Open access

Thermodynamics of Metal Hydrides: Tailoring Reaction Enthalpies of Hydrogen Storage Materials

Written By

Martin Dornheim

Submitted: 15 November 2010 Published: 02 November 2011

DOI: 10.5772/21662

From the Edited Volume

Thermodynamics - Interaction Studies - Solids, Liquids and Gases

Edited by Juan Carlos Moreno-Pirajan

Chapter metrics overview

9,218 Chapter Downloads

View Full Metrics

1. Introduction

Considering the increasing pollution and exploitation of fossil energy resources, the implementation of new energy concepts is essential for our future industrialized society. Renewable sources have to replace current energy technologies. This shift, however, will not be an easy task. In contrast to current nuclear or fossil power plants renewable energy sources in general do not offer a constant energy supply, resulting in a growing demand of energy storage. Furthermore, fossil fuels are both, energy source as well as energy carrier. This is of special importance for all mobile applications. Alternative energy carriers have to be found. The hydrogen technology is considered to play a crucial role in this respect. In fact it is the ideal means of energy storage for transportation and conversion of energy in a comprehensive clean-energy concept. Hydrogen can be produced from different feedstocks, ideally from water using regenerative energy sources. Water splitting can be achieved by electrolysis, solar thermo-chemical, photoelectrochemical or photobiological processes. Upon reconversion into energy, by using a fuel cell only water vapour is produced, leading to a closed energy cycle without any harmful emissions. Besides stationary applications, hydrogen is designated for mobile applications, e.g. for the zero-emission vehicle. In comparison to batteries hydrogen storage tanks offer the opportunity of fast recharging within a few minutes only and of higher storage densities by an order of magnitude. Hydrogen can be produced from renewable energies in times when feed-in into the electricity grid is not possible. It can be stored in large caverns underground and be utilized either to produce electricity and be fed into the electricity grid again or directly for mobile applications.

However, due to the very low boiling point of hydrogen (20.4 K at 1 atm) and its low density in the gaseous state (90 g/m3) dense hydrogen storage, both for stationary and mobile applications, remains a challenging task. There are three major alternatives for hydrogen storage: compressed gas tanks, liquid hydrogen tanks as well as solid state hydrogen storage such as metal hydride hydrogen tanks. All of these three main techniques have their special advantages and disadvantages and are currently used for different applications. However, so far none of the respective tanks fulfils all the demanded technical requirements in terms of gravimetric storage density, volumetric storage density, safety, free-form, ability to store hydrogen for longer times without any hydrogen losses, cyclability as well as recyclability and costs. Further research and development is strongly required.One major advantage of hydrogen storage in metal hydrides is the ability to store hydrogen in a very energy efficient way enabling hydrogen storage at rather low pressures without further need for liquefaction or compression. Many metals and alloys are able to absorb large amounts of hydrogen. The metal-hydrogen bond offers the advantage of a very high volumetric hydrogen density under moderate pressures, which is up to 60% higher than that of liquid hydrogen (Reilly & Sandrock, 1980).

Depending on the hydrogen reaction enthalpy of the specific storage material during hydrogen uptake a huge amount of heat (equivalent to 15% or more of the energy stored in hydrogen) is generated and has to be removed in a rather short time, ideally to be recovered and used as process heat for different applications depending on quantity and temperature. On the other side, during desorption the same amount of heat has to be applied to facilitate the endothermic hydrogen desorption process – however, generally at a much longer time scale. On one side this allows an inherent safety of such a tank system. Without external heat supply hydrogen release would lead to cooling of the tank and finally hydrogen desorption necessarily stops. On the other side it implies further restrictions for the choice of suitable storage materials. Highest energy efficiencies of the whole tank to fuel combustion or fuel cell system can only be achieved if in case of desorption the energy required for hydrogen release can be supplied by the waste heat generated in case of mobile applications on-board by the hydrogen combustion process and the fuel cell respectively.

Advertisement

2. Basics of hydrogen storage in metal hydrides

Many metals and alloys react reversibly with hydrogen to form metal hydrides according to the reaction (1):

Me + x / 2H 2 « MeH x + Q . E1

Here, Me is a metal, a solid solution, or an intermetallic compound, MeHx is the respective hydride and x the ratio of hydrogen to metal, x=cH [H/Me], Q the heat of reaction. Since the entropy of the hydride is lowered in comparison to the metal and the gaseous hydrogen phase, at ambient and elevated temperatures the hydride formation is exothermic and the reverse reaction of hydrogen release accordingly endothermic. Therefore, for hydrogen release/desorption heat supply is required.

Metals can be charged with hydrogen using molecular hydrogen gas or hydrogen atoms from an electrolyte. In case of gas phase loading, several reaction stages of hydrogen with the metal in order to form the hydride need to be considered. Fig. 1 shows the process schematically.

The first attractive interaction of the hydrogen molecule approaching the metal surface is the Van der Waals force, leading to a physisorbed state. The physisorption energy is typically of the order EPhys ≈ 6 kJ/mol H2. In this process, a gas molecule interacts with several atoms at the surface of a solid. The interaction is composed of an attractive term, which diminishes with the distance of the hydrogen molecule and the solid metal by the power of 6, and a repulsive term diminishing with distance by the power of 12. Therefore, the potential energy of the molecule shows a minimum at approximately one molecular radius. In addition to hydrogen storage in metal hydrides molecular hydrogen adsorption is a second technique to store hydrogen. The storage capacity is strongly related to the temperature and the specific surface areas of the chosen materials. Experiments reveal for carbon-based nanostructures storage capacities of less than 8 wt.% at 77 K and less than 1wt.% at RT and pressures below 100 bar (Panella et al., 2005; Schmitz et al., 2008).

Figure 1.

Reaction of a H2 molecule with a storage material: a) H2 molecule approaching the metal surface. b) Interaction of the H2 molecule by Van der Waals forces (physisorbed state). c) Chemisorbed hydrogen after dissociation. d) Occupation of subsurface sites and diffusion into bulk lattice sites.

In the next step of the hydrogen-metal interaction, the hydrogen has to overcome an activation barrier for the formation of the hydrogen metal bond and for dissociation, see Fig. 1c and 2. This process is called dissociation and chemisorption. The chemisorption energy is typically in the range of EChem ≈ 20 - 150 kJ/mol H2 and thus significantly higher than the respective energy for physisorption which is in the order of 4-6 kJ/mol H2 for carbon based high surface materials (Schmitz et al., 2008).

Figure 2.

Schematic of potential energy curves of hydrogen in molecular and atomic form approaching a metal. The hydrogen molecule is attracted by Van der Waals forces and forms a physisorbed state. Before diffusion into the bulk metal, the molecule has to dissociate forming a chemisorbed state at the surface of the metal (according to Züttel, 2003).

After dissociation on the metal surface, the H atoms have to diffuse into the bulk to form a M-H solid solution commonly referred to as α-phase. In conventional room temperature metals / metal hydrides, hydrogen occupies interstitial sites - tetrahedral or octahedral - in the metal host lattice. While in the first, the hydrogen atom is located inside a tetrahedron formed by four metal atoms, in the latter, the hydrogen atom is surrounded by six metal atoms forming an octahedron, see Fig. 3.

Figure 3.

Octahedral (O) and tetrahedral (T) interstitial sites in fcc-, hcp- and bcc-type metals. (Fukai, 1993).

In general, the dissolution of hydrogen atoms leads to an expansion of the host metal lattice of 2 to 3 Å3 per hydrogen atom, see Fig. 4. Exceptions of this rule are possible, e.g. several dihydride phases of the rare earth metals, which show a contraction during hydrogen loading for electronic reasons.

Figure 4.

Volume expansion of the Nb host metal with increasing H content. (Schober & Wenzl, 1978)

In the equilibrium the chemical potentials of the hydrogen in the gas phase and the hydrogen absorbed in the metal are the same:

1 2 μ g a s = μ m e t a l E2
.

Since the internal energy of a hydrogen molecule is 7/2 kT the enthalpy and entropy of a hydrogen molecule are

h g a s = 7 2 k T E Diss E3
and
s g a s = 7 2 k k ln p p 0 ( T ) with p 0 ( T ) = 8 ( πk T ) 7 2 M H-H 5 2 r H-H 2 h 5 E4

Here k is the Boltzmann constant, T the temperature, p the applied pressure, EDiss the dissociation energy for hydrogen (EDiss = 4.52 eV eV/H2), MH-H the mass of the H2 molecule, rH-H the interatomic distance of the two hydrogen atoms in the H2 molecule.

Consequently the chemical potential of the hydrogen gas is given by

μ g a s = k T ln p p 0 ( T ) E Diss = k T ln p p 0 + μ g a s 0 E5

with p0 = 1.01325 105 Pa.

In the solid solution (α-phase) the chemical potential is accordingly

μ α = h α T s α mit s α = s α conf + s α vibr ,electr E6
.

Here, s α,conf is the configuration entropy which is originating in the possible allocations of NH hydrogen atoms on Nis different interstitial sites:

S α ,conf = k ln N is ! N H !(N is -N H )! E7

and accordingly for small cH using the Stirling approximation we get

s α ,conf = -k ln c H n is - c H E8

with nis being the number of interstitial sites per metal atom: nis = Nis/NMe and c H the number of hydrogen atoms per metal atom: c H = N H/NMe.

Therefore the chemical potential of hydrogen in the solid solution (α-phase) is given by

μ α = h α T ( s α vibr ,electr k ln c H n is c H ) E9

Taking into account the equilibrium condition (2) the hydrogen concentration c H can be determined via

c H n is c H = p p 0 ( T ) e - Δ g s k T with Δ g s = h α T s α vibr + 1 2 μ g 0 E10
or
c H n is c H = p p 0 ( T ) e - Δ G s R T with Δ G s = Δ H s T Δ S E11
.

Here μ g0 is the chemical potential of the hydrogen molecule at standard conditions and R being the molar gas constant.

For very small hydrogen concentrations c H cH << nis in the solid solution phase α the hydrogen concentration is directly proportional to the square root of the hydrogen pressure in the gas phase. This equation is also known as the Sievert’s law, i.e.

c H = 1 K S p E12

with KS being a temperature dependent constant. As the hydrogen pressure is increased, saturation occurs and the metal hydride phase MeH cβ starts to form.

For higher hydrogen pressures/concentrations metal hydride formation occurs.

The conversion from the saturated solution phase to the hydride phase takes place at constant pressure p according to:

Me-H c α | α + 1 2 ( c β c α ) H 2 MeH c β | β + Q α β E13
.

In the equilibrium the chemical potentials of the gas phase, the solid solution phase α and the hydride phase β coincide:

μ α ( p , T , c α ) = μ β ( p , T , c β ) = 1 2 μ gas ( p , T ) = 1 2 k T ln ( p eq ( T ) p 0 ) + 1 2 μ gas 0 E14
.

Following Gibb’s Phase Rule f=c-p+2 with f being the degree of freedom, k being the number of components and p the number of different phases only one out of the four variables p, T, c α, c β is to be considered as independent. Therefore for a given temperature all the other variables are fixed.

Therefore the change in the chemical potential or the Gibbs free energy is just a function of one parameter, i.e. the temperature T:

Δ G = 1 2 R T ln ( p ( T ) p 0 ) E15
.

From this equation follows the frequently-used Van’t Hoff equation (16):

1 2 ln p p 0 = Δ H R T Δ S R E16

The temperature dependent plateau pressure of this two phase field is the equilibrium dissociation pressure of the hydride and is a measure of the stability of the hydride, which commonly is referred to as β-phase.

After complete conversion to the hydride phase, further dissolution of hydrogen takes place as the pressure increases, see Fig. 5.

Figure 5.

Schematic Pressure/Composition Isotherm. The precipitation of the hydride phase β starts when the terminal solubility of the α-phase is reached at the plateau pressure.

Multiple plateaus are possible and frequently observed in composite materials consisting of two hydride forming metals or alloys. The equilibrium dissociation pressure is one of the most important properties of a hydride storage material.

If the logarithm of the plateau pressure is plotted vs 1/T, a straight line is obtained (van’t Hoff plot) as seen in Fig. 6.

Figure 6.

Schematic pcT-diagram and van’t Hoff plot. The α-phase is the solid solution phase, the β-phase the hydride phase. Within the (α − β) two phase region both the metal-hydrogen solution and the hydride phase coexist.

2.1. Conventional metal hydrides

Fig. 7 shows the Van’t Hoff plots of some selected binary hydrides. The formation enthalpy of these hydrides H0 f determines the amount of heat which is released during hydrogen absorption and consequently is to be supplied again in case of desorption. To keep the heat management system simple and to reach highest possible energy efficiencies it is necessary to store the heat of absorption or to get by the waste heat of the accompanying hydrogen utilizing process, e.g. energy conversion by fuel cell or internal combustion system. Therefore the reaction enthalpy has to be as low as possible. The enthalpy and entropy of the hydrides determine the working temperatures and the respective plateau pressures of the storage materials. For most applications, especially for mobile applications, working temperatures below 100 ºC or at least below 150 ºC are favoured. To minimize safety risks and avoid the use of high pressure composite tanks the favourable working pressures should be between 1 and 100 bar.

Figure 7.

Van’t Hoff lines (desorption) for binary hydrides. Box indicates 1-100 atm, 0-100 ºC ranges, taken from Sandrock et al. (Sandrock, 1999).

However, the Van’t Hoff plots shown in Fig. 7 indicate that most binary hydrides do not have the desired thermodynamic properties. Most of them have rather high thermodynamic stabilities and thus release hydrogen at the minimum required pressure of 1 bar only at rather high temperatures (T>300 ºC). The values of their respective reaction enthalpies are in the range of 75 kJ/(mol H2) (MgH2) or even higher. Typical examples are the hydrides of alkaline metals, alkaline earth metals, rare earth metals as well as transition metals of the Sc-, Ti- and V-group. The strongly electropositive alkaline metals like LiH and NaH and CaH2 form saline hydrides, i.e. they have ionic bonds with hydrogen. MgH2 marks the transition between these predominantly ionic hydrides and the covalent hydrides of the other elements in the first two periods.

Examples for high temperature hydrides releasing the hydrogen at pressures of 1 bar at extremely high temperatures (T > 700 ºC) are ZrH2 and LaH2 (Dornheim & Klassen, 2009). ZrH2 for example is characterized by a high volumetric storage density NH. NH values larger than 7 × 1022 hydrogen atoms per cubic centimetre are achievable. This value corresponds to 58 mol H2/l or 116 g/l and has to be compared with the hydrogen density in liquid hydrogen (20 K): 4.2 × 1022 (35 mol H2/l or 70 g/l) and in compressed hydrogen (350 bar / 700 bar): 1.3 / 2.3 × 1022 atoms/cm3 ( 11 mol H2/l or 21 g/l and 19 mol H2/l or 38 g/l respectively). The hydrogen density varies a lot between different hydrides. VH2 for example has an even higher hydrogen density which amounts to 11.4 × 1022 hydrogen atoms per cubic centimetre and accordingly 95 mol H2/l or 190 g/l. As in the case of many other transition metal hydrides Zr has a number of different hydride phases ZrH2-x with a wide variation in the stoichiometry (Hägg, 1931). Their compositions extend from about ZrH1.33 up to the saturated hydride ZrH2. Because of the limited gravimetric storage density of only about 2 wt.% and the negligibly low plateau pressure within the temperature range of 0 – 150 ºC Zr as well as Ti and Hf are not suitable at all as a reversible hydrogen storage material. Thus, they are not useful for reversible hydrogen storage if only the pure binary hydrides are considered (Dornheim & Klassen, 2009).

Libowitz et al. (Libowitz et al., 1958) could achieve a breakthrough in the development of hydrogen storage materials by discovering the class of reversible intermetallic hydrides. In 1958 they discovered that the intermetallic compound ZrNi reacts reversibly with gaseous hydrogen to form the ternary hydride ZrNiH3. This hydride has a thermodynamic stability which is just in between the stable high temperature hydride ZrH2fH0= -169 kJ/mol H2) and the rather unstable NiH (ΔfH0= -8.8 kJmol-1H2). Thus, the intermetallic Zr-Ni bond exerts a strong destabilizing effect on the Zr-hydrogen bond so that at 300 ºC a plateau pressure of 1bar is achieved which has to be compared to 900 ºC in case of the pure binary hydride ZrH2. This opened up a completely new research field. In the following years hundreds of new storage materials with different thermodynamic properties were discovered which generally follow the well-known semi-empirical rule of Miedema (Van Mal et al., 1974):

Δ H ( A n B m H x + y ) = Δ H ( A n H x ) + Δ H ( B m H y ) Δ H ( A n B m ) E17

Around 1970, hydrides with significantly lowered values of hydrogen reaction enthalpies, such as LaNi5 and FeTi but also Mg2Ni were discovered. While 1300 ºC are necessary to reach a desorption pressure of 2 bar in case of the pure high temperature hydride LaH2, in case of LaNi5H6 a plateau pressure of 2 bar is already reached at 20 ºC only. The value of the hydrogen reaction enthalpy is lowered to ΔHLaNi5H6 = 30.9 kJmol-1H2. The respective values for NiH are ΔHf,NiH = 8.8 kJmol-1H2 and Pdiss,NiH,RT=3400 bar.

In the meantime, several hundred other intermetallic hydrides have been reported and a number of interesting compositional types identified (table 1). Generally, they consist of a high temperature hydride forming element A and a non hydride forming element B, see fig. 8.

COMPOSITION A B COMPOUNDS
A2B Mg, Zr Ni, Fe, Co Mg2Ni, Mg2Co, Zr2Fe
AB Ti, Zr Ni, Fe TiNi, TiFe, ZrNi
AB2 Zr, Ti, Y, La V, Cr, Mn, Fe, Ni LaNi2, YNi2,YMn2, ZrCr2, ZrMn2,ZrV2, TiMn2
AB3 La, Y, Mg Ni, Co LaCo3,YNi3,LaMg2Ni9
AB5 Ca, La, Rare Earth Ni, Cu, Co, Pt, Fe CaNi5, LaNi5, CeNi5, LaCu5, LaPt5, LaFe5

Table 1.

Examples of intermetallic hydrides, taken from Dornheim et al. (Dornheim, 2010).

Figure 8.

Hydride and non hydride forming elements in the periodic system of elements.

Even better agreement with experimental results than by use of Miedema’s rule of reversed stability is obtained by applying the semi-empirical band structure model of Griessen and Driessen (Griessen & Driessen, 1984) which was shown to be applicable to binary and ternary hydrides. They found a linear relationship of the heat of formation ΔH = H0 f of a metal hydride and a characteristic energy ΔE of the electronic band structure of the host metal which can be applied to simple metals, noble metals, transition metals, actinides and rare earths:

Δ H = α Δ E + β E18

with ΔE = E F-E S (E F being the Fermi energy and E S the center of the lowest band of the host metal, α = 59.24 kJ (eV mol H2)-1 and β = -270 kJ (mol H2)-1 and ΔE in eV.

As described above, most materials experience an expansion during hydrogen absorption, wherefore structural effects in interstitial metal hydrides play an important role as well. This can be and is taken as another guideline to tailor the thermodynamic properties of interstitial metal hydrides. Among others Pourarian et al. (Pourarian, 1982), Fujitani et al. (Fujitani, 1991) and Yoshida & Akiba (Yoshida, 1995) report about this relationship of lattice parameter or unit cell volume and the respective plateau pressures in different material classes.

Intensive studies let to the discovery of a huge number of different multinary hydrides with a large variety of different reaction enthalpies and accordingly working temperatures. They are not only attractive for hydrogen storage but also for rechargeable metal hydride electrodes and are produced and sold in more than a billion metal hydride batteries per year. Because of the high volumetric density, intermetallic hydrides are utilized as hydrogen storage materials in advanced fuel cell driven submarines, prototype passenger ships, forklifts and hydrogen automobiles as well as auxiliary power units.

2.2. Hydrogen storage in light weight hydrides

Novel light weight hydrides show much higher gravimetric storage capacities than the conventional room temperature metal hydrides. However, currently only a very limited number of materials show satisfying sorption kinetics and cycling behaviour. The most prominent ones are magnesium hydride (MgH2) and sodium alanate (NaAlH4). In both cases a breakthrough in kinetics could be attained in the late 90s of the last century / the early 21st century.

Magnesium hydride is among the most important and most comprehensively investigated light weight hydrides. MgH2 itself has a high reversible storage capacity, which amounts to 7.6 wt.%. Furthermore, magnesium is the eighth most frequent element on the earth and thus comparably inexpensive. Its potential usage initially was hindered because of rather sluggish sorption properties and unfavourable reaction enthalpies. The overall hydrogen sorption kinetics of magnesium-based hydrides is as in case of all hydrides mainly determined by the slowest step in the reaction chain, which can often be deduced e.g. by modelling the sorption kinetics (Barkhordarian et al, 2006; Dornheim et al., 2006). Different measures can be taken to accelerate kinetics. One important factor for the sorption kinetics is the micro- or nanostructure of the material, e.g. the grain or crystallite size. Because of the lower packing density of the atoms, diffusion along grain boundaries is usually faster than through the lattice. Furthermore, grain boundaries are favourable nucleation sites for the formation and decomposition of the hydride phase. A second important parameter is the outer dimension of the material, e.g. in case of powdered material, its particle size. The particle size (a) determines the surface area, which is proportional to the rate of the surface reaction with the hydrogen, and (b) is related to the length of the diffusion path of the hydrogen into and out of the volume of the material. A third major factor by which hydrogen sorption is improved in many hydrogen absorbing systems is the use of suitable additives or catalysts. In case of MgH2 it was shown by Oelerich et al. (Oelerich et al., 2001; Dornheim et al., 2007) that already tiny amounts of transition metal oxides have a huge impact on the kinetics of hydrogen sorption. Using such additives Hanada et al. (Hanada et al., 2007) could show that by using such additives hydrogen uptake in Mg is possible already at room temperature within less than 1 min. The additives often do not just have one single function but multiple functions. Suitable additives can catalyze the surface reaction between solid and gas. Dispersions in the magnesium-based matrix can act as nucleation centres for the hydride or the dehydrogenated phase. Furthermore, different additives, such as liquid milling agents and hard particles like oxides, borides, etc., can positively influence the particle size evolution during the milling process (Pranzas et al., 2006 ; Pranzas et al., 2007; Dornheim et al, 2007) and prevent grain i.e. crystallite growth. More detailed information about the function of such additives in MgH2 is given in (Dornheim et al., 2007). Beyond that, a preparation technique like high-energy ball milling affects both the evolution of certain particle sizes as well as very fine crystallite sizes in the nm range and is also used to intermix the hydride and the additives/catalysts. Thus, good interfacial contact with the light metal hydride as well as a fine dispersion of the additives can be achieved.

As in case of MgH2 dopants play also an important role in the sorption of Na-Al-hydride, the so-called Na-alanate. While hydrogen liberation is thermodynamically favorable at moderate temperatures, hydrogen uptake had not been possible until in 1997 Bogdanovic et al. demonstrated that mixing of NaAlH4 with a Ti-based catalyst leads to a material, which can be reversibly charged with hydrogen (Bogdanovic, 1997). By using a tube vibration mill of Siebtechnik GmbH Eigen et al. (Eigen et al., 2007; Eigen et al., 2008) showed that upscaling of material synthesis is possible: After only 30 min milling under optimised process conditions in such a tube vibration mill in kg scale, fast absorption and desorption kinetics with charging/discharging times of less than 10 min can be obtained. The operation temperatures of this complex hydride are much lower than compared to MgH2 and other light weight hydrides. Fast kinetics is achieved at 100 ºC to 150 ºC which is much less than what is required in case of MgH2, however, still significantly higher than in case of the conventional hydrides which show only a very limited storage capacity. Such hydride working temperatures offer the possibility for combinations of metal hydride tanks based on these complex hydrides with e. g. combustion engines, high temperature PEM fuel cells or other medium to high temperature fuel cells. However, compared to MgH2 the gravimetric hydrogen storage capacity is significantly reduced. Having a maximum theoretical storage capacity of about 5.6 wt. % NaAlH4 exhibits a long term practical storage capacity of 3.5-4.5 wt. % H2 only. Furthermore, in difference to MgH2 NaAlH4 decomposes in two reaction steps upon dehydrogenation which implies two different pressure plateaus instead of just one:

NaAlH 4   1 / 3Na 3 AlH 6 + 2 / 3Al + H 2 ( g )   NaH + Al + 3 / 2H 2 ( g )   E19

The first decomposition step has an equilibrium pressure of 0.1 MPa at 30 ºC, the second step at about 100 ºC (Schüth et al., 2004). A maximum of 3.7 wt.% H2 can be released during the first desorption step, 5.6 wt.% in total. The remaining hydrogen bonded to Na is technically not exploitable due to the high stability of the respective hydride.

While the reaction kinetics was optimized significantly, the desorption enthalpy of NaAlH4 of 37 kJ/molH2 and Na3AlH6 of 47 kJ/mol H2 respectively remains a challenge. For many applications even this value which is much below that of MgH2 is still too large.

Advertisement

3. Tailoring thermodynamics of light weight metal hydrides

While there are plenty of known hydrides with suitable thermodynamics for hydrogen uptake and release at ambient conditions (several bar equilibrium pressure at or nearby room temperature) currently no hydride is known which combines suitable thermodynamics and kinetics with such a high gravimetric storage capacity that a hydrogen storage tank based on such a material could compete with a 700 bar compressed composite vessel in regard to weight. Depending on the working temperature and pressure as well as the reversible gravimetric storage capacity of the selected hydride the achievable capacity of a metal hydride based storage tank is usually better than half of the capacity of the metal hydride bed itself (Buchner & Povel, 1982). Since modern composite pressurized gas tanks meanwhile show gravimetric hydrogen storage capacities of around 4 wt.% according to conservative extrapolations the possible choice of hydrides should be limited to those having the ability to reversibly store at least 6 wt.%H2. All currently known high capacity hydrides, however, show either too small values of the respective reaction enthalpy and are therefore not reversible or would require several thousand bar hydrogen pressure or alternatively electrochemical loading or on the other hand are too stable and have an equilibrium pressure which around room temperature is much below the required pressures. The value of reaction enthalpy aimed at is between 20 and 30 kJ/mol H2. Fig. 9 shows the potentially available hydrogen content of some well known hydrides plotted against their hydrogen reaction enthalpies.

Figure 9.

Theoretically achievable reversible storage capacities and reaction enthalpies of selected hydrides. LaNi5H6 and FeTiH2 are taken as examples for conventional room temperature hydrides. The reaction enthalpies and achievable hydrogen storage capacities are ΔH = -31 kJ/mol H2, CH,max = 1.4 wt.% for LaNi5H6 and for the Fe-Ti system ΔH = -31.5 kJ/mol H2, CH,max = 1.8 wt.%(average over two reaction steps with ΔH(FeTiH2) = -28 kJ/mol H2 and ΔH(FeTiH) = -35 kJ/mol H2 respectively) ( Buchner, 1982 ). The respective values for NaAlH4 are ΔH = -40.5 kJ/mol H2, CH,max = 5.6 wt.%(average over two reaction steps with ΔH(NaAlH4) = -37 kJ/mol H2 and ΔH(NaAl3H6) = -47 kJ/mol H2 (Bogdanovic et al., 2009)), for MgH2: ΔH = -78 kJ/mol H2 (Oelerich, 2000) and CH,max = 7.6 wt.%, for LiBH4: ΔH = -74 kJ/mol H2 (Mauron, 2008) and CH,max = 7.6 wt.%, for Mg(BH4)2: ΔH = -57 kJ/mol H2 (Li, 2008) and CH,max = 14.9 wt.%.

As shown in Fig. 9 none of the plotted hydrides, neither the conventional room temperature hydrides with their rather low gravimetric capacity nor the sophisticated novel chemical hydrides with their unsuitable reaction enthalpy, show the desired combination of properties. Therefore the tailoring of the thermodynamic properties of high capacity light weight and complex hydrides is a key issue, an imperative for future research in the area of hydrides as hydrogen storage materials.

3.1. Thermodynamic tuning of single phase light weight hydrides

The traditional way of tailoring the thermodynamic properties of metal hydrides is by formation of alloys with different stabilities as described in chapter 2.1. Thereby the value of reaction enthalpy can be reduced by stabilising the dehydrogenated state and/or destabilising of the hydride state, see Fig. 10 a. Accordingly, the total amount of reaction enthalpy is increased by destabilising the dehydrogenated state and/or stabilising the hydride, see Fig. 10 b.

This approach has been successfully applied to light weight metal hydrides also.

Mg-based hydrides

One of the first examples using this approach for tuning the thermodynamic properties of light weight metal hydrides was the discovery of the Mg-Ni –system as potential hydrogen storage system by Reilly and Wiswall (Reilly & Wiswall, 1968). Mg2Ni has a negative heat of

Figure 10.

Tailoring of the reaction enthalpy by altering the stability of the hydrogenated or dehydrogenated state of the metal hydrides: a) Reduction of total reaction enthalpy by stabilising the dehydrogenated phase by ΔHds or destabilising the hydride phase by ΔHhd. b) Increase of total reaction enthalpy by destabilising the dehydrogenated state by ΔHdd or stabilising the hydrogenated state by ΔHhs.

formation of H0 f(Mg2Ni) = -42 kJ/mol. Therefore, compared to pure Mg the dehydrogenated state is stabilised by ΔHds = -21 kJ/(mol Mg). The enthalpy of formation of Mg2NiH4 is H0 f(Mg2NiH4) = -176 kJ/mol (= -88 kJ/(mol Mg)), wherefore the hydride phase is stabilised by ΔHhd = -10 kJ/(mol Mg) if compared to pure MgH2. In total the hydrogen reaction enthalpy of Mg2Ni

| Δ H Mg 2 Ni-H | = | Δ H Mg-H | + Δ H ds Δ H hd E20

is reduced by 11 kJ/mol H2 to aboutΔH(Mg2Ni-H) = 67 kJ/mol H2. While pure MgH2 exhibits a hydrogen plateau pressure of 1 bar around 300 ºC, in case of Mg2NiH4 such a plateau pressure is reached already at around 240 ºC and in case of further alloying and substituting Ni by Cu at around 230 ºC in Mg2Ni0.5Cu0.5 (Klassen et al., 1998). Unfortunately, the gravimetric storage capacity of Mg2NiH4 is reduced to 3.6 wt.% H2 only and thus is less than half the respective value in the MgH2 system. Darnaudery et al. (Darnaudery et al., 1983) were successful to form several quaternary hydrides by hydrogenating Mg2Ni0.75M0.25 with different 3d elements M ∈ {V, Cr, Fe, Co and Zn} showing stabilities very similar Mg2NiH4.

Increasing the amount of 3d metals Tsushio et al. (Tsushio et al., 1998) investigated the hydrogenation of MgNi0.86M0.03 with M ∈ {Cr, Fe, Co, Mn}. Consequently, they observed a dramatic decrease in hydrogen storage capacity to 0.9 wt.% and in hydrogen reaction enthalpy which amounts to 50 kJ/(mol H2) for MgNi0.86Cr0.03. This reaction enthalpy value is in very good agreement with the value 54 kJ/(mol H2) given by Orimo et al. for amorphous MgNi (Orimo et al., 1998).

Lowering even more the content of Mg Terashita et al. (Terashita et al., 2001) found (Mg1-xCax)Ni2 based alloys desorbing hydrogen at room temperature. They determined the hydride formation enthalpy and entropy of (Mg0.68Ca0.32)Ni2 to be H = -37 kJ/(mol H2) and S = - 94 J/(mol H2 K) respectively, which is already quite near to the envisioned target. Unfortunately, with lowering the Mg content the hydrogen storage capacity dropped down to 1.4 wt.% only.

On the other side, as schematically shown in Fig. 10b the absolute value of reaction enthalpy can be increased by either stabilising the hydride phase or destabilising the dehydrogenated phase. In case of Mg-based hydrogen absorbing alloys this is not at all of interest for hydrogen storage itself since MgH2 is too stable for most hydrogen storage applications, however, this is of interest for other applications like the storage of thermal energy (Dornheim & Klassen, 2009). Mg2FeH6 is an example of such materials with increased amount of reaction enthalpy. Furthermore, it is the one with the highest known volumetric hydrogen density which amounts to 150 kg m-3. This enormously high hydrogen density is more than double the value found in case of liquid hydrogen at 20 K and moderate pressures of up to 20 bar (Klell, 2010). The gravimetric storage capacity is 5.6 wt.% and thus still rather high. Since Mg and Fe are immiscible the dehydrogenated state is destabilised compared to pure Mg: ΔHdd> 0 kJ/(mol H2). Accordingly the hydride phase is more difficult to be synthesised and reversibility as well as long term stability is more difficult to be accomplished.

Nevertheless, hydrogenation is possible at hydrogen pressures of at least 90 bar and temperatures of at least 450 ºC (Selvam & Yvon, 1991). Bogdanovic et al. (Bogdanovic et al., 2002) achieved very good reversibility and cycling stability with the hydrogen storage capacities remaining unchanged throughout 550-600 cycles at a level of 5-5.2 wt.% H2. The reaction enthalpy value is reported to be in between 77 kJ/(mol H2) and 98 kJ/(mol H2) (Bogdanovic et al., 2002), (Konstanchuk et al, 1987), (Puszkiel et al., 2008), (Didisheim et al., 1984).

The large reaction enthalpies of MgH2 and Mg2FeH6 lead to weight and volume related heat storage densities in the temperature range of 500 C which are many times higher than that of the possible sensible or latent heat storage materials (Bogdanovic et al., 2002). The calculated and experimental heat storage densities to weight given by Bogdanovic et al. are 2814 kJ/kg and 2204 kJ/kg for the MgH2-Mg system and 2106 and 1921 kJ/kg for the Mg2FeH6 – 2Mg+Fe system respectively. The corresponding calculated and experimental values for the volumetric thermal energy storage density are 3996 kJ/dm³ and 1763 kJ/dm³ for the MgH2-Mg system and 5758 kJ/dm³ and 2344 kJ/dm³ respectively (Bogdanovic et al., 2002). These thermal energy densities ought not to be mistaken with the energy stored in the hydrogen (lower heating value) which is more than a factor of three larger.

Aluminum-based complex hydrides

As Mg2FeH6 decomposes during hydrogen release into 2 Mg, Fe and 3 H2 NaAlH4 decomposes during hydrogen release in 1/3 Na3AlH6 + 2/3 Al + H2 and finally NaH + Al + 3/2 H2. As written in chapter 2.2 while much lower than those of the Mg-based hydrides the reaction enthalpies of |ΔH|= 37 kJ/(mol H2) and |ΔH|= 47 kJ/(mol H2) are still two high for many applications especially for the usage in combination with low temperature PEM fuel cells. LiAlH4 on the other hand is much less stable. It decomposes in two steps as is the case of the NaAlH4:

6LiAlH 4 2Li 3 AlH 6 + 4Al + 6H 2 ? 6LiH + 6Al + 9H 2 E21
.

The first reaction step, however, the decomposition of LiAlH4 is found to be exothermic with ΔHdecomposition = -10 kJ/(mol H2). Since the entropy of decomposition is positive. Rehydrogenation is not possible at all. The second reaction step, the decomposition of Li3AlH6 is endothermic with ΔHdecomposition = 25 kJ/(mol H2). The decomposition of LiH itself takes place at much higher temperatures with ΔH = 140 kJ/(mol H2) (Orimo et al., 2007). While the second reaction step, the decomposition of Li3AlH6 and rehydrogenation of LiH + Al shows rather suitable thermodynamic properties, sluggish kinetics prevent this system so far from being used for hydrogen storage.

To increase the storage capacity and tailor the reaction enthalpy of the NaAlH4 system it is a comprehensible approach to replace some of the Na by Li. Indeed Huot et al. (Huot et al., 1999) proved the existence of Na2LiAlH6 and the possible formation by high energy ball-milling of NaH + LiH + NaAlH4. Reversible hydrogen sorption is found to be possible in the Na-Li-Al-H system according to the following reaction:

2Na 2 LiAlH 6 4NaH + 2LiH + Al + 3H 2 E22

As in case of the pure Na-Al-H system and the Li-Al-H system kinetics can be improved by the addition of transition metal compounds like metal oxides, chlorides and fluorides, see (Ares Fernandez et al., 2007), (Ma et al., 2005) and (Martinez-Franco et al., 2010). However, due to the lack of any stable compound in the dehydrogenated state and the formation of a rather stable hydride the value of reaction enthalpy isn’t decreased but increased if compared to the original single Na and Li based aluminium hydrides. Fossdal et al. (Fossdal et al., 2005) has determined the pressure-composition isotherms of TiF3-doped Na2LiAlH6 in the temperature range of 170 ºC – 250 ºC. They determined the dissociation enthalpy and the corresponding entropy from the Van’t Hoff plot: |DH| = 56 kJ/(mol H2) and S = 138 J/(K mol H2). Therefore, instead of a lowering the heat of reaction the opposite is observed. The heat of reaction of the hexa-hydride phase is increased by about 10 kJ/(mol H2) if compared to the pure Na3AlH6 hydride phase.

In 2007 Yin et al. (Yin et al., 2007) presented DFT calculations about the doping effects of TiF3 on Na3AlH6. Their calculations suggested F- substitution for the H-anion leading to a reduction of the desorption enthalpy and therefore for a favourable thermodynamic modification of the Na3AlH6 system which was experimentally confirmed by Brinks et al. (Brinks et al., 2008) and Eigen et al. (Eigen et al., 2009).

Borohydrides

Only a very few hydrides show a higher gravimetric storage capacity than MgH2. For this they must be composed from very light elements. Knowing that Al-containing compounds can form reversible complex metal hydrides it is a reasonable approach to look for Boron-containing compounds as reversible hydrogen storage materials with even higher storage capacity. Borohydrides are known since 1940 when Schlesinger and Brown report about the successful synthesis of LiBH4 by reaction of LiEt and diborane (Schlesinger & Brown, 1940). Despite the early patent from Goerrig in 1958 (Goerrig, 1960) direct synthesis from gaseous H2 was not possible for long times. Until in 2004 three different groups from the USA (Vajo et al., 2005), South Korea (Cho et al., 2006) and Germany (Barkhordarian et al., 2007) independently discovered that by using MgB2 instead of pure Boron as starting material formation of the respective borohydrides occurs at rather moderate conditions of 5 MPa H2 pressure. Orimo et al. (Orimo et al., 2005) reports on the rehydrogenation of previously dehydrogenated LiBH4 at 35 MPa H2 pressure at 600 ºC. Mauron et al. (Mauron et al., 2008) report that rehydrogenation is also possible at 15 MPa. As in case of the Mg-based alloys and the aluminum hydrides the reaction enthalpy of many borohydrides is rather unsuitable for most applications. LiBH4 as one of the most investigated borohydrides with a very high gravimetric hydrogen density of 18.5 wt.% shows an endothermic desorption enthalpy of |DH| = 74 kJ/(mol H2) (Mauron et al., 2008) which is almost the same as in MgH2. Therefore the tailoring of the reaction enthalpy by substitution is a key issue for these materials as well. As in case of the aluminium hydrides there are two different possibilities for substitution in these complex hydrides: cation substitution and anion substitution. Nakamori et al. (Nakamori et al., 2006) reports about a linear relationship between the heat of formation ΔHboro of M(BH4)n determined by first principle methods and the Pauling electronegativity of the cation χp:

Δ H boro kJ ( molBH 4 ) 1 = 248 .7 χ P 390.8 E23

Aiming to confirm their theoretical results the same group performed hydrogen desorption experiments which show that the experimentally determined desorption temperature Td shows correlates with the Pauling electronegativity χp as well, see Fig. 11.

Figure 11.

The desorption temperature Td as a function of the Pauling electronegativity χP and estimated desorption enthalpies ΔHdes (Nakamori et al., 2007).

Based on these encouraging results several research groups started to investigate the partial substitution of one cation by another studying several bialkali metal borohydrides. The decomposition temperature of the bialkali metal borohydrides like LiK(BH4)2 is approximately the average of the decomposition temperature of the mono alkali borohydrides (Rude et al., 2011). Investigations of Li et al. (Li et al., 2007) and Seballos et al. (Seballos et al., 2009) confirmed that this correlation between desorption enthalpy / observed Td holds true for many double cation MM’(BH4)n systems, see Fig. 12.

Several experiments are indicating that transition metal fluorides are among the best additives for borohydrides (Bonatto Minella et al., 2011). While for some cases the function of the transition metal part as additive is understood (Bösenberg et al., 2009; Bösenberg et al., 2010; Deprez et al., 2010; Deprez et al., 2011), the function of F so far remained unclear. DFT calculations performed by Yin et al. (Yin et al., 2008) suggest a favourable modification of hydrogen reaction enthalpy in the LiBH4 system by substitution of the H--ion with the F--ion. However, no clear indicative experimental results for F- -substitution in borohydrides are found yet. In contrast to the F the heavier and larger halides Cl, Br, I are found to readily substitute in some borohydrides for the BH4 --ion and form solid solutions or stoichiometric compounds and are so far reported to stabilize the hydride phase leading to an increase of the desorption enthalpy |ΔH| (Rude et al., 2011).

Figure 12.

Decomposition temperatures, Tdec for metal borohydrides plotted as a function of the electronegativity of the metal, M’. (Rude et al., 2011)

3.2. Thermodynamic tuning using multicomponent systems: reactive additives and reactive hydride composites

In 1967 Reilly and Wiswall (Reilly & Wiswall, 1967) found another promising approach to tailor reaction enthalpies of hydrides (MHx) by mixing them with suitable reactants (A):

MH x + yA MA y + x 2 H 2 E24

They investigated the system MgH2/MgCu2 which reversibly reacts with hydrogen according to:

3 MgH 2 + MgCu 2 2Mg 2 Cu + 3H 2 E25

The formation of MgCu2 from Mg2Cu and Cu is exothermic and thus counteracts the endothermic release of hydrogen. Thereby, the total amount of hydrogen reaction enthalpy is reduced to roughly |ΔH| = 73 kJ/(mol H2) (Wiswall, 1978). The equilibrium temperature for 1 bar hydrogen pressure is reduced to about 240 ºC. In spite of the lower driving force for rehydrogenation, Mg2Cu is much more easily hydrogenated than pure Mg. A fact found in many other systems like the Reactive Hydride Composites as well.

Aluminum is another example of a reactive additive for MgH2. The reaction occurs via two steps (Bouaricha et al., 2000):

17 MgH 2 + 12Al 9MgH 2 + 4Mg 2 Al 3 + 8H 2 Mg 17 Al 12 + 17H 2 E26

The system can reversibly store 4.4 wt.% H2. Since the formation enthalpy ΔHForm of Mg17Al12 is -102 kJ/mol the total value of reaction enthalpy of reaction (26) is reduced by ~ 6 kJ/(mol H2) if compared to pure MgH2. An equilibrium pressure of 1 bar is reached at around 240 ºC again.

To further decrease the reaction enthalpy of a Mg-based system a much more stable compound would have to be formed during dehydrogenation. A system investigated by many groups is the MgH2-Si system. Mg2Si has an enthalpy of formation of ΔHForm = -79 kJ/mol. Due to the formation of Mg2Si the value of reaction enthalpy of MgH2/Si should therefore be reduced by 37 kJ/(mol H2) to about |ΔH| = 41 kJ/(mol H2) (Dornheim, 2010).

Theoretically 5 wt.% H2 can be stored via the reaction

2MgH 2 + Si Mg 2 Si + 4H 2 E27

The thermodynamic data indicate a very favourable equilibrium pressure of about 1 bar at 20 C and 50 bar at 120 ºC (Vajo, 2004). While so far rehydrogenation of Mg2Si was not shown to be possible the system LiH-Si turned out to be reversible. The enthalpy of dehydrogenation of LiH being 190 kJ/(mol H2) an equilibrium H2 pressure of 1 bar is reached at 910 C (Sangster, 2000; Dornheim, 2010). LiH reversibly reacts with Si via a two step reaction with the equilibrium pressure being more than 104 times higher and the dehydrogenation enthalpy being reduced by 70 kJ/(mol H2) (Vajo, 2004).

This approach has recently also been applied to borohydrides. According to Cho et al. (Cho et al., 2006) the decomposition temperature of pure LiBH4 is determined by CALPHAD to 1 bar H2 pressure at 403 C while the corresponding equilibrium temperature for the reaction

2LiBH 4 + Al 2LiH + AlB 2 + 3H 2 E28

is reduced to 188 ºC. Kang et al. (Kang et al., 2007) and Jin et al. (Jin et al., 2008) could show that this system indeed is reversible if suitable additives are used.

The only disadvantage of this approach is that the total reversible storage capacity per weight is reduced if something is added to the hydrogen storing material which contains no hydrogen.

The problem of reduced hydrogen capacity by using reactive additives has recently overcome by the approach of the Reactive Hydride Composites (Dornheim, 2006). Thereby, different high capacity hydrogen storage materials are combined which react exothermically with each other during decomposition, see Fig. 13.

One of the first examples of such a system is the LiNH2-LiH system which was discovered by Chen et al. (Chen et al., 2002):

L i N H 2 + 2 L i H L i 2 N H + L i H + H 2 L i 3 N + 2 H 2 E29

However, the value of reaction enthalpy is |ΔH| = 80 kJ/(mol H2) and therefore for most applications still much to high. In contrast the system

Mg ( NH 2 ) 2 + 2LiH Li 2 Mg ( NH ) 2 + 2H 2 E30

shows a much more suitable desorption enthalpy of |ΔH|~40 kJ/(mol H2) with an expected equilibrium pressure of 1 bar at approximately 90 ºC (Xiong et al., 2005; Dornheim, 2010).

Figure 13.

Schematic of the reaction mechanism in Reactive Hydride Composite.

In 2004 Vajo et al. (Vajo et al., 2005), Cho et al. (Cho et al., 2006) and Barkhordarian et al. (Barkordarian et al., 2007) independently discovered that the usage of borides especially MgB2 as a starting material facilitates the formation of different borohydrides. This finding initiated the development and investigation of several new reversible systems with high storage capacities of 8 – 12 wt.% H2 and improved thermodynamic and kinetic properties such as 2 LiBH4+MgH2 (Bösenberg et al., 2009; 2010; 2010b), 2 NaBH4+MgH2 (Garroni et al., 2010; Pistidda et al., 2010; 2011; Pottmaier et al., 2011), Ca(BH4)2+MgH2 (Barkhordarian et al., 2008), 6 LiBH4+CeH2, 6 LiBH4+CaH2 (Jin et al., 2008b), LiBH4/Ca(BH4)2 (Lee et al., 2009).

One of the most intensely studied systems hereof is the 2 LiBH4 + MgH2 system. The indended reaction pathway is:

2 L i B H 4 + M g H 2 2 L i H + M g B 2 + 4 H 2 E31

However, several other reaction pathways are possible leading to products such as LiB2, amorphous B, Li2B12H12 or Li2B10H10. Bösenberg et al. (Bösenberg et al., 2010b) could show that due to a higher thermodynamic driving force for the favoured reaction the competing reactions can be suppressed by applying a hydrogen back pressure and limiting the dehydrogenation temperature. Nevertheless, since long-range diffusion of metal atoms containing species is required, see Fig. 13, in bulk ball-milled samples dehydrogenation so far occurs only at temperatures higher than 350 C, hydrogenation at temperatures higher than 250 C.

The dehydrogenation temperatures of this Reactive Hydride Composite, however, can be significantly reduced by using nanoconfined 2 LiBH4 + MgH2 stabilised in inert nanoporous aerogel scaffold materials whereby long-range phase separation is hindered and thus the diffusion path length reduced (Gosalawit-Utke, 2011).

Advertisement

4. Conclusion

Metal hydrides offer a safe and compact alternative for hydrogen storage. The thermodynamic properties of them determine both their reaction heat as well as hydrogen equilibrium pressure at given temperature and, therefore, are important parameters to be taken into account. Optimised system integration for a given application is not possible without selecting a hydride with suitable thermodynamic properties. To achieve highest possible energy efficiencies the heat of reaction and temperature of operation of the metal hydride should be adapted to the waste heat and temperature of operation of the fuel cell / fuel combustion system. It has been found that the thermodynamic properties of metal hydrides can be tailored in a wide range. Unfortunately, so far all the known conventional metal hydrides with more or less ideal reaction enthalpies and hydrogen equilibrium pressures above 5 bar at room temperature suffer from a rather limited reversible hydrogen storage capacity of less than 2.5 wt.%. With such a material it is not possible to realise a solid storage hydrogen tank with a total hydrogen storage density of more than 1.8 wt.% H2. Such tank systems still have advantages for the storage of small quantities of hydrogen for larger quantities, however, modern high pressure composite tank shells have a clear advantage in respect of gravimetric storage density. To realise a solid storage tank for hydrogen with a comparable gravimetric storage density it is required that novel hydrogen storage materials based on light weight elements are developed. There are several promising systems with high gravimetric storage densities in the range of 8 – 12 wt.% H2. For the applications of these novel material systems it is important to further adapt thermodynamic properties as well as the temperatures of operation towards the practical requirements of the system.

The discovery of the approach of combining different hydrides which react with each other during hydrogen release by forming a stable compound, the so-called Reactive Hydride Composites, show a great promise for the development of novel suitable hydrogen storage material systems with elevated gravimetric storage densities. However, so far, the ideal storage material with low reaction temperatures, a reaction heat in the range of |ΔH| = 20-30 kJ/(mol H2) and a on-board reversible hydrogen storage density of more than 6 wt.% H2 has not been found.

References

  1. 1. Ares Fernandez. J. R. Aguey-Zinsou F. Elsaesser M. Ma X. Z. Dornheim M. Klassen T. Bormann R. 2007 Mechanical and thermal decomposition of LiAlH4 with metal halides. International Journal of Hydrogen Energy, 32 8 1033 1040 ), 0360-3199
  2. 2. Barkhordarian G. Klassen T. Bormann R. 2006 Kinetic investigation of the effect of milling time on the hydrogen sortpion reaction of magnesium catalyzed with different Nb2O5 contents. Journal of Alloys and Compounds, 407 1-2 , 249 255 ), 0925-8388
  3. 3. Barkhordarian G. Klassen T. Dornheim M. Bormann R. 2007 Unexpected kinetic effect of MgB2 in reactive hydride composites containing complex borohydrides. Journal of Alloys and Compounds, 440 1-2 , L18 L21 ), 0925-8388
  4. 4. Barkhordarian G. Jensen T. R. Doppiu S. Bösenberg U. Borgschulte A. . Gremaud R. Cerenius Y. Dornheim M. Klassen T. Bormann R. 2008 Formation of Ca(BH4)2 from Hydrogenation of CaH2+MgB2 Composite. Journal of Physical Chemistry C, 112 7 2743 2749 ), 1932-7447
  5. 5. Bösenberg U. Vainio U. Pranzas P. K. Bellosta von. Colbe J. M. Goerigk G. Welter E. Dornheim M. Schreyer A. Bormann R. 2009 On the chemical state and distribution of Zr- and V-based additives in Reactive Hydride Composites. Nanotechnology, 20 20 204003 1-204003/9) 1361-6528
  6. 6. Bösenberg U. Kim J. W. Gosslar D. Eigen N. Jensen T. R. Bellosta von. Colbe J. M. Zhou Y. Dahms M. Kim D. H. Guenther R. cho Y. W. Oh K. H. Klassen T. Bormann R. Dornheim M. 2010 Role of additives in LiBH4-MgH2 Reactive Hydride Composites for sorption kinetics. Acta Materialiea, 58 9 3381 3389 ), 1359-6454
  7. 7. Bösenberg U. Ravnsbaek D. B. Hagemann H. D’Anna V. Bonatto Minella. C. Pistidda C. van Beek W. Jensen T. R. Bormann R. Dornheim M. 2010b Pressure and Temperature Influence on the Desorption Pathway of the LiBH4-MgH2 Composite System. Journal of Physical Chemistry C, 114 35 15212 15217 )
  8. 8. Bogdanovic B. Schwickardi M. 1997 Ti-doped alkali metal aluminum hydrides as potential novel reversible hydrogen storage materials. Journal of Alloys and Compounds, 253-254 , 1 9 ), 0925-8388
  9. 9. Bogdanovic B. Reiser A. Schlichte K. Spliethoff B. Tesche B. 2002 Thermodynamics and dynamics of the Mg-Fe-H system and its potential for thermochemical thermal energy storage. Journal of Alloys and Compounds, 345 1-2 , 77 89 ), 0925-8388
  10. 10. Bogdanovic B. Felderhoff M. Streukens G. 2009 Hydrogen storage in complex metal hydrides. Journal of the Serbian Chemical Society, 74 2 183 196 ), 0352-5139
  11. 11. Bonatto Minella. Christian Garroni Sbastiano. Pistidda Claudio. Gosalawit-Utke R. Barkhordarian G. Rongeat C. Lindeman I. Gutfleisch O. Jensen T. R. Cerenius Y. Christnsen J. Baro M. D. Bormann R. Klassen T. Dornheim M. 2011 Effect of Transition Metal Fluorides on the Sorption Properties and Reversible Formation of Ca(BH4)2. Journal of Physical Chemistry C, 115 5 2497 2504 ), 1932-7447
  12. 12. Bouaricha S. Dodelet J. P. Guay D. Huot J. Boily S. Schulz R. 2000 Hydriding behaviour of Mg-Al and leached Mg-Al compounds prepared by high-energy ball-milling. Journal of Alloys and Compounds, 297 282 293 )
  13. 13. Brinks H. Fossdal A. Hauback B. 2008 Adjustment of the stability of complex hydrides by anion substitution. Journal of Physical Chemistry C, 112 14 5658 5661 ), 1932-7447
  14. 14. Buchner H. 1982 Energiespeicherung in Metallhydriden, Springer-Verlag, 3-211-81703-4, Wien
  15. 15. Buchner H. Povel R. 1982 The Daimler-Benz Hydride Vehicle Project. International Journal of Hydrogen Energy, 7 3 259 266 ), 0360-3199/82/030259-08
  16. 16. Chen P. Xiong Z. T. Luo J. Z. Lin J. Tan K. L. 2002 Interaction of hydrogen with metal nitrides and imides. Nature, 420 302 304 )
  17. 17. Cho Y. W. Shim J. H. Lee B. J. 2006 Thermal destabilization of binary and complex metal hydrides by chemical reaction: A thermodynamic analysis. CALPHAD: Computer Coupling of Phase Diagrams and Thermochemistry, 30 1 65 69 ), 0364-5916
  18. 18. Darnaudry J. P. Darriet B. Pezat M. 1983 The Mg2Ni0.75M0.25 alloys (M = 3d element): their application to hydrogen storage. International Journal of Hydrogen Energy, 8 705 708 )
  19. 19. Deprez E. Justo A. Rojas T. C. Lopez Cartes. C. Bonatto Minella. C. Bösenberg U. Dornheim M. Bormann R. Fernandez A. 2010 Microstructural study of the LiBH4-MgH2 Reactive Hydride Composite with and without Ti isopropoxide additive. Acta Materialia, 58 17 5683 5694 ), 1359-6454
  20. 20. Deprez E. Munoz-Marquez M. A. Jimenez de Haro. M. C. Palomares F. J. Foria F. Dornheim M. Bormann R. Fernandez A. 2011 Combined x-ray photoelectron spectroscopy and scanning electron microscopy studies of the LiBH4-MgH2 Reactive Hydride Composite with and without a Ti-based additive. Journal of Applied Physics, 109 1 014913 1-014913/10), 0021-8979
  21. 21. Didisheim J. J. Zolliker P. Yvon K. Fischer P. Schefer J. Gubelmann M. Williams A. F. 1984 Dimagnesium iron(II) hydride; Mg2FeH6, containing octahedral FeH64- anions. Inorganic Chemistry, 23 13 1953 1957 ), 0020-1669
  22. 22. Dornheim M. Eigen N. Barkhordarian G. Klassen T. Bormann R. 2006 Tailoring Hydrogen Storage Materials Towards Application. Advanced Engineering Materials, 8 5 377 385 ), 1438-1656
  23. 23. Dornheim M. Doppiu S. Barkhordarian G. Boesenberg U. Klassen T. Gutfleisch O. Bormann R. 2007 Hydrogen storage in magnesium-based hydrides and hydride composites. Viewpoint paper in: Scripta Materialia, 56 841 846 ), 1359-6462
  24. 24. Dornheim M. Klassen T. 2009 High Temperature Hydrides, In: Encyclopedia of Electrochemical Power Sources, 3 J. Garche, C. Dyer, P. Moseley, Z. Ogumi, D. Rand, B. Scrosati, 459 472 ), Elsevier, 100444520937
  25. 25. Dornheim M. 2010 Tailoring Reaction Enthalpies of Hydrides, In: Handbook of Hydrogen Storage, Michael Hirscher, 187 214 ), Wiley-VCH Verlag GmbH & Co, 978-3-52732-273-2 Weinheim
  26. 26. Eigen N. Keller C. Dornheim M. Klassen T. Bormann R. 2007 Industrial production of light metal hydrides for hydrogen storage. Viewpoint Set in Scripta Materialia, 56 10 847 851 ), 1359-6462
  27. 27. Eigen N. Gosch F. Dornheim M. Klassen T. Bormann R. 2008 Improved hydrogen sorption of sodium alanate by optimized processing. Journal of Alloys and Compounds, 465 1-2 , 310 316 ), 0925-8388
  28. 28. Eigen N. Bösenberg U. Bellosta von. Colbe J. Jensen T. R. Cerenius Y. Dornheim M. Klassen T. Bormann R. 2009 Reversible hydrogen storage in NaF-Al composites. Journal of Alloys and Compounds. 477 1-2 , 76 80 ), 0925-8388
  29. 29. Fossdal A. Brinks H. W. Fonneloep J. E. Hauback B. C. 2005 Pressure-composition isotherms and thermodynamic properties of TiF3-enhanced Na2LiAlH6. Journal of Alloys and Compounds, 397 1-2 , 135 139 ), 0925-8388
  30. 30. Fujitani S. Yonezu I. Saito T. Furukawa N. Akiba E. Hayakawa H. Ono S. 1991 Relation between equilibrium hydrogen pressure and lattice parameters in pseudobinary Zr-Mn alloy systems. Journal of the Less Common Metals, 172-174 , 1 220 230 )
  31. 31. Fukai Y. 1993 The Metal-Hydrogen System. Springer Series in Materials Science, 21 Springer, Berlin
  32. 32. Garroni S. Milanese C. Girella A. Marini A. Mulas G. Menendez E. Pistidda C. Dornheim M. Surinach S. Baro M. D. 2010 Sorption properties of NaBH4/MH2 (M = Mg, Ti) powder systems. International Journal of Hydrogen Energy, 35 11 5434 5441 )
  33. 33. Goerrig D. 1960 Borohydrides of alkali and alkaline earth metals. German Patent 1077644, F27373 IVa/12i, Application No. DE 1958 -F27373
  34. 34. Gosalawit-Utke R. Nielsen T. K. Saldan I. Laipple D. Cerenius Y. Jensen T. R. Klassen T. Dornheim M. 2011 Nanoconfined 2LiBH4-MgH2 prepared by direct melt infiltration into nanoporous materials. Journal of Physical Chemistry C, 115 21 10903 10910 )
  35. 35. Griessen R. Driessen A. 1984 Heat of formation and band structure of binary and ternary metal hydrides. Phys. Rev. B, 30 8 4372 4381 ), 0163-1829
  36. 36. Hägg G. 1931 Röntgen investigations on the hydrides of titanium zirconium, vanadium and tantalum. Zeitschrift für physikalische Chemie B, 11 433 445 )
  37. 37. Hanada N. Ichikawa T. Fujii H. 2007 Hydrogen absorption kinetics of the catalyzed MgH2 by niobium oxide. Journal of Alloys and Compounds, 446-447 , 67 71 ), 0925-8388
  38. 38. Huot J. Boily S. Güther V. Schulz R. 1999 Synthesis of Na3AlH6 and Na2LiAlH6 by mechanical alloying. Journal of Alloys and Compounds, 283 1-2 , 304 306 ), 0925-8388
  39. 39. Jin S. A. Shim J. H. Cho Y. W. Yi K. W. Zabara O. Fichtner M. 2008 Reversible hydrogen storage in LiBH4-Al-LiH composite powder. Scripta Materialia, 58 11 963 965 ), 1359-6462
  40. 40. Jin S. A. Lee Y. S. Shim J. H. Cho Y. W. 2008b Reversible Hydrogen Storage in LiBH4-MH2 (M = Ce, Ca) Composites. Journal of Physical Chemistry C, 112 25 9520 9524 ), 1932-7447
  41. 41. Kang X. D. Wang P. Ma L. P. Cheng H. M. 2007 Reversible hydrogen storage in LiBH4 destabilized by milling with Al. Applied Physics A: Materials Science & Processing, 89 4 963 966 )
  42. 42. Klassen T. Oelerich W. Zeng K. Bormann R. Huot J. 1998 Nanocrystalline Mg-based alloys for hydrogen storage, In: Magnesium Alloys and their Applications, B.L. Mordike and K. U. Kainer, 308 311 ), Werkstoff-Informationsgesellschaft mbH Frankfurt, CODEN: 68TSA9, Wolfsburg, Germany
  43. 43. Klell M. 2010 Storage of Hydrogen in the Pure Form, In: Handbook of Hydrogen Storage, Michael Hirscher, 187 214 ), Wiley-VCH Verlag GmbH & Co, 978-3-52732-273-2 Weinheim
  44. 44. Konstanchuk, I.G.; Ivanov, E.Y.; Pezat, M.; Darriet, B.; Bodyrev, V.V.; Hagenmüller, P. (1987). The hydrideng properties of a mechanical alloy with composition Mg-25% Fe. J. Less-Common Met., Vol. 131, pp. (181-189)
  45. 45. Lee J. Y. Ravnsbaek D. Lee Y. S. Kim Y. Cerenius Y. Shim J. H. Jensen T. R. Hur N. H. Cho Y. W. 2009 2009). Decomposition Reactions and Reversibility of the LiBH4-Ca(BH4)2 Composite. Journal of Physical Chemistry C, 113 33 15080 15086 ), 1932-7447
  46. 46. Libowitz G. G. Hayes H. F. Gibb T. R. G. Jr 1958 The system zirconium-nickel and hydrogen. Journal of Physical Chemistry, 62 76 79 ), 0022-3654
  47. 47. Li H. W. Orimo S. Nakamori Y. Miwa K. Ohba N. Towata S. Züttel A. 2007 Materials designing of metal borohydrides: Viewpoints from thermodynamical stabilities. Journal of Alloys and Compounds, 446-447 , 315 318 )
  48. 48. Li H. W. Kikuchi K. Nakamori Y. Ohba N. Miwa K. Towata S. Orimo S. 2008 Dehydriding and Rehydriding Processes of Well-Crystallized Mg(BH4)2 Accompanying with Formation of Intermediate Compounds. Acta Materialia, 56 1342 1347 )
  49. 49. Ma X. Z. Martinez-Franco E. Dornheim M. Klassen T. Bormann R. 2005 Catalyzed Na2LiAlH6 for hydrogen storage. Journal of Alloys and Compounds, 404-406 , 771 774 ), 0925-8388
  50. 50. Martinez-Franco E. Klassen T. Dornheim M. Bormann R. Jaramillo-Vigueras D. 2010 Hydrogen sorption properties of Ti-oxide/chloride catalyzed Na2LiAlH6. Ceramic Transactions, 209 13-20 , 13 20 ), 1042-1122
  51. 51. Mauron P. Buchter F. Friedrichs O. Remhof A. Bielmann M. Zwichy C. N. Züttel A. 2008 Stability and Reversibility of LiBH4. J. Phys. Chem. B, 112 906 910 )
  52. 52. Nakamori Y. Miwa K. Ninomiya A. Li H. Ohba N. Towata S. Züttel A. Orimo S. 2006 Correlation between thermodynamical stabilities of metal borohydrides and cation electronegativites: First-principles calculations and experiments. Physical Review B, 74 045126 045121 -9)
  53. 53. Nakamori Y. Miwa K. Li H. Ohba N. Towata S. Orimo S. 2007 Tailoring of Metal Borohydrides for Hydrogen Storage Applications. Mater. Res. Soc. Symp. Proc. 971 0971-Z02-01
  54. 54. Oelerich W. 2000 Sorptionseigenschaften von nanokristallinen Metallhydriden für die Wasserstoffspeicherung, Helmholtz-Zentrum Geesthacht Zentrum für Material- und Küstenforschung GmbH (formerly GKSS-Forschungszentrum Geesthacht GmbH), 0344-9626 Geesthacht, Germany
  55. 55. Oelerich W. 2001 Metal oxides as catalysts for improved hydrogen sorption in nanocrystalline Mg-based materials. Journal of Alloys and Compounds, 315 1-2 , 237 242 ), 0925-8388
  56. 56. Orimo S. Ikeda K. Fujii H. Saruki S. Fukunaga T. Züttel A. Schlapbach L. 1998 Structural and hydriding properties of (Mg1-xAlx)Ni-H(D) with amorphous or CsCl-type cubic structure (x=0-0.5), Acta Materialia, 46 13 4519 4525 ), 1359-6454
  57. 57. Orimo S. Nakamori Y. Kitahara G. Miwa K. Ohba N. Towata S. Zuettel A. 2005 Dehydriding and rehydriding reactions of LiBH4. Journal of Alloys and Compounds, 404-430 , 0925 8388 ), 0925-8388
  58. 58. Orimo S. Nakamori Y. Eliseo J. R. Züttel A. Jensen C. M. 2007 Complex Hydrides for Hydrogen Storage. Chemical Reviews, 107 10 4111 4132 ). 0009-2665
  59. 59. Panella B. Hirscher M. Roth S. 2005 Hydrogen adsorption in different carbon nanostructures. Carbon, 43 2209 2214 )
  60. 60. Pistidda C. Garroni S. Bonatto Minella. C. Coci F. Jensen T. R. Nolis P. Bösenberg U. Cerenius Y. Lohstroh W. Fichtner M. Baro M. D. Bormann R. Dornheim M. 2010 Pressure Effect on the 2NaH+MgB2 Hydrogen Absorption Reaction. Journal of Physical Chemistry C, 114 49 21816 21823 )
  61. 61. Pistidda C. Barkhordarian G. Rzeszutek A. Garroni S. Minella C. Bonatto Baro M. D. Nolis P. Bormann Ruediger. Klassen T. Dornheim M. 2011 Activation of the reactive hydride composite 2NaBH4 + MgH2. Scripta Materialia, 64 11 1035 1038 )
  62. 62. Pottmaier D. Pistidda C. Groppo E. Bordiga S. Spoto G. Dornheim M. Baricco M. 2011 Dehydrogenation reactions of 2NaBH4 + MgH2 system. International Journal of Hydrogen Energy. 36 13 7891 7896 )
  63. 63. Pourarian F. Shinha V. K. Wallace W. E. Smith H. K. 1982 Kinetics and thermodynamics of ZrMn2-based hydrides. Journal of the Less Common Metals, 88 2 451 458 )
  64. 64. Pranzas P. K. Dornheim M. Bellmann D. Aguey-Zinsou K. F. Klassen T. Schreyer A. 2006 SANS/USANS investigations of nanocrystalline MgH2 for reversible storage of hydrogen. Physica B: Condensed Matter, 385-386 , 1 630 632 ), 0921-4526
  65. 65. Pranzas P. K. Dornheim M. Boesenberg U. Ares Fernandez. J. R. Goerigk G. Roth S. V. Gehrke R. Schreyer A. 2007 Small-angle scattering investigations of magnesium hydride used as a hydrogen storage material. Journal of Applied Crystallography, 40 S1 383 387 ), 0021-8898
  66. 66. Puszkiel J. A. Larochette P. A. Gennari F. C. 2008 Thermodynamic-kinetic characterization of the synthesized Mg2FeH6-MgH2 hydrides mixture. International Journal of Hydrogen Energy, 33 13 3555 3560 ), 0360-3199
  67. 67. Reilly J. J. Wiswall R. H. 1967 Reaction of hydrogen with alloys of magnesium and copper. Inorganic Chemistry, 6 2220 2223 ), 0020-1669
  68. 68. Reilly J. J. Wiswall R. H. 1968 Reaction of hydrogen with alloys of magnesium and nickel and the formation of Mg2NiH4. Inorganic Chemistry, 7 2254 2256 ), 0020-1669
  69. 69. Reilly J. J. Sandrock G. D. 1980 Hydrogen storage in metal hydrides. Scientific American, 242 2 5118 5129 ), 0036-8733
  70. 70. Rude L. H. Nielsen T. K. Ravnsbaek D. B. Bösenberg U. Ley M. B. Richter B. Arnbjerg L. M. Dornheim M. Filinchuk Y. Besenbacher F. Jensen T. R. 2011 Tailoring properties of borohydrides for hydrogen storage: A review. Phys. Status Solidi A, DOI: 10.1002/pssa.201001214
  71. 71. Sandrock G. (1999 1999).A panoramic overview of hydrogen storage alloys from a gas reaction point of view. Journal of Alloys and Compounds, 293-295 , 877 888 )
  72. 72. Sangster J. J. Pelteon A. D. 2000 In: Phase Diagrams of Binary Hydrogen Alloys; Manchester, F.D. ; ASM International: Materials Park, OH, 74
  73. 73. Schlesinger H. I. Brown H. C. 1940 Metallo borohydrides. III. Lithium borohydride. Journal of the American Chemical Society, 62 3429 3435 ), 0002-7863
  74. 74. Schmitz B. Mueller U. Trukhan N. Schubert M. Ferey G. Hirscher M. 2008 Heat of adsorption in microporous high-surface-area materials. Chem Phys Chem, 9 2181 2184 )
  75. 75. Schober T. Wenzl H. 1978 The systems niobium hydride (deuteride), tantalum hydride (deuteride), vanadium hydride (deuteride): structures, phase diagrams, morphologies, methods of preparation, In: Topics in Applied Physics 29 Hydrogen in Metals, Vol. 2, G. Alefeld and J. Völkl, 11 71 ), Springer, 0303-4216 Berlin
  76. 76. Schüth F. Bogdanovic B. Felderhoff M. 2004 Light metal hydrides and complex hydrides for hydrogen storage. Chemical Communications, 20 2249 2258 ), 1359-7345
  77. 77. Seballos L. Zhang J. Z. Rönnebro E. Herberg J. L. Majzoub E. H. 2009 Metastability and crystal structure of the bialkali complex metal borohydrides NaK(BH4)2. Journal of Alloys and Compounds, 476 446 450 )
  78. 78. Selvam P. Yvon K. 1991 Synthesis of magnesium iron hydride (Mg2FeH6), magnesium cobalt hydride (Mg2CoH5) and magnesium nickel hydride (Mg2NiH4) by high-pressure sintering of the elements, International Journal of Hydrogen Energy, 16 9 615 617 ), 0360-3199
  79. 79. Terashita N. Kobayashi K. Sasai T. Akiba E. 2001 Structural and hydriding properties of (Mg1-xCax)Ni2 Laves phase alloys. Journal of Alloys and Compounds, 327 1-2 , 275 280 ), 0925-8388
  80. 80. Tsushio Y. Enoki H. Akiba E. 1998 Hydrogenation properties of MgNi0.86M10.03 (M1=Cr, Fe, Co, Mn) alloys. Journal of Alloys and Compounds, 281 301 305 ).
  81. 81. Vajo J. J. Mertens F. Ahn C. C. Bowman R. C. Jr Fultz B. 2004 Altering Hydrogen Storage Properties by Hydride Destabilization through Alloy Formation: LiH and MgH2 Destabilized with Si. Journal of Physical Chemistry B, 108 37 13977 13983 ), 1520-6106
  82. 82. Vajo J. J. Skeith S. L. Mertens F. 2005 Reversible Storage of Hydrogen in Destabilized LiBH4. Journal of Physical Chemistry B, 109 9 3719 3722 ), 1520-6106
  83. 83. Van Mal H. H. Buschow K. H. H. Miedema A. R. 1974 Hydrogen absorption in lanthanium-nickel (LaNi5) and related compounds. Experimental observations and their explanation. Jounal of the Less-Common Metals, 35 1 65 76 ), 0022-5088
  84. 84. Wiswall R. 1978 Hydrogen Storage in Metals. In: Topics of Appl. Phys., 29-Hydrogen in Metals II, G. Alefeld and J. Völkl, 201 242 ), Springer-Verlag, ISS: 0303-4216, Berlin, Heidelberg, New York
  85. 85. Xiong Z. T. Hu J. J. Wu G. T. Chen P. Luo W. Gross K. Wang J. 2005 Thermodynamic and kinetic investigations of the hydrogen storage in the Li-Mg-N-H system. Journal of Alloys and Compounds, 398 1-2 , 235 239 ), 0000-0925-8388
  86. 86. Yin L. C. Wang P. Kang X. D. Sun C. H. Cheng H. M. 2007 Functional anion concept: Effect of fluorine anion on hydrogen storage of sodium alanate. Physical Chemistry Chemical Physics, 9 12 1499 1502 ), 1463-9076
  87. 87. Yin L. Wang P. Fang Z. Cheng H. 2008 Thermodynamically tuning LiBH4 by fluorine anion doping for hydrogen storage: A density functional study. Chemical Physics Letters, 450 4-6 , 318 321 ), 0009-2614
  88. 88. Yoshida M. . Akiba E. 1995 Hydrogen absorbing-desorbing properties and crystal structure of the Zr-Ti-Ni-Mn-V AB2 Laves phase alloys. Journal of Alloys and Compds., 224 121 126 )
  89. 89. Züttel A. 2003 Materials for Hydrogen Storage. Materials Today, 6 24 33 ), 1369-7021

Written By

Martin Dornheim

Submitted: 15 November 2010 Published: 02 November 2011