Next Article in Journal
Role of the Sulphur Source in the Solvothermal Synthesis of Ag-CdS Photocatalysts: Effects on the Structure and Photoactivity for Hydrogen Production
Previous Article in Journal
Controllable H2 Generation by Formic Acid Decomposition on a Novel Pd/Templated Carbon Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enhancing the Hydrogen Storage Properties of AxBy Intermetallic Compounds by Partial Substitution: A Short Review

1
Center for Energy Materials Research, Korea Institute of Science and Technology, Seoul 02792, Korea
2
Division of Nano & Information Technology, University of Science and Technology KIST School, Seoul 02792, Korea
3
Strategy & Technology Division, Hyundai Motor Company, Uiwang 16082, Korea
*
Author to whom correspondence should be addressed.
Submission received: 6 October 2020 / Revised: 2 December 2020 / Accepted: 9 December 2020 / Published: 15 December 2020

Abstract

:
Solid-state hydrogen storage covers a broad range of materials praised for their gravimetric, volumetric and kinetic properties, as well as for the safety they confer compared to gaseous or liquid hydrogen storage methods. Among them, AxBy intermetallics show outstanding performances, notably for stationary storage applications. Elemental substitution, whether on the A or B site of these alloys, allows the effective tailoring of key properties such as gravimetric density, equilibrium pressure, hysteresis and cyclic stability for instance. In this review, we present a brief overview of partial substitution in several AxBy alloys, from the long-established AB5 and AB2-types, to the recently attractive and extensively studied AB and AB3 alloys, including the largely documented solid-solution alloy systems. We not only present classical and pioneering investigations, but also report recent developments for each AxBy category. Special care is brought to the influence of composition engineering on desorption equilibrium pressure and hydrogen storage capacity. A simple overview of the AxBy operating conditions is provided, hence giving a sense of the range of possible applications, whether for low- or high-pressure systems.

1. Introduction

Humankind is on the verge of facing a worldwide energy crisis considering the soon-to-come fossil fuels shortage. The transition to environmentally friendly energy sources is a challenge that many countries are already tackling by reformatting their economy to implement alternative and sustainable solutions. As such, the hydrogen-based economy became one of the main candidates for the transition towards cleaner energy source, in the light of hydrogen’s positive impact on the environment and its intrinsic great potential as an abundant energy carrier: (i) high gravimetric energy density of 142 MJ kg−1 (against only 47 MJ kg−1 for petroleum) and (ii) high energy efficiency (fuel cells electrochemical processes show ~50–60% efficiency whereas that of combustion engines is as low as 25% for hydrogen-air mixtures, but still slightly higher than petrol-air) [1].
In the most general context, there are three different hydrogen storage methods: (i) compressed gas, (ii) liquid (cryogenic liquid H2 or liquid organic hydrogen carriers) or (iii) solid state storage as metal hydrides (see the flow chart in Figure 1, which elaborates the different techniques for hydrogen storage). To this day, gaseous storage of hydrogen is the most utilized method due to its relative simplicity. However, the low volumetric energy density of hydrogen at ambient temperature and atmospheric pressure (1 kg H2 occupies 11 m3) remains a major technical limitation to the widespread use of gaseous hydrogen [2,3]. Indeed, a high level of pressurizing is needed to meet the volume efficiency requirement of industrial-scale energy storage systems, causing additional energy consumption and costs. Liquid-state hydrogen storage greatly improves volumetric characteristics (from lower than 40 kg H2 m−3 of compressed hydrogen gas to 70.8 kg H2 m−3), but requires either cryogenic conditions (~21.2 K at ambient pressure) to avoid boil-off (hydrogen critical temperature is 33 K), or up to 104 atm of pressure for room temperature closed storage systems [2]. Either way, liquid-state hydrogen storage has to overcome technical and economic barriers for actual applications [4], since hydrogen liquefaction process (compressing and cooling) consumes about 30% of the energy stored [5], and 104 atm is challenging on an engineering point of view.
Hydrogen storage in solids has the advantage of bypassing the aforementioned limitations of gaseous and liquid-state storage, and provides safe and efficient storage conditions [6,7]. Solid-state hydrogen storage in metal hydrides appears to be the safest way of storing hydrogen since metal hydrides can be operated at relatively moderate temperatures and pressures compared to other storage states. Besides, some metal hydrides can be regarded as heat-storage systems, since hydrogen absorption/desorption is an exothermic/endothermic process [8], respectively, that can easily be triggered by operating at conditions different from the equilibrium (above or below the equilibrium pressure for a fixed temperature) [9]. Up to this date, various metal hydrides have been investigated due to their high hydrogen volumetric and gravimetric density. As such, complex hydrides like MBH4 (M = Li, Mg and Na, displaying 18.5 wt% H2 for LiBH4, 14.8 wt% H2 for Mg(BH4)2, and 10.6 wt% H2 for NaBH4) [10,11,12,13,14], and AlH3 (10.1 wt% H2) [15] were studied. Additionally, the properties of MgH2 (7.6 wt% H2) [16] enhanced with various catalysts (metal oxides, carbides, nitrides or metalloids like Si) [17,18,19,20,21], or even with transition metals to form ternary hydrides (Ni, Co and Fe respectively forming Mg2NiH4 (3.6 wt% H2), Mg2CoH5 (4.4 wt% H2), and Mg2FeH6 (5.5 wt% H2)) [22,23,24,25,26,27,28,29,30,31,32,33,34,35] were also explored. Although attractive, these high capacity hydrides all suffer major limitations absolutely preventing mass production and their use at ambient conditions; namely a high thermal stability (MBH4 and Mg-based ternary hydrides), a low reversibility (MBH4), and the need of extreme conditions for mechanochemical synthesis (AlH3) [10,15,26].
For this reason, room temperature hydrides such as intermetallics have drawn significant attention. Not only their thermodynamics is suitable for large-scale applications, but also they display high reversibility and a decent energy density per unit volume superior to those of gaseous and liquid phase (see Table 1) [36].
In the simplest case, intermetallic hydrides are AxByHz ternary compounds, because variations in elemental nature and their amount allow tailoring the sorption and storage properties of these hydrides. Element A is usually rare earth or transition metal and tends to form a stable hydride. Element B, on the other hand, is often a transition metal and does not form stable hydrides. It has been found that B:A ratios of 0.5, 1, 2, 5 form hydrides with a hydrogen-to-metal ratio of up to two [2]. The main hydride families are summarized in Table 2.
The main requirements for a large scale application of metal hydrides for on-board applications are (i) low hydrogen release temperature in the typical working conditions of a PEM fuel cell, (ii) high hydrogen absorption and desorption rates, (iii) acceptable costs and most importantly (iv) high storage capacity of 8 wt% according to the recent European VII FP call [38], which sets the bar even higher than the 6 wt% targeted by the American Department of Energy (DOE) [39]. It is difficult to achieve the gravimetric capacity target, especially for intermetallic hydrides. Hence, the main application for the intermetallic hydrides would be stationary applications, which are essential parts of renewable energy systems.
Many research groups have been trying to improve the characteristics of existing alloys in order to meet the EU’s and DOE’s requirements. Past, present and future developments in the field of hydrogen-based energy storage have been extensively documented recently (since 2016) in several exhaustive review articles, from general energy storage methods and delivery systems [40,41,42], to more specific storage technologies such as metal hydrides [8,43,44], including for instance Mg-based materials for energy storage [45,46,47]. However, to our knowledge, there are only very little recent reviews focusing on room temperature AxBy hydrides, aside from the comprehensive and comparative overview of AB3 alloys for stationary fuel-cell applications by Liu et al. [48], and that of vanadium-based hydrides for hydrogen storage by Kumar et al. [49]. Therefore, in this review article, we provide a brief overview of partial substitution in AxBy intermetallics and solid solutions for room temperature applications. We compare some of the most promising achievements and findings for each AxBy alloy category to identify and suggest the most promising representatives for further development. Material capabilities and performance are compared and discussed for both classical and recent works on the topic, notably in terms of desorption pressure-composition-isotherm, hysteresis, cycling performance and storage capacity.

2. AB5-Type Alloys

The AB5-type hydrides have been intensively studied during the last decades for their high potential for practical applications [51]. They have reversible and fast hydrogen absorption/desorption kinetics at near-ambient temperatures, simple activation process, and moderate pressure-temperature conditions of hydrogenation/dehydrogenation, which can easily be controlled. However, the maximum discharge capacity is limited to only around 1.5 wt% for the single CaCu5-type hexagonal structure [8,52,53].
In this section, LaNi5 is taken as the reference material of the AB5 family, in the light of its remarkable properties and features in comparison with other AB5 compounds that were recently studied. Indeed, LaNi5-based hydrides show good hydrogen absorption/desorption characteristics under near-atmospheric conditions and excellent kinetics [54,55]. The amount of hydrogen desorbed from a typical LaNi5-type metal-hydride system ranges from much less than 1 wt% up to 1.2 wt% H2 between room temperature and 373 K, with a theoretical maximum reversible storage capacity of 1.5 wt% H2 (still below the DOE’s target) [39,56]. Despite attractive properties, LaNi5-based compounds have a high cost in comparison with other alloys and show a significant capacity loss (higher than 30% after 800 cycles under impure hydrogen gas containing 100 ppm of O2 [57]), therefore urging to develop other materials with higher discharge capacity, better cyclic stability and lower cost [58].
The costly lanthanum in LaNi5 can thus be replaced by cheaper rare earth elements such as Ce [59], or by a cheaper rare earth mixture called mischmetal (Mm) consisting of La, Ce, Pr and Nd [60], which was investigated in many studies. MmNi5 possesses a hexagonal crystal structure similar to that of LaNi5 and tends to form stable hydrides. However, it shows a very high activation pressure (120 atm at 298 K), a high hydride formation pressure (30–60 atm at 298 K), large hysteresis between the absorption and desorption pressures and a maximum storage capacity of about 20% lower than that of LaNi5 [61,62]. Many groups have attempted to reduce the high hydride formation pressure in MmNi5 by partially substituting A and B components with various elements [61,63,64,65].
To enhance the hydrogen storage capacity, Ca may partially replace Mm in MmNi5 because of its lightweight (at. wt. 40) in comparison to Mm (at. wt. 140, corresponding to the following composition: La 22%, Ce 52%, Nd 15% and Pr 11%). Hence, for H/M = 1.0 the storage capacity of MmNi5H6 is 1.38 wt%, while that of Mm0.66Ca0.34Ni5H6 corresponds to 1.5 wt% [66,67]. The studies on Mm1-xCaxNi5 were first reported by Sandrock [68] and Shinar et al. [69]. Sandrock’s results show that the hydride dissociation pressure decreased with increasing Ca content, while Shinar’s results indicate that the substitution of Ca for Mm or La caused an increase in hydride dissociation pressure. Such contradictory behaviour results from the variation in Ca content, as elucidated by Wang et al. [70]. Indeed, they reported that the dissociation pressure of the hydrides (at 298 K) increased when x < 0.3 but decreased when 0.3 < x < 0.9, which was attributed to the effect of geometrical and electronic factors. In addition, the first hydrogenation incubation time shortened and its absorption rate increased along with increasing x in Mm1−xCaxNi5, and the hysteresis reduced.
Different from Mm and Ca (A substitutes), substitutions for B element were reported to be effective in tailoring the plateau pressure. Among them, Al was used for reducing the plateau pressure, for instance from 50 atm for MmNi5 down to 0.5 atm for MmNi4.2Al0.8. However, the maximum storage capacity decreased from 1.44 to 1.3 wt% and the plateau slope increased [71]. Meanwhile, Fe is known to increase hydrogen storage capacity (1.5 wt% for MmNi4.6Fe0.4), and reduce sloping and hysteresis [72,73].
Srivastava et al. [66,67] reported the effects of simultaneously substituting Ca, Al and Fe by preparing a series of Mm1−xCaxNi5−y−zAlyFez alloys. This composition turned out to compensate some drawbacks of previously described alloys, and result in a larger storage capacity. Mm0.9Ca0.1Ni4.7Fe0.2Al0.1 is thus showing a maximum storage capacity of 2.2 wt%, however Mm0.9Ca0.1Ni4.8Fe0.1Al0.1 (smaller Fe content) desorbed 1.9 wt% (see desorption PCIs plotted in Figure 2). The desorption behaviour of Mm1−xCaxNi5−y−zAlyFez alloys is plotted in Figure 2, together with those of the reference alloys (LaNi5 and MmNi5). Additional information on the alloys is available in Table 3, which summarizes key properties and main remarks for each alloy shown in the figure. From all the above alloy modifications, we can note the overall increase of desorption plateau pressures, and of the reversible storage capacity from 1 and 1.25 wt% (MmNi5 and LaNi5, respectively) to 1.65 wt% in Mm0.9Ca0.1Ni4.6Fe0.3Al0.1 at ambient temperature and pressure up to 60 atm.
Other than Al and Fe, various elements have also been used for B site substitutions. For instance, in 2000, Rożdżyńska-Kiełbik et al. [76] prepared a series of pseudo binary LaNi5 alloys by substituting 0, 5, 10, 15 and 20 at% of Zn for Ni. For an increasing Zn content, the produced LaNi5−xZnx alloys showed a linear increase of the unit cell volume, accompanied with a decrease of the absorption plateau pressures (in the range of 293 to 353 K) as well as a slight decrease in the hydrogen storage capacity as compared to the parent LaNi5 compound.
A lowered absorption plateau pressure and decreased hydrogen content is similarly observed when substituting by metalloids like Si (forming La28.9Ni67.55Si3.55 which yields a H/M ratio of 1.0 at 1.04 atm against 1.2 for LaNi5 at 1.3 atm) [77], while Sn mostly yields a better degradation resistance to thermal cycling [78].
Recently, more complex composition manipulations have been attempted by alloying Co, Al and Mn. The experimental study conducted by Briki et al. [79] reports that the synthesized LaNi3.6Mn0.3Al0.4Co0.7 (hexagonal CaCu5-type structure) reversibly absorbs/desorbs hydrogen in normal operating conditions (293 K and 6 bar), exhibits a significant reduction of hysteresis between hydriding and dehydriding, and a larger size of interstitial voids leading to a higher number of hydrogen atoms in the cell. Similar improvements without storage capacity decrease were also achieved thanks to Al in multicomponent alloys such as melt-spun LaNi4.7−xAl0.3Bix (x = 0.0, 0.1, 0.2, 0.3), whereas Bi substitution increased the absorption/desorption plateau pressure and reduced the hydrogen capacity [80]. In this investigation, Yilmaz et al. also evidenced the formation of BiLa and AlNi3 intermetallic phases at the grain boundaries, which results in an increased pulverization resistance of the alloy.

3. AB2-Type Alloys

AB2 Laves phase is another type of alloy with high potential for hydrogen storage. Usually, these alloys exist in three different crystal structures: cubic C15 (for instance MgCu2, ZrV2), hexagonal C14 (MgZn2, ZrMn2) and double hexagonal C36 (MgNi2). Laves phases with A = Zr show relatively high capacities (ZrV2H5.3, ZrMn2H3.6, ZrCr2H3.4), faster kinetics, longer lifetime and a relatively low cost in comparison to the LaNi5-based alloys. However, their hydrides are too stable at room temperature and more sensitive to contaminants [38,81]. This high stability of Zr-containing alloys is also seen in various type of materials, notably in amorphous structures in which hydrogen is irreversibly immobilized either in trapping sites [82,83,84], or by forming stable ZrH2 phase [85,86].
In this section, we take Ti–Mn Laves phase alloys as the reference material of the AB2 family, because of their easy activation, good hydriding-dehydriding kinetics, high hydrogen storage capacity and relatively low cost. Besides, they display high plateau pressure at room temperature (over 20 atm) and a sloping plateau often accompanied with a large hysteresis that requires major improvements [87,88].
In 2005, Toyota’s group demonstrated the use of Ti1.1MnCr alloys in a high-pressure metal hydride (MH) tank. This alloy has a maximum storage capacity of 1.9 wt%, but it has been reached only for a hydrogen pressure of around 350 atm at room temperature [89]. Kandavel et al. [90] substituted Zr in Ti1.1CrMn to provide favorable hydrogen sorption conditions and maximize the storage capacity. The increase in Zr content leads to a decrease in the equilibrium plateau pressure and faster absorption kinetics, together with an increase in the hydrogen storage capacity from 1.9 to 2.2 wt% for Ti1.1CrMn and (Ti0.9Zr0.1)1.1CrMn, respectively. Besides, Park et al. [87] conducted studies on Ti–Zr–Mn–Cr based metal hydrides and concluded that when Zr/Ti ratio increases, the lattice strain increases. This is partially responsible for a drastic increase of sloping, while the use of Cu was found very effective to mitigate the sloping.
In 1995, Morii et al. [91] prepared and investigated (Ti, Zr)(Ni, Mn, X)2 alloys, where X is V or/and Fe. The results showed that V lowers both hysteresis and plateau pressure. On the other hand, Ni raises the plateau pressure and reduces the width of the plateau region, while Fe flattens and lengthens it.
Improvements of the hydrogen storage properties of Laves phase AB2-type alloys at 303–308 K and 1–15 atm have been achieved by introducing non-stoichiometry at the A site of (Ti0.65Zr0.35)1+xMnCr0.8Fe0.2 alloys. From pressure-composition-temperature (PCT) measurements, the maximum hydrogen storage capacity was found to be around 2.2 wt% at 35 atm and 305 K for (Ti0.65Zr0.35)1.1MnCr0.8Fe0.2, which is approximately 16% higher than that of the commercially available “Hydralloy C5” (Ti0.955Zr0.045Mn1.52V0.43Fe0.12Al0.03). These alloys show remarkable hydrogenation kinetics: the full capacity is reached within 10 min without any need for activation [92].
Alloys without zirconium (such as Ti1.02Cr1.0Fe0.75Mn0.25) display 1.55 wt% of reversible hydrogen storage capacity when the temperature is as low as 233 K. However, without zirconium the effective hydrogen capacity is optimal only when the pressure is higher than 70 atm [93], proving the effectiveness of Zr in Laves phase alloys. Figure 3 shows the desorption behavior of some noteworthy AB2 alloys.
Recent developments (<5 years) on AB2-type materials have highlighted their significant potential for high-pressure compressors, notably (Ti,Zr)(Mn,Cr)-based alloys. Indeed, Corgnale et al. [96] proposed a techno-economic analysis of metal hydride systems for efficient and novel high-pressure compressors. Among various materials, TiCr1.9, Ti1.1CrMn, TiCrMn0.4Fe0.4V0.2, and (Ti0.97Zr0.03)1.1Cr1.6Mn0.4, they suggested the last one as the best candidate for their novel two-stage hybrid electrochemical and metal hydride compression system, since pressures about 863 atm can be reached with a thermal power provided at approximately 423 K.
Pickering et al. [97] further demonstrated the high capability of (Ti,Zr)(Mn,Cr)-based alloys for both hydrogen storage and high-pressure compression by producing industrial volumes (~10 kg) of tailored AB2 intermetallics (A = Ti + Zr, B = Cr + Mn + Ni+Fe + V) by means of vacuum induction melting process. They successfully tuned the hydrogenation properties of the alloy, showing that at a fixed quite low Zr/(Ti + Zr) ratio the PCT properties of the materials can be adjusted in a wide range by the variation of V content which, in addition, results in the increase of the hydrogen storage capacity. Cheaper alternatives to pristine Ti and V nevertheless exist, notably by replacing those high purity raw materials by their low-cost and low-purity counterparts, namely Ti sponge and ferrovanadium (FeV), respectively. Such substitution in (Ti,Zr)(V,Fe,Cr,Mn) reduces the raw material cost by 83%, without altering the dissociation pressure (15 atm), nor the reversibility (1.4 and 1.5 wt% H2 after 1000 cycles, against an initial capacity of 2 and 1.7 wt% H2 for pristine and modified alloys, respectively) [98].
The development of hybrid hydrogen storage system is equally appealing to the scientific community. For instance, rare earth elements (RE) such as La, Ce or Ho in Ti1.02Cr1.1Mn0.3Fe0.6RE0.03 have been shown in 2018 to yield better activation behaviour, larger storage capacity but lower desorption plateau pressure [99]. This study suggests Ti1.02Cr1.1Mn0.3Fe0.6La0.03 alloy as the best overall candidate since it can be fully activated at room temperature, and has a hydrogen storage capacity as high as ~1.7 wt%. Another example of hybrid system is reported by Puszkiel et al. [100], who demonstrated that mixing expanded natural graphite (ENS) into (Ti0.9Zr0.1)1.25Cr0.85Mn1.1Mo0.05 alloy not only improves the heat transfer properties, but also yields a reversible capacity of about 1.5 wt%, together with decent cycling stability and rapid reaction kinetics (25 to 70 s).
Although all the above-mentioned (Ti,Zr)(Mn,Cr)-based Laves phase alloys are widely investigated in the light of their superior potential for high-pressure compressors (and hybrid hydrogen storage), Zr-based AB2 materials are nevertheless not to be discarded although they display significantly lower desorption plateau pressures. Wu et al. [101] thus elucidated the role of Ni addition on the hydrogen storage characteristics of Zr(V1−xNix)2 (x = 0.02, 0.05, 0.1, 0.15, 0.25) intermetallic compounds. The hydrogen absorption capacity turns out to decrease, and the equilibrium pressure increases with increasing Ni content. The alloys exhibit fast absorption kinetics at room temperature and a remarkable cyclic stability even after 100 hydrogen absorption/desorption cycles.
Owing to fast kinetics, high equilibrium pressure and impressive volumetric hydrogen storage density at ambient temperature, ZrFe2 based alloys are similarly good candidates for high pressure compressed hydrogen tanks. To bypass its rather large hysteresis, Mn, Ti, V and Cr addition [102,103] has been considered. On one hand, V addition is suggested to improve the hysteresis, while Ti helps to lower plateau sloping as well as to increase the plateau pressure. Zr1.05Fe1.6Mn0.4 shows a relatively high dehydriding pressure of 20.6 atm at 298 K, while (Zr0.5Ti0.5)1.05Fe0.95MnV0.05 delivers a maximum capacity of 1.64 wt% H2 and shows a dehydriding pressure of 6.8 atm at 298 K (calculated from Van’t Hoff plots) [102]. Additionally, the simultaneous Cr/V substitution for Fe decreases the equilibrium pressure (due to the enlarged unit cell), and Zr1.05Fe1.85Cr0.075V0.075 seems to exhibit decent overall hydrogen storage properties (1.54 wt%, and a desorption equilibrium pressure of 9.7 atm at 243 K) [103].
Figure 4 summarizes desorption PCT curves of some representative materials for high pressure compressor described above. Unlike Figure 3, most of alloys shown here display significantly higher desorption plateau pressures that seem to be achieved at the expense of the storage capacity. It is interesting to note the excellent capacity of AB2 alloys to cover this broad range of properties with relatively simple manipulations of the composition. Indeed, essential properties such as absorption/desorption plateau pressures, maximum/reversible storage capacity, activation and cyclic performance (among others) can be tuned to adapt the alloy to the requirements of the target applications. This outstanding ability is even more obvious when carefully comparing the effect of substitutional modifications on each alloy presented in Figure 3 and Figure 4, as shown in the comparative Table 4.

4. AB-Type Alloys

AB-type alloys are attractive materials for hydrogen storage because of their light molar mass and high weight capacities. TiFe alloys with cubic CsCl-type structures are the most known alloys of this class and stand among the best hydrogen storage materials up to this date [8,104].
TiFe intermetallic compound is one of the most promising hydrogen storage alloys, due to its relatively high theoretical hydrogen storage capacity (1.9 wt%) at near-ambient conditions compared to other AxBy families. Besides, its economical merit based on the abundance and low cost of the constituting elements encourages extensive investigations on the TiFe system.
The hydrogen sorption and desorption in TiFe was first described by Reilly and Wiswall in the year 1974 [105]. They reported two stable intermetallics of TiFe system (TiFe and TiFe2) and a third, Ti2Fe that forms only above 1273 K (dissociates to TiFe and Ti below that temperature). Only TiFe is known to make two ternary hydrides, TiFeH and TiFeH2.
The hydrogen absorption in TiFe alloy depends on two factors: (i) the Fe/Ti ratio and (ii) the oxygen amount in the alloy. TiFe intermetallic exists over a narrow composition range of ~2.5 at% (from 49.5 to 52 at% Ti). Slightly less than 49.5 at% Ti results in a two-phase mixture of TiFe2 and TiFe, the first being of no use since it is a non-hydride former. If Ti content is higher than 52 at%, the alloy consists of TiFe and (α or β) Ti solid solution [105]. Although Ti itself readily forms hydrides, they are highly stable and are non-reversible at the temperatures of interest (ambient).
The lower plateau level and general shape of the curve is not significantly affected but the maximum hydrogen storage capacity substantially reduces with the increase in oxygen content (Figure 5) [106]. Additionally, TiFe usually requires heating over 573 K for activation, which again suggests the low poisoning tolerance resulting in significant deterioration of hydrogen sorption even for trace amounts of gas species (oxygen and water vapor for instance) [105,107,108]. Most importantly, surface oxidation issues induce significant difficulties notably in the first hydrogenation. The problem with first activation can be resolved by partial replacement of the base element [109,110,111,112,113,114], mechanical alloying [115,116], surface modifications [108], groove rolling and high-pressure torsion [107,117]. Most of these studies did not lead to an improvement in hydrogen storage properties, and the result was usually a decreased maximum hydrogen absorption capacity and increased desorption temperature of the intermetallic hydrides.
There is 10% volume increase when initial hydrogenation occurs. This exerts stresses on unhydrided core, thus results in cracks. The presence of second phase particles (TiFe2, Ti10Fe7O3, Ti) promotes activation: (i) lowers fracture toughness of TiFe and (ii) provides interface for preferential hydride nucleation and penetration [106].
The intimacy of the alloy to a minor level of oxygen creates another feature to be noted; TiFe microstructure with minor oxygen contamination exhibits at least two phases (TiFe and an oxygen stabilized Ti10Fe7O3 as fine eutectic distribution). For this oxygen stabilized phase, each oxygen binds with 5.7 metal atoms, so even 1 wt% O-contamination results in 19 wt% Ti10Fe7O3 phase, which does not form any hydride [106].
The level of plateau pressure determines the stability of the hydride. Partial substitution of Fe by 3d-transition metals can disrupt and thus modify the stability of the resulting hydride (TiFe1−xMx). This allows the alloy to be tailor-made with appropriate properties for particular application. Mn can be used in that purpose, for instance by providing a heat-treatment free novel activation route [118]. Shang et al. [119] synthesized Ti1.1Fe0.8Mn0.2 (Figure 6), and demonstrated that partial replacement of Fe with Mn as well as excess Ti helped to reduce the activation process temperature from 573 K to 423 K and to increase the amount of stored hydrogen from 1.35 to 1.5 wt% (mostly due to Mn) under ambient temperature and a pressure of 30 atm (first plateau).
Comparable improvements of the activation procedure are also reported for off-stoichiometric TiFe0.9 alloys, for which Mn substitution for Fe is additionally shown to reduce the equilibrium pressure at room temperature [113]. Plateau pressures can also be tuned by Cu substitution for which the cell parameter linearly increases with Cu content [120]. Hence, the combination of Cu, Mn and off-stoichiometry is of great interest for tailoring the properties of TiFe-based alloys, as recently demonstrated by Dematteis et al. [121]. They investigated the effect of Mn and Cu substitution for Fe in TiFe0.9 system, and the thorough structural and thermodynamic study shows that all synthesized alloys display fast kinetics and high storage capacity. The report suggests that (i) both Mn and Cu substitutions increase the cell parameter of TiFe (decreased first plateau pressure), whereas (ii) Cu substitution increases the second plateau pressure, and (iii) the hydride stability is not solely driven by cell volume, but may also strongly depend on the electronic properties of the substituting elements.
Jang et al. in 1986 [111] studied Zr substituting Ti, rather than Fe, in TiFe alloy for improved activation properties. Particularly, Ti0.9Zr0.1Fe activated at room temperature and required no heat treatment. There was a visible enhancement in the β phase hydride (TiFeH) formation and suppression of γ phase (TiFeH2). Nagai et al. [122] in the year 1988 further studied the result of Zr addition in TiFe. Partial substitution of Zr (~1–15 at%) results in TiFe and two other phases ((Ti1−yZry)2Fe, a hydride former, and Ti(Fe1−xZrx)2, a non-hydride former). They reported activation of the alloy at ambient temperatures with reduced incubation time for hydrogenation kinetics, without any loss in hydrogen storage capacity.
Lee and Perng in 1999 [123] studied partial substitution of Co, Ni and Al in TiFe. They observed that Co and Ni (similar size with Fe) addition led to the formation of a small fraction of α phase (solid solution) with hydriding characteristics similar to that of pure TiFe, but the addition of Al (large atomic size as compared to Fe) resulted in a much larger α phase fraction (see Figure 6). All three alloys did not require any activation treatment.
Kuziora et al. [124] very recently explored the effect of refractory metals (Ta and Mo) on the hydrogen storage properties of TiFe alloys prepared via suspended droplet alloying (SDA). The resultant alloys, Ti0.5Fe0.45Ta0.05 and Ti0.5Fe0.4Mo0.1, absorbed ~1 and 1.4 wt% H2, respectively. The alloys displayed sloping plateau, and despite the encouraging results, the correlation between alloy composition and absorption plateau pressure could not be established.
Different from refractory metals, recent research is focused on the use of Zr and other alloying elements for possible room temperature activation. Jain et al., in 2015, presented a comparative study on the effect of Zr, Ni and Zr7Ni10 alloy on the TiFe hydrogenation properties [125]. They concluded that Zr addition annihilates the initial activation requirements and reduces the incubation time without affecting the reversible storage capacity.
Very recently, in the year 2020, Yang et al. [126], documented the effect of Cr, Mn and Y substitution for Fe on the hydrogen storage properties. They concluded that Cr substituted alloys (TiFe0.9Cr0.1, TiFe0.9Cr0.1Y0.05) have lower equilibrium pressure and sloped plateaus, thus providing better hydrogenation kinetics as compared to Mn substituted alloys (TiFe0.9Mn0.1, TiFe0.9Mn0.1Y0.05), which have higher equilibrium pressure but flat plateaus and thus better dehydrogenation kinetics. Y substitution in Ti–Fe–Mn and Ti–Fe–Cr based alloys resulted in αY phase, which transforms to YH3 during hydrogenation.
Ha et al. [127] investigated the contrast in the microstructure of as cast and heat treated TiFe-6 wt% ZrCr2 alloys. They reported that the as cast alloy has 65 wt% TiFe and 35 wt% TiFe2 (C14 Laves phase) while the heat-treated alloy has a portion of TiFe2 transformed to TiFe phase (84 wt%). The activation profile reveals that both the alloys can be activated at room temperature under 30.6 atm H2 but the as cast alloy displays enhanced absorption kinetics (activation starts without any delay while its heat-treated counterpart requires 40 h of incubation time). Both specimens show approximately equal maximum hydrogen storage capacity of 1.7 wt%. The first plateau for the annealed alloy is flatter in shape and the desorption isotherm shows less retained hydrogen as compared to the as cast alloy. In parallel, Jung et al. [128] conducted a study on tailoring the equilibrium plateau pressure of TiFe monohydride and dihydride via V substitution for both Ti and Fe, in order to achieve maximum reversible capacity under a narrow pressure range. When V substitutes for Ti, the monohydride plateau pressure rises whereas a pronounced opposite trend is seen if V substitutes for Fe. Interestingly, the plateau pressure for dihydride is lowered in both the cases.
To summarize all the above, similarly to AB2 alloys, there exists an appreciable opportunity to fabricate TiFe based AB alloys with ternary element substitution resulting in tailored hydrogen storage and hydrogenation kinetics properties as per the application requirement, as summarized in Table 5.

5. AB3-Type Alloys

Research on AB3 alloys, whose structure consist of combined AB2 and AB5 (see equation below), was initially motivated by their strong potential for Ni-MH batteries [48,129]. Indeed, negative electrode materials based on AB3 can offer a higher hydrogen storage capacity than AB5-types alloys (already commercialized), but unfortunately suffer a severe degradation of their cyclic properties due to pulverization and oxidation/corrosion [130,131].
AB5 + 2(AB2) = 3(AB3)
Most frequently based on La2MgNi9, AB3 alloys however turn out to be promising for stationary hydrogen storage applications as well, considering their good activation and hydrogenation/dehydrogenation kinetics on one hand, and their relatively high storage capacity and low cost on the other (thus combining the best features of AB5 and AB2 respectively) [48]. Additionally, the phase composition of AB3 alloys (hence their properties) can be tuned by means of element substitution, heat treatment and different material processing methods, similarly to AB2 alloys [48].
Pioneering work in the seventies [132,133] first reported hydrogen solubility and hydride forming ability of AB3 alloys based on rare earth elements (A side) and transition metals (B side). Later on, Kadir et al. further investigated such alloys, by providing exhaustive reports on the effect of rare earth elements on the hydrogenation properties of AB3 alloys (La, Ce, Pr, Nd, Sm, Gd) [134], as well as on the effect of La and Mg partial replacement by Ca and/or Y in La-Mg-Ni based alloys [135,136].
The hydriding characteristics of LaNi3/CaNi3 and RT3 phases (R = Dy, Ho, Er, Tb, Gd; T = Fe or Co) showed that the hydrogen storage capacity of the AB3 phases exceeds that of the well-known hydrogen absorber LaNi5 [137]. Due to the special crystal structure of AB3 compounds, it is possible to combine Mg, Ca, and rare earth elements in the A side. Kadir et al. [136] synthesized (La0.65Ca0.35)(Mg1.32Ca0.68)Ni9 which absorbs ~1.87 wt% H2 at ~33 atm H2 and 283 K. Under identical pressure condition, Chen et al. [137] reached up to 1.8 wt% H2 at 293 K for LaCaMgNi9.
In order to improve the performance of La–Mg–Ca–Ni AB3-type alloy, Lim et al. investigated the effects of partial substitution with Ce and Al on the hydrogenation properties of La0.65−xCexCa1.03Mg1.32Ni9−yAly alloys [138]. Their results indicated that the hydrogen storage capacity significantly decreased after Ce and Al substitution. Xin et al. [139] investigated the effects of Y partial substitution on overall hydrogen storage properties of (La0.65Ca0.35)(Mg1.32Ca0.68)Ni9. At 1 atm H2, the hydrogen desorption capacity of La0.60Y0.05Mg1.32Ca1.03Ni9 was approximately 1.624, 1.616, and 1.610 wt% at 298, 313, and 333 K, respectively. In addition, the equilibrium pressure could be tailored by altering the Y amount to range 1–10 atm.
The effect of half replacement of Ca by R (R = Nd, Gd and Er) on the phase structure and hydrogen storage property of Ca2MgNi9 compound was investigated in 2019 by Zang et al. [140]. Results showed that alloys with Gd, Er or Nd instead of La have lower maximum storage capacity (1.4, 1.2, and 1.5 wt% H2, respectively, against 1.87 wt% H2 for La). Desorption behaviours of some remarkable AB3 alloys (plotted in Figure 7) show flatter plateau pressures than some AB2 and AB alloys (Figure 3, Figure 4 and Figure 6) while displaying comparable storage capacity (see detailed summary in Table 6).
To summarize, partial substitution in the B site of Ni for elements with larger atomic radius increases the unit cell volume and results in a decrease of the absorption and desorption plateau pressures. As such, increasing Co concentration (in La0.7Mg0.3Ni3.4−xMn0.1Cox [141] or in La2Mg(Ni1−xCox)9 (x = 0.1–0.5) [142]), or Al and Mo (in La0.7Mg0.3Ni3.5−x(Al0.5Mo0.5)x with x = 0–0.8 [143]) decreased the desorption equilibrium pressure.
“Pseudo AB3” alloys like A2B7 (Ce2Ni7-type structure) or even A5B19 (Ce5Co19) with stacked super structures were also considered both for battery and hydrogen storage applications. As such, in the aim of developing a new type of Mg-free AB3-type alloy system, Yan et al. [144] investigated the effect of La and Mg replacement by Y on one side, and that of Ni by Mn and Al on the other. Similar to La-Mg-Ni based system, the studied LaY2Ni8.2Mn0.5Al0.3 (AB3-type), LaY2Ni9.7Mn0.5Al0.3 (A2B7-type) and LaY2Ni10.6Mn0.5Al0.3 (A5B19-type) alloys are multiphase structures, with hydrogen storage capacities at 313 K of 0.85, 1.48 and 1.45 wt% for AB3, A2B7 and A5B19-type alloys, respectively (A2B7 and A5B19 being larger than that of the AB5-type alloy they used for comparison purpose: 1.38 wt%). However, such alloy system still needs major improvement, since the decomposition pressure is very low, ranging from about 10−2 to 0.4 atm, which is still impractical for solid-state hydrogen storage applications.

6. Solid Solutions

Metallurgically speaking, the term ”solid solution alloy” designates a primary element (solvent) into which one or more minor elements (solutes) are dissolved. Unlike the intermetallic compound, the solute does not need to be present at an integer or near-integer stoichiometric ratio and is present in a random (disordered) substitutional or interstitial distribution within the basic crystal structure. Several solid solution alloys form reversible hydrides, in particular those based on Pd, Ti, Zr, Nb and V solvents [145].
Despite excellent properties such as fast absorption/desorption kinetics and large hydrogen gravimetric density of maximum 3.8 wt% at moderate temperatures, V-based alloys suffer major drawbacks preventing their rapid and widespread applications. These limitations are (i) the relatively difficult first activation, and (ii) the high thermal stability of its hydride phases yielding poor cyclic performance (reversible capacity down to ~2 wt% H2 at room temperature) [49,146].
Upon hydrogenation, V forms a solid solution α followed by β phase (V2H with body-centered tetragonal structure) and then the γ phase (VH2 with CaF2 crystal structure), whose respective thermal stability drastically differs. Indeed, the β phase is so stable that its hydrogen desorption reaction never occurs under moderate conditions, its desorption pressure usually ranging 10−5–0.1 atm. On the other hand, the γ phase is not as stable as its β counterpart and its hydrogen absorption/desorption reaction occurs at moderate temperatures and pressures (over 1 atm at room temperature). Therefore, due to the stability of the β phase, only about half of the amount of hydrogen absorbed in vanadium metal can be used in the hydrogen absorption and desorption processes under practical conditions [146].
Thermodynamic destabilization of the β phase of pristine V stands out as the main solution to tackle the issues mentioned above. Hence, similarly to any other AxBy alloy category, the use of alloying elements of diverse nature and simultaneous addition (binary, ternary and quaternary systems for instance) can destabilize the hydride phases, by altering the ionicity, electronic density of states and lattice parameters [49].
Binary V-based systems cover a broad range of elements, with Ti being the most studied one in the light of its high solubility in V [147], the improved hydrogenation rates and increased terminal solid solubility (TSS) of hydrogen [49]. Although Ti is widely utilized, other elements such as Si, Al and Fe are also considered, but turn out to decrease the hydrogenation rates [148,149], while Mo addition increases hydrogenation-dehydrogenation pressure and decreases the hydrogen storage capacity for instance [150].
To push further the enhancement brought by binary alloys, ternary systems have been developed, notably V–Ti–Cr which remains the most documented ternary alloy due to its excellent improvement of the cyclic stability (as compared to its former binary V–Ti counterpart) while maintaining high effective capacity at room temperature [151,152,153]. Storage capacity can be controlled and increased by tuning the compositional ratio of those three elements, for instance in a mixture of 60 at% V, 15 at% Ti and 25 at% Cr which reaches as high as 2.62 wt% [154]. V–Ti–Cr alloys however show a steep slope of hydrogen absorption–desorption plateaus, requiring homogenization by heat treatment [155,156] and melt-quenching treatment [157,158]. Besides, the formation of an enriched Ti phase during heat treatment and the oxidation of Ti during melt-quenching both reduce the amount of stored hydrogen and complicate the activation process [159].
In spite of the attractive storage capacity of V–Ti–Cr alloys, they remain expensive since the price of pure V is very high. Fe can thus be used as a replacement of V in ternary systems, and excellent storage capacity of 3.9 wt% with a reversible capacity of 2.4 wt% are reported for Ti43.5V49Fe7.5 (at 253 K) [160]. Fe also shows a great potential for tailoring plateau pressures, for instance in (V0.9Ti0.1)1−xFe alloys (with x = 0–0.075) [161]. The reduction of costs by Fe addition has also been attempted for quaternary alloys, notably by Luo et al. [162] who synthesized V48Fe12Ti15Cr25. The maximum hydrogen storage capacity of this alloy reached 1.98 wt% at 315 K, which is lower than that of other V–Ti–Cr series alloys, due to smaller lattice constant and cell volume.
The lattice constant of the alloys is closely related to the amount of hydrogen absorbed/desorbed [154,163,164]: V48Fe12Ti15Cr25 has smaller interstitial sites, which could lead to a lower hydrogen storage capacity, higher plateau pressure, and smaller hysteresis. Similar to Fe addition, the use of Ce is shown by Liu et al. [165] to improve the flatness of plateau of the Ti32Cr46V22 BCC alloy, as a result of the microstructural homogenization during heat-treatment (Ce also increases the hydrogen capacity by lowering the oxygen concentration). The heat-treated Ti32Cr46V22Ce0.4 alloy can release 2.00 and 2.52 wt% H2 at 343 and 298 K, respectively, under 1 atm.
In general, quaternary alloys compile the advantages of the already optimized properties of ternary V–Ti–Cr alloys, and display an improved cyclic stability without noticeable change of the storage capacity after the addition of various atoms such as Fe [166], Nb [167] or even C [168]. However, even more complex systems exist, as shown by Yang et al. [169], who conducted partial substitution studies on V–Ti–Cr–Fe alloys using Co and Zr for improving the storage and cyclic properties. They found out that the hydrogen absorption-desorption capacities of the (VFe)60(TiCrCo)40−xZrx alloys decrease with increasing Zr content. The maximum desorption capacity reaches 2.10 wt% when x = 0, against 1.88 wt% when x = 2. This could be ascribed to the decrease of the volume fraction of the BCC phase while the other phases increase with the Zr content. At the same time, the rate of cyclic degradation decreases with higher Zr content, from 10.9% after 10 cycles (for x = 0) down to 4.5% (when x = 2). Moreover, as the Zr content increases, the hydriding incubation period shortens from 120 s for x = 0 down to 4 s for x = 2. Additionally, more than 90% of the maximum hydrogen absorption capacity is achieved in 400 s when x = 0, while only about 150 s when x = 2. Figure 8 shows the desorption behaviour of some representative solid solution alloys described in this section (see Table 7 for more information on the plotted alloys).
In summary, the effect of partial substitution on the microstructure and subsequent hydrogenation properties (plateau pressure, storage capacity, hysteresis and so on) shown in this section are not limited to solid solutions like Ti–V–Cr mentioned earlier. Similar observations can be made on other classes of alloys, as described earlier in this review. In principle, the substitution by a smaller element (smaller radius) leads to a smaller cell volume and induces an increased plateau pressure, for instance in (Ti,Mn)-based AB2-type alloys (Fe substitution for Mn [95], or simply by increasing Mn content to contract the lattice [170]), in La–Mg–Ni-based AB3 alloys (Y substitution for La [139]), or in V solid solution (Fe substitution for V [146]). On the other hand, the use of larger radius enlarges the cell volume and decreases the plateau pressure, for LaNi5 (increasing Zn substitution for Ni) [76], for TiFe (Mn substitution for Fe) [119], and also for La–Mg–Ni-based AB3 alloy (Co, Mo and Al substitution for Ni) [141,142,143] for example. It is however difficult to generalize the trends from this non-exhaustive list, notably to establish clear effects on the storage capacity. Indeed, since the substitution may lead to the formation of multicomponent systems, the formation of secondary phases may additionally alter the storage properties (see Table 3, Table 4, Table 5, Table 6 and Table 7 for a more detailed case-by-case comparison).

7. Conclusions

Hydrogen is a sustainable energy carrier that can totally redefine and transform the future global energy industry. The actual barrier for implementing hydrogen economy is not only the lack of adequate infrastructures, but also the safe and long-term storage methods. As described in this review, solid-state storage systems based on intermetallic compounds and solid solutions are recognized as one of the most feasible solutions to store hydrogen for hydrogen-powered systems. Overall, the alloys described here cannot store large quantities of hydrogen (most gravimetric densities being around and under 2 wt%). Therefore, developing new kinds of metal hydrides with larger hydrogen storage capacities remains a significant challenge for scientists and engineers. In this review, we evaluated and compared several alloys and presented the most successful and promising modifications aiming to improve their hydrogen sorption properties. In an attempt to overview the most promising representatives of each alloy category, both classical and recent publications have been reviewed. In our opinion, Mm0.9Ca0.1Ni4.6Fe0.3Al0.1 (AB5), (Ti0.65Zr0.35)1.05MnCr0.8Fe0.2 and Ti1.02Cr1.1Mn0.3Fe0.6La0.03 (low/high pressure AB2), Ti1.1Fe0.8Mn0.2 (AB), La0.60V0.05Mg1.32Ca1.03Ni9 (AB3), and Ti32Cr46Y22Ce0.4 (solid solutions) all gather some of the best properties among other compounds presented here (see Table 8). The enhancement of room temperature properties they show deserves additional investigations to open the route to further developments.
Partial substitution, even as trivial as a few weight percent of one element for another, can induce drastic changes in all hydrogen-related properties of AxBy alloys. This unique potential for tuning the properties by manipulating the composition enables the development of tailor-made alloys as per specific application targets. Whether for storage or high-pressure compression of hydrogen, AB5 and AB2 alloys illustrate particularly well this fact, which explains the extensive research they have undergone. Hence, tremendous progress has been made especially in the past two decades, leading to actual commercialization. Those are few among many examples, which encourage application-driven research, stimulating the hope to see one day AxBy alloys move from laboratory to industrial-scale in a society based on hydrogen economy.

Author Contributions

Conceptualization, A.L., J.O.F., J.P. and J.-Y.S.; Writing—Original Draft Preparation, A.L., J.O.F. and M.F.; Writing—Review and Editing, J.O.F., J.-Y.S., Y.-S.L., J.-H.S., J.P. and Y.W.C.; Visualization, A.L., J.O.F. and M.F.; Supervision; J.O.F. and J.-Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the “Technology Development Program to Solve Climate Changes” of the National Research Foundation (NRF) funded by the Ministry of Science, ICT & Future Planning (Grant number: 2015M1A2A2074688) and the APC was funded by Korea Institute of Science and Technology (2E30201).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schlapbach, L.; Züttel, A. Hydrogen-Storage Materials for Mobile Applications. In Materials for Sustainable Energy: A Collection of Peer-Reviewed Research and Review Articles from Nature Publishing Group; World Scientific: Singpore, 2011; pp. 265–270. [Google Scholar]
  2. Orimo, S.; Matsushima, T.; Fujii, H.; Fukunaga, T.; Majer, G.; Züttel, A.; Schlapbach, L. Nanostructured graphite-hydrogen systems prepared by mechanical milling method. Mol. Cryst. Liq. Cryst. 2002, 386, 173–178. [Google Scholar] [CrossRef]
  3. Von Colbe, J.B.; Ares, J.-R.; Barale, J.; Baricco, M.; Buckley, C.; Capurso, G.; Gallandat, N.; Grant, D.M.; Guzik, M.N.; Jacob, I.; et al. Application of hydrides in hydrogen storage and compression: Achievements, outlook and perspectives. Int. J. Hydrog. Energy 2019, 44, 7780–7808. [Google Scholar] [CrossRef]
  4. Liu, T.; Wang, C.; Wu, Y. Mg-based nanocomposites with improved hydrogen storage performances. Int. J. Hydrog. Energy 2014, 39, 14262–14274. [Google Scholar] [CrossRef]
  5. Trudeau, M.L. Advanced Materials for Energy Storage. MRS Bull. 1999, 24, 23–26. [Google Scholar] [CrossRef]
  6. Wang, P.; Kang, X.-D. ChemInform Abstract: Hydrogen-Rich Boron-Containing Materials for Hydrogen Storage. ChemInform 2009, 40, 5400–5413. [Google Scholar] [CrossRef]
  7. Hong, S.-H.; Song, M.Y. Hydrogen desorption and absorption properties of Pd and MgO or nano-sized Ni-added MgH2 + LiBH4 composites. Mater. Res. Bull. 2013, 48, 3453–3458. [Google Scholar] [CrossRef]
  8. Rusman, N.; Dahari, M. A review on the current progress of metal hydrides material for solid-state hydrogen storage applications. Int. J. Hydrog. Energy 2016, 41, 12108–12126. [Google Scholar] [CrossRef]
  9. Jain, I.; Lal, C.; Jain, A. Hydrogen storage in Mg: A most promising material. Int. J. Hydrog. Energy 2010, 35, 5133–5144. [Google Scholar] [CrossRef]
  10. Dolotko, O.; Gupta, S.; Kobayashi, T.; McDonald, E.; Hlova, I.; Majzoub, E.; Balema, V.P.; Pruski, M.; Pecharsky, V.K. Mechanochemical reactions and hydrogen storage capacities in MBH4–SiS2 systems (MLi or Na). Int. J. Hydrog. Energy 2019, 44, 7381–7391. [Google Scholar] [CrossRef] [Green Version]
  11. Jiang, Z.; Yuan, J.; Han, H.; Wu, Y. Effect of carbon nanotubes on the microstructural evolution and hydrogen storage properties of Mg(BH4). J. Alloy. Compd. 2018, 743, 11–16. [Google Scholar] [CrossRef]
  12. Afonso, G.; Bonakdarpour, A.; Wilkinson, D.P. Hydrogen Storage Properties of the Destabilized 4NaBH4/5Mg2NiH4 Composite System. J. Phys. Chem. C 2013, 117, 21105–21111. [Google Scholar] [CrossRef]
  13. Javadian, P.; Zlotea, C.; Ghimbeu, C.M.; Latroche, M.; Jensen, T.R. Hydrogen Storage Properties of Nanoconfined LiBH4–Mg2NiH4 Reactive Hydride Composites. J. Phys. Chem. C 2015, 119, 5819–5826. [Google Scholar] [CrossRef]
  14. Severa, G.; Ronnebro, E.; Jensen, C.M. Direct hydrogenation of magnesium boride to magnesium borohydride: Demonstration of >11 weight percent reversible hydrogen storage. Chem. Commun. 2010, 46, 421–423. [Google Scholar] [CrossRef] [PubMed]
  15. Duan, C.; Cao, Y.; Hu, L.; Fu, D.; Ma, J.; Youngblood, J. An efficient mechanochemical synthesis of alpha-aluminum hydride: Synergistic effect of TiF3 on the crystallization rate and selective formation of alpha-aluminum hydride polymorph. J. Hazard. Mater. 2019, 373, 141–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bogdanović, B.; Hartwig, T.; Spliethoff, B. The development, testing and optimization of energy storage materials based on the MgH2 Mg system. Int. J. Hydrog. Energy 1993, 18, 575–589. [Google Scholar] [CrossRef]
  17. Barkhordarian, G.; Klassen, T.; Bormann, R. Kinetic investigation of the effect of milling time on the hydrogen sorption reaction of magnesium catalyzed with different Nb2O5 contents. J. Alloy. Compd. 2006, 407, 249–255. [Google Scholar] [CrossRef]
  18. Barkhordarian, G.; Klassen, T.; Bormann, R. Fast hydrogen sorption kinetics of nanocrystalline Mg using Nb2O5 as catalyst. Scr. Mater. 2003, 49, 213–217. [Google Scholar] [CrossRef]
  19. Oelerich, W.; Klassen, T.; Bormann, R. Comparison of the catalytic effects of V, V2O5, VN, and VC on the hydrogen sorption of nanocrystalline Mg. J. Alloy. Compd. 2001, 322, L5–L9. [Google Scholar] [CrossRef]
  20. Sargent, A.-L.; Sheppard, D.A.; Paskevicius, M.; Pistidda, C.; Dornheim, M.; Buckley, C.E. Reaction kinetic behaviour with relation to crystallite/grain size dependency in the Mg–Si–H system. Acta Mater. 2015, 95, 244–253. [Google Scholar] [CrossRef] [Green Version]
  21. Vajo, J.J.; Mertens, F.; Ahn, C.C.; Bowman, J.R.C.; Fultz, B. Altering Hydrogen Storage Properties by Hydride Destabilization through Alloy Formation: LiH and MgH2 Destabilized with Si. J. Phys. Chem. B 2004, 108, 13977–13983. [Google Scholar] [CrossRef]
  22. Su, W.; Zhu, Y.; Zhang, J.; Liu, Y.; Yang, Y.; Mao, Q.; Li, L. Effect of multi-wall carbon nanotubes supported nano-nickel and TiF3 addition on hydrogen storage properties of magnesium hydride. J. Alloy. Compd. 2016, 669, 8–18. [Google Scholar] [CrossRef]
  23. Asselli, A.A.C.; Botta, W.J.; Huot, J. Formation reaction of Mg2FeH6: Effect of hydrogen absorption/desorption kinetics. Mater. Res. 2013, 16, 1373–1378. [Google Scholar] [CrossRef] [Green Version]
  24. Huot, J.; Hayakawa, H.; Akiba, E. Preparation of the hydrides Mg2FeH6 and Mg2CoH5 by mechanical alloying followed by sintering. J. Alloy. Compd. 1997, 248, 164–167. [Google Scholar] [CrossRef]
  25. Bogdanovic, B.; Hofmann, H.; Neuy, A.; Reiser, A.; Schlichte, K.; Spliethoff, B.; Wessel, S. Ni-doped versus undoped Mg-MgH2 materials for high temperature heat or hydrogen storage. J. Alloy. Compd. 1999, 292, 57–71. [Google Scholar] [CrossRef]
  26. Felderhoff, M.; Bogdanović, B. High Temperature Metal Hydrides as Heat Storage Materials for Solar and Related Applications. Int. J. Mol. Sci. 2009, 10, 325–344. [Google Scholar] [CrossRef] [PubMed]
  27. Puszkiel, J.; Riglos, M.V.C.; Ramallo-López, J.M.; Mizrahi, M.; Gemming, T.; Pistidda, C.; Larochette, P.A.; Von Colbe, J.B.; Klassen, T.; Dornheim, M.; et al. New Insight on the Hydrogen Absorption Evolution of the Mg–Fe–H System under Equilibrium Conditions. Metals 2018, 8, 967. [Google Scholar] [CrossRef] [Green Version]
  28. Witek, K.; Karczewski, K.; Karpowicz, M.; Polański, M. Mg2FeH6 Synthesis Efficiency Map. Crystals 2018, 8, 94. [Google Scholar] [CrossRef] [Green Version]
  29. Polański, M.; Nawra, D.; Zasada, D. Mg2FeH6 synthesized from plain steel and magnesium hydride. J. Alloy. Compd. 2019, 776, 1029–1040. [Google Scholar] [CrossRef]
  30. Fadonougbo, J.O.; Jung, J.-Y.; Suh, J.-Y.; Lee, Y.-S.; Shim, J.-H.; Cho, Y.W. Low temperature formation of Mg2FeH6 by hydrogenation of ball-milled nano-crystalline powder mixture of Mg and Fe. Mater. Des. 2017, 135, 239–245. [Google Scholar] [CrossRef]
  31. Jung, J.Y.; Fadonougbo, J.O.; Suh, J.-Y.; Lee, Y.-S.; Huh, J.-Y.; Cho, Y.W. Synthesis of Mg2FeH6 by hydrogenation of Mg/Fe powder mixture prepared by cold roll milling in air: Effects of microstructure and oxygen distribution. Int. J. Hydrog. Energy 2018, 43, 16758–16765. [Google Scholar] [CrossRef]
  32. Fadonougbo, J.O.; Jung, J.-Y.; Suh, J.-Y.; Lee, Y.-S.; Shim, J.-H.; Fleury, E.; Cho, Y.W. The role of Fe particle size and oxide distribution on the hydrogenation properties of ball-milled nano-crystalline powder mixtures of Fe and Mg. J. Alloy. Compd. 2019, 806, 1039–1046. [Google Scholar] [CrossRef]
  33. Bobet, J. Study of Mg–M (M = Co, Ni and Fe) mixture elaborated by reactive mechanical alloying: Hydrogen sorption properties. Int. J. Hydrog. Energy 2001, 26, 493–501. [Google Scholar] [CrossRef]
  34. Rzeszotarska, M.; Czujko, T.; Polański, M. Mg2(Fe, Cr, Ni)HX complex hydride synthesis from austenitic stainless steel and magnesium hydride. Int. J. Hydrog. Energy 2020, 45, 19440–19454. [Google Scholar] [CrossRef]
  35. Fadonougbo, J.O.; Kim, H.-J.; Suh, B.-C.; Suh, J.-Y.; Lee, Y.-S.; Shim, J.-H.; Yim, C.D.; Cho, Y.W. Kinetics and thermodynamics of near eutectic Mg–Mg2Ni composites produced by casting process. Int. J. Hydrog. Energy 2020, 45, 29009–29022. [Google Scholar] [CrossRef]
  36. 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]
  37. Iba, H.; Akiba, E. Hydrogen absorption and modulated structure in Ti–V–Mn alloys. J. Alloy. Compd. 1997, 253, 21–24. [Google Scholar] [CrossRef]
  38. Principi, G.; Agresti, F.; Maddalena, A.; Russo, S.L. The problem of solid state hydrogen storage. Energy 2009, 34, 2087–2091. [Google Scholar] [CrossRef]
  39. US Department of Energy. DOE Technical Targets for Onboard Hydrogen Storage for Light-Duty Vehicles; Office of Energy Efficiency and Renewable Energy: Washington, DC, USA, 2017.
  40. Hirscher, M.; Yartys, V.; Baricco, M.; Von Colbe, J.B.; Blanchard, D.; Bowman, R.C.; Broom, D.P.; Buckley, C.E.; Chang, F.; Chen, P.; et al. Materials for hydrogen-based energy storage—past, recent progress and future outlook. J. Alloy. Compd. 2020, 827, 153548. [Google Scholar] [CrossRef]
  41. Moradi, R.; Groth, K.M. Hydrogen storage and delivery: Review of the state of the art technologies and risk and reliability analysis. Int. J. Hydrog. Energy 2019, 44, 12254–12269. [Google Scholar] [CrossRef]
  42. 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. Hydrog. Energy 2017, 42, 289–311. [Google Scholar] [CrossRef]
  43. Abe, J.; Popoola, A.; Ajenifuja, E.; Popoola, O. Hydrogen energy, economy and storage: Review and recommendation. Int. J. Hydrog. Energy 2019, 44, 15072–15086. [Google Scholar] [CrossRef]
  44. Bannenberg, L.; Heere, M.; Benzidi, H.; Montero, J.; Dematteis, E.; Suwarno, S.; Jaroń, T.; Winny, M.; Orłowski, P.; Wegner, W. Metal (boro-) hydrides for high energy density storage and relevant emerging technologies. Int. J. Hydrog. Energy 2020, 45, 33687–33730. [Google Scholar] [CrossRef]
  45. Yartys, V.; Lototskyy, M.; Akiba, E.; Albert, R.; Antonov, V.; Ares, J.; Baricco, M.; Bourgeois, N.; Buckley, C.; Von Colbe, J.B.; et al. Magnesium based materials for hydrogen based energy storage: Past, present and future. Int. J. Hydrog. Energy 2019, 44, 7809–7859. [Google Scholar] [CrossRef]
  46. Crivello, J.-C.; Denys, R.V.; Dornheim, M.; Felderhoff, M.; Grant, D.M.; Huot, J.; Jensen, T.R.; De Jongh, P.E.; Latroche, M.; Walker, G.; et al. Mg-based compounds for hydrogen and energy storage. Appl. Phys. A 2016, 122, 1–17. [Google Scholar] [CrossRef] [Green Version]
  47. Baran, A.; Polański, M. Magnesium-Based Materials for Hydrogen Storage—A Scope Review. Materials 2020, 13, 3993. [Google Scholar] [CrossRef]
  48. Liu, W.; Webb, C.; Gray, E. Review of hydrogen storage in AB3 alloys targeting stationary fuel cell applications. Int. J. Hydrog. Energy 2016, 41, 3485–3507. [Google Scholar] [CrossRef]
  49. Kumar, S.; Jain, A.; Ichikawa, T.; Kojima, Y.; Dey, G. Development of vanadium based hydrogen storage material: A review. Renew. Sustain. Energy Rev. 2017, 72, 791–800. [Google Scholar] [CrossRef]
  50. Klebanoff, L. Hydrogen storage technology: Materials and applications. Choice Rev. Online 2013, 50, 50. [Google Scholar] [CrossRef] [Green Version]
  51. Borzone, E.; Baruj, A.; Blanco, M.; Meyer, G. Dynamic measurements of hydrogen reaction with LaNi5−xSnx alloys. Int. J. Hydrog. Energy 2013, 38, 7335–7343. [Google Scholar] [CrossRef] [Green Version]
  52. Modibane, K.; Lototskyy, M.V.; Davids, M.; Williams, M.; Hato, M.; Molapo, K. Influence of co-milling with palladium black on hydrogen sorption performance and poisoning tolerance of surface modified AB5-type hydrogen storage alloy. J. Alloy. Compd. 2018, 750, 523–529. [Google Scholar] [CrossRef]
  53. Pan, H.; Liu, Y.; Gao, M.; Zhu, Y.; Lei, Y. The structural and electrochemical properties of La0.7Mg0.3 (Ni0.85Co0.15)x (x = 3.0–5.0) hydrogen storage alloys. Int. J. Hydrog. Energy 2003, 28, 1219–1228. [Google Scholar] [CrossRef]
  54. Sharma, V.K.; Kumar, E.A. Effect of measurement parameters on thermodynamic properties of La-based metal hydrides. Int. J. Hydrog. Energy 2014, 39, 5888–5898. [Google Scholar] [CrossRef]
  55. Prigent, J.; Joubert, J.-M.; Gupta, M. Modification of the hydrogenation properties of LaNi5 upon Ni substitution by Rh, Ir, Pt or Au. J. Alloy. Compd. 2012, 511, 95–100. [Google Scholar] [CrossRef]
  56. Georgiadis, M.C.; Kikkinides, E.; Makridis, S.S.; Kouramas, K.; Pistikopoulos, E.N. Design and optimization of advanced materials and processes for efficient hydrogen storage. Comput. Chem. Eng. 2009, 33, 1077–1090. [Google Scholar] [CrossRef]
  57. Liu, J.J.; Li, K.; Cheng, H.H.; Yan, K.; Wang, Y.; Liu, Y.; Jin, H.M.; Zheng, Z. New insights into the hydrogen storage performance degradation and Al functioning mechanism of LaNi5−xAlx alloys. Int. J. Hydrog. Energy 2017, 42, 24904–24914. [Google Scholar] [CrossRef]
  58. Zhu, Y.; Yang, C.; Zhu, J.; Li, L. Structural and electrochemical hydrogen storage properties of Mg2Ni-based alloys. J. Alloy. Compd. 2011, 509, 5309–5314. [Google Scholar] [CrossRef]
  59. Pęska, M.; Dworecka-Wójcik, J.; Płociński, T.; Polański, M. The Influence of Cerium on the Hydrogen Storage Properties of La1−xCexNi5 Alloys. Energies 2020, 13, 1437. [Google Scholar] [CrossRef] [Green Version]
  60. Chen, J.; Dou, S.; Liu, H.K. Effect of partial substitution of La with Ce, Pr and Nd on the properties of LaNi5-based alloy electrodes. J. Power Sources 1996, 63, 267–270. [Google Scholar] [CrossRef]
  61. Balasubramaniam, R.; Mungole, M.; Rai, K. Hydriding properties of MmNi5 system with aluminium, manganese and tin substitutions. J. Alloy. Compd. 1993, 196, 63–70. [Google Scholar] [CrossRef]
  62. Mungole, M. Hysteresis in MmNi5 systems with aluminium, manganese and tin substitutions. Int. J. Hydrog. Energy 1995, 20, 151–157. [Google Scholar] [CrossRef]
  63. Molinas, B.; Pontarollo, A.; Scapin, M.; Peretti, H.; Melnichuk, M.; Corso, H.; Aurora, A.; Gattia, D.M.; Montone, A. The optimization of MmNi5−xAlx hydrogen storage alloy for sea or lagoon navigation and transportation. Int. J. Hydrog. Energy 2016, 41, 14484–14490. [Google Scholar] [CrossRef]
  64. Iosub, V.; Latroche, M.; Joubert, J.-M.; Percheron-Guégan, A. Optimisation of MmNi5−xSnx (Mm = La, Ce, Nd and Pr, 0.27 < x < 0.5) compositions as hydrogen storage materials. Int. J. Hydrog. Energy 2006, 31, 101–108. [Google Scholar]
  65. Zhou, W.; Tang, Z.; Zhu, D.; Ma, Z.; Wu, C.; Huang, L.; Chen, Y. Low-temperature and instantaneous high-rate output performance of AB5-type hydrogen storage alloy with duplex surface hot-alkali treatment. J. Alloy. Compd. 2017, 692, 364–374. [Google Scholar] [CrossRef]
  66. Srivastava, S.; Upadhyaya, R. Investigations on synthesis, characterization and hydrogenation behavior of hydrogen storage alloys, Mm1−xCaxNi5−y−zAlyFez (x = 0, 0.05, 0.1, 0.2, 0.3; y = 0, 0.1; z = 0, 0.1) Mm1−xCaxNi5−y−zAlyFez (x = 0, 0.05, 0.1, 0.2, 0.3; y = 0, 0.1; z = 0, 0.1). Int. J. Hydrog. Energy 2007, 32, 4195–4201. [Google Scholar] [CrossRef]
  67. Srivastava, S.; Upadhyaya, R. Investigations of AB5-type hydrogen storage materials with enhanced hydrogen storage capacity. Int. J. Hydrog. Energy 2011, 36, 7114–7121. [Google Scholar] [CrossRef]
  68. Sandrock, G. A New Family of Hydrogen Storage Alloys Based on the System Nickel-Mischmetal-Calcium. In Proceedings of the 12th Intersociety Energy Conversion Engineering Conference, Washington, DC, USA, 28 August–2 September 1977; pp. 951–958. [Google Scholar]
  69. Shinar, J.; Shaltiel, D.; Davidov, D.; Grayevsky, A. Hydrogen sorption properties of the La1−xCaxNi5 and La (Ni1−xCux)5 systems. J. Less Comm. Metals 1978, 60, 209–219. [Google Scholar] [CrossRef]
  70. Wang, X.; Chen, C.; Wang, C.; Wang, Q. Hydrogen storage properties of Ml1−xCaxNi5 pseudobinary intermetallic compounds. J. Alloy. Compd. 1996, 232, 192–196. [Google Scholar] [CrossRef]
  71. Kumar, E.A.; Sharma, V. Simulation of Pressure Concentration Isotherms of MmNi5−xAlx Hydrides. In Proceedings of the 13th International Conference on Fuel Cell & Hydrogen Technologies, Kuala Lumpur, Malaysia, 22–23 November 2011. [Google Scholar]
  72. Sarma, V. On the mechanically pulverized MmNi4.6Fe0.4 as a viable hydrogen storage material. Int. J. Hydrog. Energy 2001, 26, 231–236. [Google Scholar] [CrossRef]
  73. Apostolov, A.; Stanev, N.; Tcholakov, P. Hydrogen desorption characteristics of MmNi5−xFex compounds. J. Less Common Metals 1985, 110, 127–129. [Google Scholar] [CrossRef]
  74. An, X.; Gu, Q.; Zhang, J.; Chen, S.; Yu, X.; Li, Q. Experimental investigation and thermodynamic reassessment of La–Ni and LaNi5–H systems. Calphad 2013, 40, 48–55. [Google Scholar] [CrossRef]
  75. Verbetsky, V.; Sirotina, R.; Umerenko, E. Absorption of hydrogen by MmNi5 alloys. Int. J. Hydrog. Energy 1996, 21, 935–938. [Google Scholar] [CrossRef]
  76. Rozdzynska-Kielbik, B.; Iwasieczko, W.; Drulis, H.; Pavlyuk, V.; Bala, H. Hydrogenation equilibria characteristics of LaNi5−xZnx intermetallics. J. Alloy. Compd. 2000, 298, 237–243. [Google Scholar] [CrossRef]
  77. Abu Dakka, M.I.; Jain, I. Comparative study of hydrogen in La(28.9)Ni(67.55)Si(3.55) and LaNi. Int. J. Hydrog. Energy 2000, 25, 773–777. [Google Scholar] [CrossRef]
  78. Bowman, R., Jr.; Luo, C.; Ahn, C.; Witham, C.; Fultz, B. The effect of tin on the degradation of LaNi5−ySny metal hydrides during thermal cycling. J. Alloy. Compd. 1995, 217, 185–192. [Google Scholar] [CrossRef]
  79. Briki, C.; Belkhiria, S.; Dhaou, M.H.; De Rango, P.; Jemni, A. Experimental study of the influences substitution from Ni by Co, Al and Mn on the hydrogen storage properties of LaNi3.6Mn0.3Al0.4Co0.7 alloy. Int. J. Hydrog. Energy 2017, 42, 10081–10088. [Google Scholar] [CrossRef]
  80. Yilmaz, F.; Ergen, S.; Hong, S.-J.; Uzun, O. Effect of Bismuth on hydrogen storage properties of melt-spun LaNi4.7−xAl0.3Bix (x = 0.0, 0.1, 0.2, 0.3) ribbons. Int. J. Hydrog. Energy 2018, 43, 20243–20251. [Google Scholar] [CrossRef]
  81. Bououdina, M.; Grant, D.M.; Walker, G. Review on hydrogen absorbing materials—structure, microstructure, and thermodynamic properties. Int. J. Hydrog. Energy 2006, 31, 177–182. [Google Scholar] [CrossRef]
  82. Kim, M.-H.; Debnath, M.R.; Wang, Y.-I.; Suh, J.-Y.; Fleury, E.; Kim, D. Effect of Co on the degradation of the hydrogen permeability of Ni–Nb–Zr amorphous membranes. Metals Mater. Int. 2014, 20, 215–219. [Google Scholar] [CrossRef]
  83. Zhao, Y.; Choi, I.-C.; Seok, M.-Y.; Kim, M.-H.; Kim, D.-H.; Ramamurty, U.; Suh, J.-Y.; Jang, J.-I. Effect of hydrogen on the yielding behavior and shear transformation zone volume in metallic glass ribbons. Acta Mater. 2014, 78, 213–221. [Google Scholar] [CrossRef]
  84. Zhao, Y.; Choi, I.-C.; Seok, M.-Y.; Ramamurty, U.; Suh, J.-Y.; Jang, J.-I. Hydrogen-induced hardening and softening of Ni–Nb–Zr amorphous alloys: Dependence on the Zr content. Scr. Mater. 2014, 93, 56–59. [Google Scholar] [CrossRef]
  85. Fadonougbo, J.O.; Suh, J.-Y.; Han, S.; Shim, C.-H.; Kim, G.-H.; Kim, M.-H.; Fleury, E.; Cho, Y.W. Hydrogen-induced decomposition of Cu–Zr binary amorphous metallic alloys. J. Alloy. Compd. 2016, 660, 456–460. [Google Scholar] [CrossRef]
  86. Fadonougbo, J.O.; Suh, J.-Y.; Shim, C.-H.; Kim, G.-H.; Fleury, E.; Cho, Y.W. Nanometer-scale phase separation and formation of delta ZrH2 in Cu–Zr binary amorphous alloys. J. Alloy. Compd. 2017, 721, 646–652. [Google Scholar] [CrossRef]
  87. Park, J.-G.; Jang, H.-Y.; Han, S.-C.; Lee, P.S.; Lee, J.-Y. The thermodynamic properties of Ti–Zr–Cr–Mn laves phase alloys. J. Alloy. Compd 2001, 325, 293–298. [Google Scholar] [CrossRef]
  88. Liu, B.-H.; Kim, N.-M.; Lee, K.-Y.; Lee, J.-Y. Hydrogen storage properties of TiMn2-based alloys. J. Alloy. Compd. 1996, 240, 214–218. [Google Scholar] [CrossRef]
  89. Kojima, Y.; Kawai, Y.; Towata, S.-I.; Matsunaga, T.; Shinozawa, T.; Kimbara, M. Development of metal hydride with high dissociation pressure. J. Alloy. Compd. 2006, 419, 256–261. [Google Scholar] [CrossRef]
  90. Kandavel, M.; Bhat, V.; Rougier, A.; Aymard, L.; Nazri, G.-A.; Tarascon, J.-M. Improvement of hydrogen storage properties of the AB2 Laves phase alloys for automotive application. Int. J. Hydrog. Energy 2008, 33, 3754–3761. [Google Scholar] [CrossRef]
  91. Morii, K.; Shimizu, T. Hydriding characteristics in (Ti,Zr)(Ni,Mn,X)2 alloys. J. Alloy. Compd. 1995, 231, 524–527. [Google Scholar] [CrossRef]
  92. Manickam, K.; Grant, D.M.; Sardari, P.T. Optimization of AB2 type alloy composition with superior hydrogen storage properties for stationary applications. Int. J. Hydrog. Energy 2015, 40, 16288–16296. [Google Scholar] [CrossRef]
  93. Li, J.; Jiang, X.; Li, G.; Li, X. Development of Ti1.02Cr2−x−yFexMny (0.6 ≤ x ≤ 0.75, y = 0.25, 0.3) alloys for high hydrogen pressure metal hydride system. Int. J. Hydrog. Energy 2019, 44, 15087–15099. [Google Scholar] [CrossRef]
  94. Lee, J.-Y.; Lee, K.-Y.; Lee, H.-H.; Kim, D.-M.; Yu, J.-S.; Jung, J.-H.; Lee, S.-G. Hydrogen-Storage Material Employing Ti-Mn Alloy System. U.S. Patent 5,888,317A, 30 March 1999. [Google Scholar]
  95. Li, J.; Xu, L.; Jiang, X.; Li, X. Study on the hydrogen storage property of (TiZr0.1)xCr1.7−y FeyMn0.3 (1.05 < x < 1.2, 0.2 < y < 0.6) alloys. Prog. Nat. Sci. Mater. Int. 2018, 28, 470–477. [Google Scholar]
  96. Corgnale, C.; Sulic, M. Techno-Economic Analysis of High-Pressure Metal Hydride Compression Systems. Metals 2018, 8, 469. [Google Scholar] [CrossRef] [Green Version]
  97. Pickering, L.; Lototskyy, M.V.; Davids, M.W.; Sita, C.; Linkov, V. Induction melted AB2-type metal hydrides for hydrogen storage and compression applications. Mater. Today Proc. 2018, 5, 10470–10478. [Google Scholar] [CrossRef]
  98. Ulmer, U.; Dieterich, M.; Pohl, A.; Dittmeyer, R.; Linder, M.P.; Fichtner, M. Study of the structural, thermodynamic and cyclic effects of vanadium and titanium substitution in laves-phase AB2 hydrogen storage alloys. Int. J. Hydrog. Energy 2017, 42, 20103–20110. [Google Scholar] [CrossRef] [Green Version]
  99. Yao, Z.; Liu, L.; Xiao, X.; Wang, C.; Jiang, L.; Chen, L. Effect of rare earth doping on the hydrogen storage performance of Ti1.02Cr1.1Mn0.3Fe0.6 alloy for hybrid hydrogen storage application. J. Alloy. Compd. 2018, 731, 524–530. [Google Scholar] [CrossRef]
  100. Puszkiel, J.; Bellosta von Colbe, J.M.; Jepsen, J.; Mitrokhin, S.V.; Movlaev, E.; Verbetsky, V.; Klassen, T. Designing an AB2-Type Alloy (TiZr-CrMnMo) for the Hybrid Hydrogen Storage Concept. Energies 2020, 13, 2751. [Google Scholar] [CrossRef]
  101. Wu, T.; Xue, X.; Zhang, T.; Hu, R.; Kou, H.; Li, J. Role of Ni addition on hydrogen storage characteristics of ZrV2 Laves phase compounds. Int. J. Hydrog. Energy 2016, 41, 10391–10404. [Google Scholar] [CrossRef]
  102. Qin, C.; Zhou, C.; Ouyang, L.; Liu, J.; Zhu, M.; Sun, T.; Wang, H. High-pressure hydrogen storage performances of ZrFe2 based alloys with Mn, Ti, and V addition. Int. J. Hydrog. Energy 2020, 45, 9836–9844. [Google Scholar] [CrossRef]
  103. Zhou, C.; Wang, H.; Ouyang, L.; Liu, J.; Zhu, M. Achieving high equilibrium pressure and low hysteresis of Zr–Fe based hydrogen storage alloy by Cr/V substitution. J. Alloy. Compd. 2019, 806, 1436–1444. [Google Scholar] [CrossRef]
  104. Blasius, A.; Gonster, U. Mössbauer surface studies on Tife hydrogen storage material. Appl. Phys. A 1980, 22, 331–332. [Google Scholar] [CrossRef]
  105. Reilly, J.J.; Wiswall, R.H. Formation and properties of iron titanium hydride. Inorg. Chem. 1974, 13, 218–222. [Google Scholar] [CrossRef]
  106. Sandrock, G.D.; Reilly, J.J.; Johnson, J.R. Metallurgical Considerations in the Production and use of FeTi Alloys for Hydrogen Storage. In Proceedings of the Intersociety Energy Conversion Engineering Conference, State Line, NV, USA, 12 September 1976. [Google Scholar]
  107. Edalati, K.; Matsuda, J.; Iwaoka, H.; Toh, S.; Akiba, E.; Horita, Z. High-pressure torsion of TiFe intermetallics for activation of hydrogen storage at room temperature with heterogeneous nanostructure. Int. J. Hydrog. Energy 2013, 38, 4622–4627. [Google Scholar] [CrossRef]
  108. Davids, M.W.; Lototskyy, M.V.; Nechaev, A.; Naidoo, Q.; Williams, M.; Klochko, Y. 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]
  109. Qu, H.; Du, J.; Pu, C.; Niu, Y.; Huang, T.; Li, Z.; Lou, Y.; Wu, Z. Effects of Co introduction on hydrogen storage properties of Ti–Fe–Mn alloys. Int. J. Hydrog. Energy 2015, 40, 2729–2735. [Google Scholar] [CrossRef]
  110. Yamashita, I.; Tanaka, H.; Takeshita, H.; Kuriyama, N.; Sakai, T.; Uehara, I. Hydrogenation characteristics of TiFe1−xPdx (0.05 ≤ x ≤ 0.30) alloys. J. Alloy. Compd. 1997, 253, 238–240. [Google Scholar] [CrossRef]
  111. Jang, T.; Han, J.; Jai-Young, L. Effect of substitution of titanium by zirconium in TiFe on hydrogenation properties. J. Less Common Metals 1986, 119, 237–246. [Google Scholar] [CrossRef]
  112. Bronca, V.; Bergman, P.; Ghaemmaghami, V.; Khatamian, D.; Manchester, F. Hydrogen absorption characteristics of an FeTi + misch metal alloy. J. Less Common Metals 1985, 108, 313–325. [Google Scholar] [CrossRef]
  113. Guéguen, A.; Latroche, M. Influence of the addition of vanadium on the hydrogenation properties of the compounds TiFe0.9Vx and TiFe0.8Mn0.1Vx (x = 0, 0.05 and 0.1). J. Alloy. Compd. 2011, 509, 5562–5566. [Google Scholar] [CrossRef]
  114. Lee, S.M.; Perng, T.P. Microstructural Correlations with the Hydrogenation Kinetics of FeTi1+ε Alloys. J. Alloy. Compd. 1991, 177, 107–118. [Google Scholar] [CrossRef]
  115. Emami, H.; Edalati, K.; Matsuda, J.; Akiba, E.; Horita, Z. Hydrogen storage performance of TiFe after processing by ball milling. Acta Mater. 2015, 88, 190–195. [Google Scholar] [CrossRef]
  116. Chiang, C.-H.; Chin, Z.-H.; Perng, T.-P. Hydrogenation of TiFe by high-energy ball milling. J. Alloy. Compd. 2000, 307, 259–265. [Google Scholar] [CrossRef]
  117. Edalati, K.; Matsuda, J.; Yanagida, A.; Akiba, E.; Horita, Z. Activation of TiFe for hydrogen storage by plastic deformation using groove rolling and high-pressure torsion: Similarities and differences. Int. J. Hydrog. Energy 2014, 39, 15589–15594. [Google Scholar] [CrossRef]
  118. Johnson, J.; Reilly, J. The Use of Manganese Substituted Ferrotitanium Alloys for Energy Storage. In Proceedings of the The Use of Manganese Substituted Ferrotitanium Alloys for Energy Storage, Miami Beach, FL, USA, 5–7 December 1977. [Google Scholar]
  119. Shang, H.; Zhang, M.; Li, Y.; Qi, Y.; Guo, S.; Zhao, D. Effects of adding over-stoichiometrical Ti and substituting Fe with Mn partly on structure and hydrogen storage performances of TiFe alloy. Renew. Energy 2019, 135, 1481–1498. [Google Scholar] [CrossRef]
  120. Wu, C.Y.; Li, J.C. Phase-Structure of the Ti1cu1-Xfex System. Metall. Trans. A Phys. Metall. Mater. Sci. 1989, 20, 981–985. [Google Scholar] [CrossRef]
  121. Dematteis, E.M.; Cuevas, F.; Latroche, M. Hydrogen storage properties of Mn and Cu for Fe substitution in TiFe0.9 intermetallic compound. J. Alloy. Compd. 2021, 851, 156075. [Google Scholar] [CrossRef]
  122. Nagai, H.; Kitagaki, K.; Shoji, K.-I. Hydrogen Storage Characteristics of FeTi Containing Zirconium. Trans. Jpn. Inst. Metals 1988, 29, 494–501. [Google Scholar] [CrossRef] [Green Version]
  123. Lee, S.-M.; Perng, T.-P. Correlation of substitutional solid solution with hydrogenation properties of TiFe1−x Mx (M= Ni, Co, Al) alloys. J. Alloy. Compd. 1999, 291, 254–261. [Google Scholar] [CrossRef]
  124. Kuziora, P.; Kunce, I.; McCain, S.; Adkins, N.J.E.; Polański, M. The influence of refractory metals on the hydrogen storage characteristics of FeTi-based alloys prepared by suspended droplet alloying. Int. J. Hydrog. Energy 2020, 45, 21635–21645. [Google Scholar] [CrossRef]
  125. Jain, P.; Gosselin, C.; Huot, J. Effect of Zr, Ni and Zr 7 Ni 10 alloy on hydrogen storage characteristics of TiFe alloy. Int. J. Hydrog. Energy 2015, 40, 16921–16927. [Google Scholar] [CrossRef]
  126. Yang, T.; Wang, P.; Xia, C.Q.; Liu, N.; Liang, C.Y.; Yin, F.X.; Li, Q. Effect of chromium, manganese and yttrium on microstructure and hydrogen storage properties of TiFe-based alloy. Int. J. Hydrog. Energy 2020, 45, 12071–12081. [Google Scholar] [CrossRef]
  127. Ha, T.; Lee, S.-I.; Hong, J.; Lee, Y.-S.; Kim, D.-I.; Suh, J.-Y.; Cho, Y.W.; Hwang, B.; Lee, J.; Shim, J.-H. Hydrogen storage behavior and microstructural feature of a TiFe–ZrCr2 alloy. J. Alloy. Compd. 2021, 853, 157099. [Google Scholar] [CrossRef]
  128. Jung, J.Y.; Lee, Y.-S.; Suh, J.-Y.; Huh, J.-Y.; Cho, Y.W. Tailoring the equilibrium hydrogen pressure of TiFe via vanadium substitution. J. Alloy. Compd. 2021, 854, 157263. [Google Scholar] [CrossRef]
  129. Ouyang, L.; Huang, J.; Wang, H.; Liu, 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]
  130. Liu, Y.; Pan, H.; Yue, Y.; Wu, X.; Chen, N.; Lei, Y. Cycling durability and degradation behavior of La–Mg–Ni–Co-type metal hydride electrodes. J. Alloy. Compd. 2005, 395, 291–299. [Google Scholar] [CrossRef]
  131. Liao, B.; Lei, Y.; Chen, L.; Lü, G.; Pan, H.; Wang, Q. The effect of Al substitution for Ni on the structure and electrochemical properties of AB3-type La2Mg(Ni1−xAlx)9 (x = 0–0.05) alloys. J. Alloy. Compd. 2005, 404, 665–668. [Google Scholar] [CrossRef]
  132. Takeshita, T.; Wallace, W.E.; Craig, R.S. Solubility of hydrogen in rare earth-tricobalt compounds. Inorg. Chem. 1974, 13, 2283–2284. [Google Scholar] [CrossRef]
  133. Oesterreicher, H.; Clinton, J.; Bittner, H. Hydrides of La–Ni Compounds. Mater. Res. Bull. 1976, 11, 1241–1247. [Google Scholar] [CrossRef]
  134. Kadir, K.; Sakai, T.; Uehara, I. Synthesis and structure determination of a new series of hydrogen storage alloys; RMg2Ni9 (R = La, Ce, Pr, Nd, Sm and Gd) built from MgNi2 Laves-type layers alternating with AB5 layers. J. Alloy. Compd. 1997, 257, 115–121. [Google Scholar] [CrossRef]
  135. Kadir, K.; Sakai, T.; Uehara, I. Structural investigation and hydrogen capacity of YMg2Ni9 and (Y0.5Ca0.5)(MgCa)Ni9: New phases in the AB2C9 system isostructural with LaMg2Ni9. J. Alloy. Compd. 1999, 287, 264–270. [Google Scholar] [CrossRef]
  136. Kadir, K.; Sakai, T.; Uehara, I. Structural investigation and hydrogen storage capacity of LaMg2Ni9 and (La0.65Ca0.35)(Mg1.32Ca0.68)Ni9 of the AB2C9 type structure. J. Alloy. Compd. 2000, 302, 112–117. [Google Scholar] [CrossRef]
  137. Chen, J.; Kuriyama, N.; Takeshita, H.; Tanaka, H.; Sakai, T.; Haruta, M. Hydrogen Storage Alloys with PuNi3-Type Structure as Metal Hydride Electrodes. Electrochem. Solid State Lett. 2000, 3, 249–252. [Google Scholar] [CrossRef]
  138. Lim, K.L.; Liu, Y.; Zhang, Q.-A.; Chan, S.L.I. Effects of partial substitutions of cerium and aluminum on the hydrogenation properties of La (0.65−x) CexCa1.03Mg1.32Ni (9−y) Aly alloy. Int. J. Hydrog. Energy 2014, 39, 10537–10545. [Google Scholar] [CrossRef]
  139. Xin, G.; Wang, S.; Yuan, H.; Yang, K.; Jiang, L.; Liu, X. Promising hydrogen storage properties of cost-competitive La(Y)–Mg–Ca–Ni AB3-type alloys for stationary applications. RSC Adv. 2016, 6, 21742–21748. [Google Scholar] [CrossRef]
  140. Zang, J.; Zhang, Q.; Sun, D. Hydrogen storage performances of RCaMgNi9 (R = Nd, Gd and Er) compounds. J. Alloy. Compd. 2019, 794, 45–52. [Google Scholar] [CrossRef]
  141. Liu, Y.; Pan, H.; Gao, M.; Li, R.; Lei, Y. Effect of Co content on the structural and electrochemical properties of the La0.7Mg0.3Ni3.4−xMn0.1Cox hydride alloys: II. Electrochemical properties. J. Alloy. Compd. 2004, 376, 304–313. [Google Scholar] [CrossRef]
  142. Liao, B.; Lei, Y.; Chen, L.; Lü, G.; Pan, H.; Wang, Q. Effect of Co substitution for Ni on the structural and electrochemical properties of La2Mg(Ni1−xCox)9 (x = 0.1–0.5) hydrogen storage electrode alloys. Electrochim. Acta 2004, 50, 1057–1063. [Google Scholar] [CrossRef]
  143. Zhang, X.; Sun, D.; Yin, W.; Chai, Y.; Zhao, M. Crystallographic and electrochemical characteristics of La0.7Mg0.3Ni3.5−x (Al0.5Mo0.5)x (x = 0–0.8) hydrogen storage alloys. J. Power Sources 2006, 154, 290–297. [Google Scholar] [CrossRef]
  144. Yan, H.; Xiong, W.; Wang, L.; Li, B.; Li, J.; Zhao, X. Investigations on AB3-, A2B7-and A5B19-type LaYNi system hydrogen storage alloys. Int. J. Hydrog. Energy 2017, 42, 2257–2264. [Google Scholar] [CrossRef]
  145. Sandrock, G. A panoramic overview of hydrogen storage alloys from a gas reaction point of view. J. Alloy. Compd. 1999, 293, 877–888. [Google Scholar] [CrossRef]
  146. Yukawa, H.; Takagi, M.; Teshima, A.; Morinaga, M. Alloying effects on the stability of vanadium hydrides. J. Alloy. Compd. 2002, 330, 105–109. [Google Scholar] [CrossRef]
  147. Seo, C.-Y.; Kim, J.-H.; Lee, P.S.; Lee, J.-Y. Hydrogen storage properties of vanadium-based b.c.c. solid solution metal hydrides. J. Alloy. Compd. 2003, 348, 252–257. [Google Scholar] [CrossRef]
  148. Kumar, S.; Taxak, M.; Krishnamurthy, N. Hydrogen absorption kinetics of V–Al alloy. J. Therm. Anal. Calorim. 2013, 112, 5–10. [Google Scholar] [CrossRef]
  149. Kumar, S.; Taxak, M.; Krishnamurthy, N. Synthesis and hydrogen absorption kinetics of V4Cr4Ti alloy. J. Therm. Anal. Calorim. 2012, 112, 51–57. [Google Scholar] [CrossRef]
  150. Verbetsky, V.N.; Zotov, T.A.; Movlaev, E.A. Absorption of hydrogen by V–Mo and V–Mo–Ti alloys. Inorg. Mater. Appl. Res. 2014, 5, 70–74. [Google Scholar] [CrossRef]
  151. Tamura, T.; Kazumi, T.; Kamegawa, A.; Takamura, H.; Okada, M. Protium absorption properties and protide formations of Ti–Cr–V alloys. J. Alloy. Compd. 2003, 356, 505–509. [Google Scholar] [CrossRef]
  152. Okada, M.; Kuriiwa, T.; Tamura, T.; Takamura, H.; Kamegawa, A. Ti–V–Cr bcc alloys with high protium content. J. Alloy. Compd. 2002, 330, 511–516. [Google Scholar] [CrossRef]
  153. Kagawa, A.; Ono, E.; Kusakabe, T.; Sakamoto, Y. Absorption of Hydrogen by Vanadium-Rich V–Ti-Based Alloys. J. Less Common Metals 1991, 172, 64–70. [Google Scholar] [CrossRef]
  154. Tsukahara, M. Hydrogenation Properties of Vanadium-Based Alloys with Large Hydrogen Storage Capacity. Mater. Trans. 2011, 52, 68–72. [Google Scholar] [CrossRef] [Green Version]
  155. Tamura, T.; Tominaga, Y.; Matsumoto, K.; Fuda, T.; Kuriiwa, T.; Kamegawa, A.; Takamura, H.; Okada, M. Protium absorption properties of Ti–V–Cr–Mn alloys with a bcc structure. J. Alloy. Compd. 2002, 330, 522–525. [Google Scholar] [CrossRef]
  156. Shudo, Y.; Ebisawa, T.; Itoh, H. Characterization of Ti–Zr–Mn–V-based Laves phase alloys for MH refrigeration system. J. Alloy. Compd. 2003, 356, 497–500. [Google Scholar] [CrossRef]
  157. Yu, X.; Wu, Z.; Xia, B.; Xu, N. Hydrogen storage performance of quenched Ti–V-based alloy. J. Alloy. Compd. 2004, 373, 134–136. [Google Scholar] [CrossRef]
  158. Yu, X.; Wu, Z.; Xu, N. Effects of melt-quenching rates on the hydrogen storage properties of Ti-based BCC phase alloy. Phys. B Condens. Matter 2004, 344, 456–461. [Google Scholar] [CrossRef]
  159. Yu, X.; Wu, Z.; Xia, B.; Xu, N. Improvement of activation performance of the quenched Ti–V-based BCC phase alloys. J. Alloy. Compd. 2005, 386, 258–260. [Google Scholar] [CrossRef]
  160. Nomura, K.; Akiba, E. H2 Absorbing-desorbing characterization of the TiVFe alloy system. J. Alloy. Compd. 1995, 231, 513–517. [Google Scholar] [CrossRef]
  161. Lynch, J.F.; Maeland, A.J.; Libowitz, G.G. Lattice Parameter Variation and Thermodynamics of Dihydride Formation in the Vanadium-Rich V–Ti–Fe/H2 System. Z. Phys. Chem. 1985, 145, 51–59. [Google Scholar] [CrossRef]
  162. Luo, L.; Li, Y.; Zhai, T.; Hu, F.; Zhao, Z.; Bian, X.; Wu, W. Microstructure and hydrogen storage properties of V48Fe12Ti15-xCr25Alx (x = 0, 1) alloys. Int. J. Hydrog. Energy 2019, 44, 25188–25198. [Google Scholar] [CrossRef]
  163. Akiba, E.; Iba, H. Hydrogen absorption by Laves phase related BCC solid solution. Intermetallics 1998, 6, 461–470. [Google Scholar] [CrossRef]
  164. Yan, Y.; Chen, Y.; Liang, H.; Wu, C.; Tao, M.; Mingjing, T. Effect of Al on hydrogen storage properties of V30Ti35Cr25Fe10 alloy. J. Alloy. Compd. 2006, 426, 253–255. [Google Scholar] [CrossRef]
  165. Liu, X.; Jiang, L.; Li, Z.; Huang, Z.; Wang, S. Improve plateau property of Ti32Cr46V22 BCC alloy with heat treatment and Ce additive. J. Alloy. Compd. 2009, 471, L36–L38. [Google Scholar] [CrossRef]
  166. Aoki, M.; Noritake, T.; Ito, A.; Ishikiriyama, M.; Towata, S.-I. Improvement of cyclic durability of Ti–Cr–V alloy by Fe substitution. Int. J. Hydrog. Energy 2011, 36, 12329–12332. [Google Scholar] [CrossRef]
  167. Towata, S.-I.; Noritake, T.; Itoh, A.; Aoki, M.; Miwa, K. Effect of partial niobium and iron substitution on short-term cycle durability of hydrogen storage Ti–Cr–V alloys. Int. J. Hydrog. Energy 2013, 38, 3024–3029. [Google Scholar] [CrossRef]
  168. Shen, C.-C.; Li, H.-C. Cyclic hydrogenation stability of γ-hydrides for Ti25V35Cr40 alloys doped with carbon. J. Alloy. Compd. 2015, 648, 534–539. [Google Scholar] [CrossRef]
  169. Yang, S.; Yang, F.; Wu, C.; Chen, Y.; Mao, Y.; Luo, L. Hydrogen storage and cyclic properties of (VFe)60(TiCrCo)40−xZrx (0≤ x ≤ 2) alloys. J. Alloy. Compd. 2016, 663, 460–465. [Google Scholar] [CrossRef]
  170. Nayebossadri, S.; Book, D. Development of a high-pressure Ti–Mn based hydrogen storage alloy for hydrogen compression. Renew. Energy 2019, 143, 1010–1021. [Google Scholar] [CrossRef]
Figure 1. Flow chart representation of hydrogen storage methods (adapted from Hydrogen Storage Technology Materials and Applications, Page 67 [50]).
Figure 1. Flow chart representation of hydrogen storage methods (adapted from Hydrogen Storage Technology Materials and Applications, Page 67 [50]).
Hydrogen 01 00004 g001
Figure 2. Desorption pressure-composition isotherms (PCI) of Mm0.9Ca0.1Ni5−xAl0.1Fex at 300 K [67] with desorption PCI of reference LaNi5 at 303 K [74] and MmNi5 at 273 K [75].
Figure 2. Desorption pressure-composition isotherms (PCI) of Mm0.9Ca0.1Ni5−xAl0.1Fex at 300 K [67] with desorption PCI of reference LaNi5 at 303 K [74] and MmNi5 at 273 K [75].
Hydrogen 01 00004 g002
Figure 3. Desorption pressure-composition isotherms of TiMn1.5 [94], (Ti1−xZrx)1+yMn0.8Cr1.2 [87], Ti0.8Zr0.1Mn1.2Cr0.2V0.1Fe0.1 [94], (Ti0.9Zr0.1)1.1Cr1.5Fe0.2Mn0.3 [95], Ti0.2Zr0.8Ni1.3Mn0.7 and Ti0.4Zr0.6Ni1.1Mn0.6V0.1Fe0.2 at 303 K [91]. (Ti0.65Zr0.35)1.05MnCr0.8Fe0.2 on the other hand was measured at 305 K [92].
Figure 3. Desorption pressure-composition isotherms of TiMn1.5 [94], (Ti1−xZrx)1+yMn0.8Cr1.2 [87], Ti0.8Zr0.1Mn1.2Cr0.2V0.1Fe0.1 [94], (Ti0.9Zr0.1)1.1Cr1.5Fe0.2Mn0.3 [95], Ti0.2Zr0.8Ni1.3Mn0.7 and Ti0.4Zr0.6Ni1.1Mn0.6V0.1Fe0.2 at 303 K [91]. (Ti0.65Zr0.35)1.05MnCr0.8Fe0.2 on the other hand was measured at 305 K [92].
Hydrogen 01 00004 g003
Figure 4. Recently reported desorption pressure-composition isotherms of Ti0.98Zr0.02V0.43Fe0.09Cr0.05Mn1.5 and its FeV and Ti sponge counterpart acquired at 298 K [98], Ti1.02Cr1.1Mn0.3Fe0.6 and Ti1.02Cr1.1Mn0.3Fe0.6La0.03 measured at 263 K [99], (Ti0.9Zr0.1)1.25Cr0.85Mn1.1Mo0.05 at 296 K [100], Zr1.05Fe1.6Mn0.4 and (Zr0.5Ti0.5)1.05Fe0.95MnV0.05 at 288 K [102], and finally Zr1.05Fe1.85Cr0.075V0.075 at 288 K [103].
Figure 4. Recently reported desorption pressure-composition isotherms of Ti0.98Zr0.02V0.43Fe0.09Cr0.05Mn1.5 and its FeV and Ti sponge counterpart acquired at 298 K [98], Ti1.02Cr1.1Mn0.3Fe0.6 and Ti1.02Cr1.1Mn0.3Fe0.6La0.03 measured at 263 K [99], (Ti0.9Zr0.1)1.25Cr0.85Mn1.1Mo0.05 at 296 K [100], Zr1.05Fe1.6Mn0.4 and (Zr0.5Ti0.5)1.05Fe0.95MnV0.05 at 288 K [102], and finally Zr1.05Fe1.85Cr0.075V0.075 at 288 K [103].
Hydrogen 01 00004 g004
Figure 5. The effect of oxygen content (in the inlet gas) on the hydrogen desorption isotherm of the reference un-modified TiFe alloy at 313 K [106]. Oxygen content ranges from 0.010 to 0.87 wt%.
Figure 5. The effect of oxygen content (in the inlet gas) on the hydrogen desorption isotherm of the reference un-modified TiFe alloy at 313 K [106]. Oxygen content ranges from 0.010 to 0.87 wt%.
Hydrogen 01 00004 g005
Figure 6. Desorption pressure-composition isotherms of TiFe and Ti0.9Zr0.1Fe at 303 K [111], and Ti1.1Fe0.8Mn0.2 at 313 K [119]. The graph also plots the PCI of alloys with varying Fe/Ti ratios at 313 K: (a) 60.5 wt% Fe and 39.5 wt% Ti, (b) 50.5 wt% Fe and 49.2 wt% Ti, (c) 36.7 wt% Fe and 63.2 wt% Ti [105] and of TiFe alloy with 4 wt% Zr [125]. Finally, TiFe0.9Ni0.1, TiFe0.9Co0.1 and TiFe0.9Al0.1 at 323 K [123].
Figure 6. Desorption pressure-composition isotherms of TiFe and Ti0.9Zr0.1Fe at 303 K [111], and Ti1.1Fe0.8Mn0.2 at 313 K [119]. The graph also plots the PCI of alloys with varying Fe/Ti ratios at 313 K: (a) 60.5 wt% Fe and 39.5 wt% Ti, (b) 50.5 wt% Fe and 49.2 wt% Ti, (c) 36.7 wt% Fe and 63.2 wt% Ti [105] and of TiFe alloy with 4 wt% Zr [125]. Finally, TiFe0.9Ni0.1, TiFe0.9Co0.1 and TiFe0.9Al0.1 at 323 K [123].
Hydrogen 01 00004 g006
Figure 7. Desorption pressure-composition isotherms of LaCaMgNi9 at 293 K [137], (La0.65Ca0.35)(Mg1.32Ca0.68)Ni9 at 283 K [136] and La(0.65−x)YxMg1.32Ca1.03Ni9 at 298 K [139].
Figure 7. Desorption pressure-composition isotherms of LaCaMgNi9 at 293 K [137], (La0.65Ca0.35)(Mg1.32Ca0.68)Ni9 at 283 K [136] and La(0.65−x)YxMg1.32Ca1.03Ni9 at 298 K [139].
Hydrogen 01 00004 g007
Figure 8. Desorption pressure-composition isotherms of V and V–Fe at 313 K [146], Ti32Cr46V22Ce0.4 at 298 K and 318 K [165], V48Fe12Ti15Cr25 at 295 K [162] and (VFe)60(TiCrCo)39.5Zr0.5 at 298 K [169].
Figure 8. Desorption pressure-composition isotherms of V and V–Fe at 313 K [146], Ti32Cr46V22Ce0.4 at 298 K and 318 K [165], V48Fe12Ti15Cr25 at 295 K [162] and (VFe)60(TiCrCo)39.5Zr0.5 at 298 K [169].
Hydrogen 01 00004 g008
Table 1. Storage properties of intermetallic hydrides compared to gas and liquid hydrogen [36].
Table 1. Storage properties of intermetallic hydrides compared to gas and liquid hydrogen [36].
MaterialH DensityEnergy Density
wt%kg m−3MJ kg−1MJ dm−3
Gas H2, 700 bar10042120.05
Liquid H2 (20 K)10071120.08.5
LaNi5H61.4901.710.8
TiFeH21.91052.312.6
MgH27.61109.213.3
Table 2. Some of the most important families of hydride-forming intermetallic compounds, with the corresponding reference alloys and structures [2,37].
Table 2. Some of the most important families of hydride-forming intermetallic compounds, with the corresponding reference alloys and structures [2,37].
Intermetallic CompoundReference AlloyStructure
AB5LaNi5Haucke phase, hexagonal
AB2TiMn2Laves phase, hexagonal or cubic
ABTiFeCubic, CsCl-type or orthorhombic, CrB-type
AB3CeNi3Hexagonal, NbBe3-type
Solid solutionsV, Ti–VBody centered cubic
Table 3. Summary of the main properties of AB5 alloys presented in this section and plotted in Figure 2.
Table 3. Summary of the main properties of AB5 alloys presented in this section and plotted in Figure 2.
AlloyTemperature (K)Structure TypeLattice Parameters (nm)Maximum/Reversible Capacity (wt%)Pa/Pd (atm)Remarks on the Effects of Partial Substitution
LaNi5 [74]303CaCu5a = 0.5020
c = 0.3980
1.56/1.252.90/2.50-
MmNi5 [75].273CaCu5a = 0.4934
c = 0.3998
1.29/1.00-/5.20Easy activation process, increased hysteresis and maximum storage capacity ~20% lower than that of LaNi5.
Mm0.9Ca0.1Ni4.8Fe0.1Al0.1 [67]300CaCu5a = 0.4949
c = 0.4013
1.90/1.25~24.00/18.00Ca enhances the hydrogen storage capacity, reduces the incubation time of the first hydrogenation, increases the absorption rate (with increasing Ca concentration), and reduces the hysteresis. Al reduces the plateau pressure, but also the maximum hydrogen storage capacity. Fe increases the storage capacity and reduces sloping and hysteresis.
Mm0.9Ca0.1Ni4.7Fe0.2Al0.1 [67]300CaCu5a = 0.4952
c = 0.4019
2.20/1.50~16.00/13.00
Mm0.9Ca0.1Ni4.6Fe0.3Al0.1 [67]300CaCu5a = 0.4962
c = 0.4018
1.96/1.65~10.00/5.00
Table 4. Summary of the main properties of AB2 alloys presented in this section and plotted in Figure 3 and Figure 4.
Table 4. Summary of the main properties of AB2 alloys presented in this section and plotted in Figure 3 and Figure 4.
AlloyTemperature (K)Structure TypeLattice Parameters (nm)Maximum/Reversible Capacity (wt%)Pa/Pd (atm)Remarks on the Effects of Partial Substitution
TiMn1.5 [94]303MgZn2a = 0.4878
c = 0.7976
1.42/0.75-/5.45-
(Ti0.8Zr0.2)1.00Mn0.8Cr1.2 [87]303MgZn2a = 0.4902
c = 0.8044
1.79/1.3512.00/10.52Increasing Zr content decreases the equilibrium plateau pressure, accelerates absorption kinetics, and increases the storage capacity. Cr increases hydrogen storage capacity and reduces equilibrium pressure. Lattice strain increases along with Zr/Ti ratio and partially results in important sloping
(Ti0.8Zr0.2)1.05Mn0.8Cr1.2 [87]303MgZn2-1.90/1.558.40/7.54
(Ti0.75Zr0.25)1.00Mn0.8Cr1.2 [87]303MgZn2a = 0.4912
c = 0.8064
1.86/1.508.72/6.99
(Ti0.75Zr0.25)1.05Mn0.8Cr1.2 [87]303MgZn2-1.91/1.605.23/4.47
(Ti0.9Zr0.1)1.1Cr1.5Fe0.2Mn0.3 [95]303MgZn2a = 0.4897
c = 0.8026
1.84/1.302.54/2.17Partial substitution of Mn by Fe shrinks the cell volume and increases the plateau pressure, but the hydrogen capacity does not change noticeably.
Ti0.2Zr0.8Ni1.3Mn0.7 [91]303MgCu2-1.65/1.351.40/0.38V lowers both hysteresis and plateau pressure. Ni raises the plateau pressure and reduces the plateau width, while Fe flattens and lengthens it.
Ti0.4Zr0.6Ni1.1Mn0.6V0.1Fe0.2 [91]303MgCu2-1.64/1.251.22/0.75
(Ti0.65Zr0.35)1.05MnCr0.8Fe0.2 [92]305MgZn2-2.2/1.755.30/2.80Unit cell volume and storage capacity increase, charge time reduces along with non-stoichiometry on the A site.
Ti0.8 Zr0.1Mn1.2Cr0.2V0.1Fe0.1 [94]303MgZn2-2.03/1.6018.06/9.02-
Ti0.98Zr0.02V0.43Fe0.09Cr0.05Mn1.5 [98]298MgZn2a = 0.4875
c = 0.7994
1.89/1.3623.3/12.1Replacing Ti by Ti sponge does not change the initial storage capacity, whereas it is reduced after substitution of V by FeV. Substitutions do not affect microstructural properties.
Tisp0.98Zr0.02(FeV)0.43Fe0.09Cr0.05Mn1.5 [98]298MgZn2a = 0.4872
c = 0.7989
1.62/1.1223.4/13.8
Ti1.02Cr1.1Mn0.3Fe0.6 [99]263MgZn2a = 0.4854
c = 0.7968
1.38/1.05208/143RE improves the activation behaviour, the sorption properties and storage capacity (saturation reached at RT), but decreases the hydrogen desorption plateau pressure.
Ti1.02Cr1.1Mn0.3Fe0.6La0.03 [99]263MgZn2a = 0.4862
c = 0.7975
1.7/1.2218/137
(Ti0.9Zr0.1)1.25Cr0.85Mn1.1Mo0.05 [100]296MgZn2a = 0.4884
c = 0.801
1.45/1.2554/34The equilibrium pressure increases with Mo amount, mostly due to the larger bulk modulus of Mo (compared to Cr) rather than the increase in cell volume. For Mo higher than x = 0.05, the hydrogen capacity greatly drops.
Zr1.05Fe1.6Mn0.4 [102]288MgCu2a = 0.70861.32/1.1260/24V addition improves the hysteresis, while an adequate Ti amount helps to achieve low slope and high plateau pressure.
(Zr0.5Ti0.5)1.05Fe0.95MnV0.05 [102]288MgZn2a = 0.4935
c = 0.8079
1.64/1.3223/6.7
Zr1.05Fe1.85Cr0.075V0.075 [103]288MgCu2a = 0.70861.54/1.0861/44Cr and V substitutions effectively decrease the equilibrium pressure due to the enlarged unit cell, while the V substitution improves the hysteresis. Too high Cr addition induces a transition from MgCu2 to MgZn2 structure type.
Table 5. Summary of the main properties of AB alloys presented in this section and plotted in Figure 6.
Table 5. Summary of the main properties of AB alloys presented in this section and plotted in Figure 6.
AlloyTemperature (K)Structure TypeLattice Parameters (nm)Maximum/Reversible Capacity (wt%)Pa/Pd (atm)Remarks on the Effects of Partial Substitution
TiFe [111]303CsCl-~1.73/~1.60-/6Enthalpy of β hydride (TiFeH) formation is −5.68 kcal (mol H2)−1 and the enthalpy for hydrogen desorption is 6.31 kcal (mol H2)−1.
Ti1.1Fe0.8Mn0.2 [119]313CsCla = 0.2990~1.75/~1.65-/~1.0 (mid-point)-Dehydrogenation enthalpy is 5.65 kcal (mol H2)−1. The addition of over-stoichiometric Ti leads to reduction in hydrogen storage capacity and substitution of Mn for Fe results in more stable hydride and improved hydrogen storage capacity with easy activation.
Ti36Fe64 (60.5 wt% Fe and 39.5 wt% Ti) [105]313CsCl-~0.80/~0.71-/7.0Second plateau is sloping.
Ti45.5Fe54.5 (50.5 wt% Fe and 49.2 wt% Ti) [105]313CsCl-~1.98/~1.74-/~4 (mid-point)Sloping plateau.
Ti59.6Fe40.4 (36.7 wt% Fe and 63.2 wt% Ti) [105]313CsCl-~2.1/~0.90-/~1.5Second plateau is sloping. The desorption is not complete due to the presence of excess Ti, which forms very stable hydride.
Ti0.9Zr0.1Fe [111]303CsCl-~1.1/~0.99-/3Enthalpy of β hydride formation is −6.25 kcal (mol H2)−1 and the enthalpy for hydrogen desorption is 6.91 kcal (mol H2)−1. Zr substitution increases the hydride stability and reduces the storage capacity.
TiFe0.9Ni0.1 [123]323CsCl-~1.4/~1.27-/~1.0Enthalpy for hydrogen desorption is 8.51 kcal (mol H2)−1. Partial substitution greatly influences the stability of the monohydride, as reflected in the broad enthalpy range reported in this table. Easy activation at RT.
TiFe0.9Co0.1 [123]323CsCl -~1.46/~1.29-/~5.0Enthalpy for hydrogen desorption is 7.32 kcal (mol H2)−1. Easy activation at RT.
TiFe0.9Al0.1 [123]323CsCl a = 0.2997~1.27/~1.15-/(sloping plateau)Sloping plateau. Easy activation at RT.
TiFe + 4 wt% Zr [125]313CsCl a = 0.2983~1.2/~0.83-/(sloping plateau)Addition of Zr results in multiphase alloy (formation of a Zr-rich inter-granular phase), RT activation but incomplete desorption at RT.
Table 6. Summary of the main properties of AB3 alloys presented in this section and plotted in Figure 7.
Table 6. Summary of the main properties of AB3 alloys presented in this section and plotted in Figure 7.
AlloyTemperature (K)Structure TypeLattice Parameters (nm)Maximum/Reversible Capacity (wt%)Pa/Pd (atm)Remarks on the Effects of Partial Substitution
LaCaMgNi9 [137]293NbBe3a = 0.4924
c = 2.3875
1.8/1.253.20/2.64AB3 alloys can be tuned by different material processing methods similarly to AB2 and AB5 alloys.
(La0.65Ca0.35)(Mg1.32Ca0.68)Ni9 [136]283NbBe3a = 0.4952
c = 2.3907
1.87/1.352.21/1.65
La0.65Mg1.32Ca1.03Ni9 [139]298NbBe3a = 0.4961
c = 2.3926
1.83/1.32.28/1.61AB5 phase fraction increases with Y substitution, and becomes the dominant phase inducing a reduction of the maximum capacity. Y substitution significantly increases both the hydrogen absorption/desorption plateau pressures due to the lattice contraction.
La0.60Y0.05Mg1.32Ca1.03Ni9 [139]298NbBe3a = 0.4955
c = 2.3935
1.79/1.45.31/3.11
La0.45Y0.20Mg1.32Ca1.03Ni9 [139]298NbBe3a = 0.4948
c = 2.3942
1.75/1.513.32/8.37
Table 7. Summary of the main properties of Solid Solutions presented in this section and plotted in Figure 8.
Table 7. Summary of the main properties of Solid Solutions presented in this section and plotted in Figure 8.
AlloyTemperature (K)Structure TypeLattice Parameters (nm)Maximum/Reversible Capacity (wt%)Pa/Pd (atm)Remarks on the Effects of Partial Substitution
V [146]313BCC-3.63/1.8-/4.57-
V–Fe [146]313BCC-3.50/1.6-/6.99Fe (smaller radius) increases the plateau pressure and inhibits the diffusion of hydrogen.
Ti32Cr46V22Ce0.4 [165]298
318
BCC-3.63/2.5
3.44/2.0
-/1.41
-/3.40
Cr and Ti additions improve cyclic stability, reaction rates, and terminal solid solubility. Ce increases the hydrogen capacity by lowering the oxygen concentration.
V48Fe12Ti15Cr25 [162]295BCCa = 0.29671.98/1.11.91/1.01Commercial ferrovanadium substitution for V lowers the alloy costs, reduces hydrogen storage capacity, hysteresis and cycle stability, and makes higher plateau pressure.
(VFe)60(TiCrCo)39.5Zr0.5 [169]298BCCa = 0.30813.61/1.6-/1.89Co and Zr enhance the storage and cyclic properties, but the hydrogen absorption/desorption capacities decrease with increasing Zr content. The rate of cyclic degradation decreases with higher Zr content and the hydriding incubation period shortens.
Table 8. Summary of the main properties of most promising alloys presented in each section.
Table 8. Summary of the main properties of most promising alloys presented in each section.
AlloyTemperature (K)Structure TypeLattice Parameters (nm)Maximum/Reversible Storage Capacity (wt%)Pa/Pd (atm)
Mm0.9Ca0.1Ni4.6Fe0.3Al0.1 [67]300CaCu5a = 0.4962
c = 0.4018
1.96/1.6510.00/5.00
(Ti0.65Zr0.35)1.05MnCr0.8Fe0.2 [92]305MgZn2-2.2/1.755.30/2.80
Ti1.02Cr1.1Mn0.3Fe0.6La0.03 [99]263MgZn2a = 0.4862
c = 0.7975
1.7/1.2218/137
Ti1.1Fe0.8Mn0.2 [119]313CsCla = 0.2990~1.75/~1.65-/~1.0 (mid-point)
La0.60Y0.05Mg1.32Ca1.03Ni9 [139]298NbBe3a = 0.4955
c = 2.3935
1.79/1.45.31/3.11
Ti32Cr46V22Ce0.4 [165]298
318
BCC-3.63/2.5
3.44/2.0
-/1.41
-/3.40
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lys, A.; Fadonougbo, J.O.; Faisal, M.; Suh, J.-Y.; Lee, Y.-S.; Shim, J.-H.; Park, J.; Cho, Y.W. Enhancing the Hydrogen Storage Properties of AxBy Intermetallic Compounds by Partial Substitution: A Short Review. Hydrogen 2020, 1, 38-63. https://0-doi-org.brum.beds.ac.uk/10.3390/hydrogen1010004

AMA Style

Lys A, Fadonougbo JO, Faisal M, Suh J-Y, Lee Y-S, Shim J-H, Park J, Cho YW. Enhancing the Hydrogen Storage Properties of AxBy Intermetallic Compounds by Partial Substitution: A Short Review. Hydrogen. 2020; 1(1):38-63. https://0-doi-org.brum.beds.ac.uk/10.3390/hydrogen1010004

Chicago/Turabian Style

Lys, Andrii, Julien O. Fadonougbo, Mohammad Faisal, Jin-Yoo Suh, Young-Su Lee, Jae-Hyeok Shim, Jihye Park, and Young Whan Cho. 2020. "Enhancing the Hydrogen Storage Properties of AxBy Intermetallic Compounds by Partial Substitution: A Short Review" Hydrogen 1, no. 1: 38-63. https://0-doi-org.brum.beds.ac.uk/10.3390/hydrogen1010004

Article Metrics

Back to TopTop