesMgH2+NaAlH40.283.10[49]
Metal oxideMgH2+SrTiO30.026.63[50]
MgH2+1 mol. % Nb2O5740.225.45[51]
Nano-sized alloysMgH2+10 wt. % Zr2Ni0.525.12[55]
MgH2+FeNb0.355.55[56]
Carbon materialsMgH2+5 wt. %SWNT1.346.72[54]
ΔHdes = Enthalpy of desorption*SWNT = Single-walled carbon nanotube.
Table 3. Hydrogenation properties of metal alloys synthesized by mechanical alloying.
Table 3. Hydrogenation properties of metal alloys synthesized by mechanical alloying.
Mg-based alloys
Alloy typeAlloy materialMilling conditions
(rotation speed and milling time)
Particle size distribution, µmHydriding kinetics,
min−1
Hydrogen capacity,
wt. %
References
As-cast alloyMg_48–701.4 × 10−37.66[57]
As-milled alloysMg
Mg-10FeTiO3
Mg-10FeTiO3-5MWCNT
500 rpm for 120 min
500 rpm for 120 min
500 rpm for 120 min
<0.05
<0.05
<0.05
0.04
0.07
0.31
4.98
2.71
5.22
[37]
Mg-10(FeV)
Mg-10(FeV)-5MWCNT
500 rpm, for 300 min
500 rpm, for 300 min
0.001–0.01
~0.01
0.02
0.02
6.96
6.69
[47]
Mg-0.5Pd
Mg-5Pd
500 rpm, for 360 min
500 rpm, for 360 min
0.05–0.1
<0.01
0.04
0.02
5.13
5.56
[48]
AB2-type alloys
Alloy typeAlloy materialMilling conditions
(rotation speed and milling time)
Particle size distribution, µmHydriding kinetics,
min−1
Hydrogen capacity,
wt. %
References
As-cast alloyTi0.5Zr0.5MnCr_15–290.041.79[65]
As-milled alloysTi0.5Zr0.5MnCrMilled for 5 min
Milled for 30 min
Milled for 60 min
~8
3–6
0.5–1.7
0.13
0.05
0.06
1.82
1.65
1.15
As-milled alloysMg(Ni0.75Mn0.25)2
Mg(Ni0.1Mn0.9)2
MgMn2
350 rpm for 600 min
350 rpm for 600 min
350 rpm for 600 min
<100.25
0.14
0.1
0.43
0.93
1.21
[66]
AB5-type alloys
Alloy typeAlloy materialMilling conditions
(rotation speed and milling time)
Particle size distribution, µmHydriding kinetics,
min−1
Hydrogen capacity,
wt. %
References
As-cast alloyMmNi5_5–500.241.20[67]
As-milled alloyMmNi5200 rpm for 180 min1–20.281.42
As-cast alloyMnNi4.6Fe0.4_~9.58_1.50[62]
As-milled alloysMnNi4.6Fe0.4200 rpm for 10 min
200 rpm for 30 min
300 rpm for 10 min
300 rpm for 30 min
_
~1.62
_
_
_
_
_
_
2.00
1.82
1.91
1.21
Table 4. Comparison of hydrogenation kinetics and capacity of as-cast alloys and those prepared by MA, SPS and melt spinning.
Table 4. Comparison of hydrogenation kinetics and capacity of as-cast alloys and those prepared by MA, SPS and melt spinning.
Mechanical Alloying/Milling
MaterialRate constant (min−1)Hydrogen absorption capacity (wt. %)ConditionsReferences
Metal alloy materialAs-castAs-milledAs-castAs-milled
Mg2Ni0.060.152.912.60473 K, 1.0 MPa[66]
TiFe+4 wt. % Zr (60 min milling)9.0 × 10−30.011.621.27273 K, 4.5 MPa[89]
TiFe+4 wt. % Zr (5 min milling)9.0 × 10−38.5 × 10−31.621.41273 K, 4.5 MPa[89]
Ti0.5Zr0.5MnCr (60 min milling)0.040.051.751.05273 K, 2.0 MPa[65]
Ti0.5Zr0.5MnCr (5 min milling)0.040.051.751.82273 K, 2.0 MPa[65]
Melt-spinning technique
MaterialRate constant (min−1)Hydrogen absorption capacity (wt. %)ConditionsReferences
Metal alloy materialAs-castMelt-spunAs-castMelt-spun
Mg11Y2Ni20.900.523.613.89523 K, 3.0 MPa[90]
Mg3LaNi0.10.250.602.732.90573 K, 4.0 MPa[24]
Mg-10Ni-2Mn0.010.024.675.09598 K, 1.0 MPa[23]
Mg2Ni0.9Co0.19.0 × 10−30.022.383.00473 K, 1.5 MPa[91]
Spark plasma sintering
MaterialRate constant (min−1)Hydrogen absorption capacity (wt. %)ConditionsReferences
Metal alloy materialAs-castSinteredAs-castSintered
Mg sintered with 77% V77.8Zr7.4Ti7.4Ni7.41.7 × 10−36.7 × 10−37.262.52573 K, 3.0 MPa[87]
V35(Ti,Cr)65 sintered with ZrMn20.010.042.862.89303 K, 0.6 MPa[88]
Mg sintered with 30% ZrMn21.7 × 10−33.3 × 10−37.266.34573 K, 3.0 MPa[87]
Table 5. Comparison between kinetics and hydrogen content of as-milled and as-spun techniques.
Table 5. Comparison between kinetics and hydrogen content of as-milled and as-spun techniques.
MaterialRate constant, min−1Absorption Capacity,
wt. %
ConditionsReferences
Alloy materialAs-milledAs-spunAs-milledAs-spun
Mg–10Ni–2Mm0.040.013.214.22573 K, 2.0 MPa[92]
YMg11Ni0.050.034.724.07593 K, 3.0 MPa[44]
Table 6. Hydrogenation properties of surface modified metal alloys prepared by mechanical alloying.
Table 6. Hydrogenation properties of surface modified metal alloys prepared by mechanical alloying.
Alloy TypeMetal AlloyRate Constant,
min−1
Hydrogen Capacity, wt. %Hydrogenation ConditionsReferences
As-milledLaNi50.045.30313 K, 1.5 MPa[28]
Pd depositionPd-LaNi52.385.18313 K, 1.5 MPa
As-milledMg2Ni0.063.55673 K, 2.0 MPa[28]
Pd depositionPd-Mg2Ni0.592.64573 K, 2.0 MPa
As-milledLaNi4.25Al0.754.8 × 10−13.21293 K, 1.5 MPa[107]
Pd depositionPd-LaNi4.25Al0.751.333.89293 K, 1.5 MPa
As-milledLa0.9Pr0.05Nd0.05Al0.3Mn0.4Co0.65Ni3.52.4 × 10−30.45293 K, 0.5 MPa[108]
Pt depositionPt-La0.9Pr0.05Nd0.05Al0.3Mn0.4
Co0.65Ni3.5
4.5 × 10−20.83298 K, 0.1 MPa
Ru depositionRu-La0.9Pr0.05Nd0.05Al0.3Mn0.4
Co0.65Ni3.5
3.3 × 10−20.34298 K, 0.1 MPa
Pt-Ru co-depositionPt-Ru-La0.9Pr0.05Nd0.05Al0.3
Mn0.4Co0.65Ni3.5
2.9 × 10−20.79298 K, 0.1 MPa
As-milledLa0.40Ce0.48(Nd,Pr)0.16Ni3.34Co0.64Al0.63Mn0.581.6 × 10−41.24293 K, 0.5 MPa[101]
Pd depositionPd-La0.40Ce0.48(Nd,Pr)0.16
Ni3.34Co0.64Al0.63Mn0.58
6.2 × 10−42.75293 K, 0.5 MPa
Pd-Ni co-depositionPd-Ni-La0.40Ce0.48(Nd,Pr)0.16
Ni3.34Co0.64Al0.63Mn0.58
1.1 × 10−33.60293 K, 0.5 MPa

Share and Cite

MDPI and ACS Style

Somo, T.R.; Maponya, T.C.; Davids, M.W.; Hato, M.J.; Lototskyy, M.V.; Modibane, K.D. A Comprehensive Review on Hydrogen Absorption Behaviour of Metal Alloys Prepared through Mechanical Alloying. Metals 2020, 10, 562. https://doi.org/10.3390/met10050562

AMA Style

Somo TR, Maponya TC, Davids MW, Hato MJ, Lototskyy MV, Modibane KD. A Comprehensive Review on Hydrogen Absorption Behaviour of Metal Alloys Prepared through Mechanical Alloying. Metals. 2020; 10(5):562. https://doi.org/10.3390/met10050562

Chicago/Turabian Style

Somo, Thabang Ronny, Thabiso Carol Maponya, Moegamat Wafeeq Davids, Mpitloane Joseph Hato, Mykhaylo Volodymyrovich Lototskyy, and Kwena Desmond Modibane. 2020. "A Comprehensive Review on Hydrogen Absorption Behaviour of Metal Alloys Prepared through Mechanical Alloying" Metals 10, no. 5: 562. https://doi.org/10.3390/met10050562

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