Você está na página 1de 5

Published on Web Date: June 11, 2010

r2010 American Chemical Society 1982 DOI: 10.1021/jz100386j


|
J. Phys. Chem. Lett. 2010, 1, 19821986
pubs.acs.org/JPCL
Tuning the Hydrogen Storage in Magnesium Alloys
S uleyman Er,

Gilles A. de Wijs,

and Geert Brocks*


,

Computational Materials Science, Faculty of Science and Technology and MESA Institute for Nanotechnology, University of
Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands, and

Electronic Structure of Materials, Institute for Molecules and
Materials, Faculty of Science, Radboud University Nijmegen, Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands
ABSTRACT Mixing Mg with Ti leads to an alloy with markedly faster hydrogena-
tion kinetics as compared to that of pure Mg, but the resulting hydrides are too
stable for hydrogen storage applications. Moreover, Mg-Ti alloys are thermody-
namically unstable with respect to decomposition into the elements. In this Letter,
we show by means of first-principles density functional theory calculations that
adding Al or Si to Mg-Ti makes the alloys more stable and their hydrides less stable.
Controlling the structure of Mg-Ti-X (X = Al or Si) alloys by growing multilayers
of Mg and TiXallows for tuning the hydrogenation energy and improves the alloys
as hydrogen storage materials.
SECTION Energy Conversion and Storage
M
agnesium, aninexpensive andabundant metal, can
store up to 7.7 weight% hydrogen. The application
of MgH
2
as a hydrogen storage material is however
obstructed by its high thermodynamic stability and very low
(de)hydrogenation rates, which lead to excessively high oper-
ating temperatures T J 300 C.
1
A large amount of research
has been devoted to reducing the stability of MgH
2
and
enhancing its (de)hydrogenation kinetics. Experiments on
thin films have shown that alloying magnesium with early
transition-metal (TM) elements can markedly improve the
(de)hydrogenation kinetics.
2
The improved kinetics has been
correlated with a cubic, fluorite-type Mg
(1-x)
TM
x
H
2
phase,
whereas the tetragonal rutile phase of MgH
2
would hamper
the (de)hydrogenation kinetics.
3,4
From first-principles cal-
culations, we have shown that a fluorite-type phase of
Mg
(1-x)
TM
x
H
2
is stable with respect to a rutile-type phase
for TM = Sc, Ti, V, and Cr, and x J 0.2.
5,6
The dehydrogenation enthalpies of Mg
(1-x)
TM
x
H
2
are how-
ever either larger or at least not significantly smaller than that of
MgH
2
, implicating that the thermodynamic properties are not
markedly improved. In addition, the stability of the Mg
(1-x)
TM
x
alloys decreases along the series Sc, Ti, V, Cr. In fact, only
Mg
(1-x)
Sc
x
is thermodynamically stable with respect to decom-
position into the elemental metals (Mg and Sc). Scandium is a
rare element, however, which prevents its large-scale use in a
hydrogen storage material. Mg
(1-x)
V
x
and Mg
(1-x)
Cr
x
are ther-
modynamically quite unstable, hampering their use in a rever-
sible system, which needs to undergo repeated (de)hydroge-
nationcycles. Mg
(1-x)
Ti
x
is aninterestingcase. The binaryphase
diagram of Mg and Ti indicates that a thermodynamically
stable alloy between these two elements cannot be formed.
Studies of kinetically stabilized bulk Mg-Ti samples have been
inconclusive,
7,8
but metastable Mg-Ti alloys are readily formed
in thin films by codepositing the elements. Such films have a
typical thickness of 200 nmand survive repeated (de)hydroge-
nation cycles at room temperature.
4,9
The high stability of Mg-Ti hydride remains a problem,
however. The calculated dehydrogenation enthalpy of
Mg
0.75
Ti
0.25
H
2
is 0.76 eV/H
2
, which is close to the values
reported for MgH
2
.
5,6
Recently, Vermeulen et al. have sug-
gested that incorporating Al or Si into Mg-Ti thin films
improves their hydrogen storage properties, whereas their
hydrogen storage capacity does not deteriorate.
10
In this
paper, we study the energetics of Mg-Ti-Al,Si alloys and
their hydrides for several possible compositions by means of
first-principles density functional theory (DFT) calculations.
Simultaneous deposition of the metals in thin films very likely
results indisorderedstructures. We showthat nanostructuring
the alloys as atomic multilayers enables one to significantly
improve their hydrogen storage properties.
Starting from the binary alloys, the phase diagrams of
Ti with Al or Si show that, for example, alloys with a 1:1
composition are very stable.
11,12
Both TiAl and TiSi crystallize
in a tetragonally distorted fcc structure, the so-called L1
0
structure. The calculated formation energies per atom (E
f
) of
TiAl and TiSi are -0.409 and -0.495 eV/atom, respectively,
with respect to the elements in their ground-state metal
structures. In contrast, MgTi alloys are particularly unstable,
andsofar, nocompositions or structures have beenfoundthat
give a negative formation energy. As an example, the calcu-
lated E
f
of MgTi in the L1
0
structure is 0.206 eV/atom.
Adding Al or Si to form ternary Mg-Ti-Al,Si compounds
could improve E
f
via the formation of stable Ti-Al or Ti-Si
bonds. In addition, the formation energies of MgAl and MgSi
(in the L1
0
structure) are -0.015 and 0.099 eV/atom, respec-
tively, which shows that Mg-Al and Mg-Si bonds are much
more advantageous than Mg-Ti bonds. So already from
arguments based upon a simple structure, one may suspect
Received Date: March 24, 2010
Accepted Date: June 8, 2010
r2010 American Chemical Society 1983 DOI: 10.1021/jz100386j |J. Phys. Chem. Lett. 2010, 1, 19821986
pubs.acs.org/JPCL
that adding Al or Si to Mg-Ti may lead to a stabilized alloy not
only by forming stable Ti-Al or Ti-Si bonds but also by
preventing unstable Mg-Ti bonds.
13
Mg and Si can formthe very stable compound Mg
2
Si with a
fluorite structure. If one adds Si to Mg-Ti, the formation of
Mg
2
Si needs to be prevented since that compound does not
absorb hydrogen.
14
If we consider the reaction Mg
2
Si Ti T
TiSi 2Mg, thenour calculations showthat the materials onthe
right-hand side of this reaction are more stable by 0.35 eV per
TiSi unit. This leaves room for designing a Mg-Ti-Si com-
poundsuchthat Ti captures theSi andprevents theformationof
Mg
2
Si. Thesamereasoningholds for MgandAl, whichcanform
stable alloys, albeit with a relatively small formation energy.
15
The large formation energy of TiAl then leaves room for pre-
venting these Mg-Al alloys from being formed.
Considering the stability of TiAl and TiSi, it makes sense to
look at ternary Mg-Ti-X, X =Al,Si, alloys whose Ti-X ratio
is fixed at 1:1, that is, Mg
(1-2y)
Ti
y
X
y
. We study the composi-
tions y = 0.125, 0.1875, and 0.25, which cover the experi-
mentally reported range of compositions that show
interesting hydrogenstorage properties.
10
One could envision
growing multilayers of TiX and Mg by atomic beam epitaxy,
for instance. Figure 1 shows examples of such multilayers
grown in the (001) direction of a fcc structure. Alternating one
layer of TiX and one layer of Mg leads to the composition
Mg
0.5
Ti
0.25
X
0.25
. Mg
0.75
Ti
0.125
X
0.125
can then be formed by
alternating one layer of TiX with three layers of Mg and
Mg
0.625
Ti
0.1875
X
0.1875
by a more complex stack of one TiX
layer and one or two Mg layers. We optimize the unit cells
without imposing any constraints. Calculations on Mg-Ti
alloys have shown that the difference between fcc and hcp
structures is only 0.04 eV/atom.
5,6
Starting from hcp struc-
tures would alter the formation energies somewhat, but it
does not change the trend.
The formation energies per atom of these Mg
(1-2y)
Ti
y
X
y
structures are given in Table 1. They can be compared to the
formation energies of layered Mg
(1-2y)
Ti
2y
compounds with
the same amount of Mg. Adding Al or Si indeed increases the
stability. Mg-Ti is quite unstable over the whole composition
range with respect to phase separation into the elements. The
formation energy of these layered Mg-Ti structures scales
linearly with the Ti content, which one may attribute to the
formation of Mg-Ti interfaces. A linear fit with the Ti content
gives aninterface energyof 0.20eVper interface andTi atom.
The increased stability of Mg
(1-2y)
Ti
y
X
y
can be attributed to
the stability of TiX layers. The effect decreases with increased
Mg content, which can be interpreted as a strain effect. The
lattice constants of TiAl and TiSi are 11 and 14%smaller than
that of Mg, respectively. The in-plane lattice constant of
multilayer Mg
(1-2y)
Ti
y
X
y
scales roughly linearly with the Mg
content in the range considered. Therefore, an increased Mg
content exerts an increased tensile strain on the TiX layers,
which increases the formation energy.
Depositing all three metals simultaneously most likely
leads to disordered structures.
10
A full search in the complete
configurationspace for the optimal structure of a ternaryalloy,
as a function of its composition, is computationally prohibi-
tive. Therefore, we apply the following scenario to model a
disordered structure. We position all metal atoms at fcc sites
and assume that binary Mg-Ti alloys have randompositional
disorder. Such randomstructures are modeled in finite super-
cells by special quasirandom structures (SQSs).
17,18
We use
supercells containing 32 metal atoms to model SQSs of
Figure 1. Layered Mg
(1-2y)
Ti
y
X
y
structures for y = (a) 0.25, (b) 0.1875, and (c) 0.125. Mg, Ti, and X (=Al or Si) atoms are shown as
blue, black, and red colored spheres, respectively.
16
Table 1. Calculated Formation Energies, E
f
(eV/atom), of
Mg
(1-2y)
Ti
y
X
y
in the Layered Structures with Respect to
the Elements (see Figure 1)
compound X = Ti X = Al X = Si
Mg
0.5
Ti
0.25
X
0.25
0.206 -0.083 -0.132
Mg
0.625
Ti
0.1875
X
0.1875
0.141 -0.015 -0.040
Mg
0.75
Ti
0.125
X
0.125
0.093 0.010 0.006
r2010 American Chemical Society 1984 DOI: 10.1021/jz100386j |J. Phys. Chem. Lett. 2010, 1, 19821986
pubs.acs.org/JPCL
Mg
(1-x)
Ti
x
, x =0.25, 0.375, 0.5.
6,19
Adding a third element in
the compositionMg
(1-2y)
Ti
y
X
y
, X=Al or Si, means that half of
the Ti atoms should be replaced by X atoms. To examine the
preferential binding sites of the X atoms, the coordination
numbers N of each Ti by Ti atoms in their nearest-neighbor
shells are calculated for the SQSs of Mg
(1-x)
Ti
x
. We order the
Ti atoms in the SQSs of Mg
(1-x)
Ti
x
according to their N values
and consider two scenarios for replacing half of them by
X atoms. In the first scenario, Ti atoms with the largest N are
replaced, which leads to a maximumcoordination of Xatoms
by Ti or X atoms. We call this the high coordination (hc)
structure in the following. In the second scenario, Ti atoms
with the smallest N are replaced, giving X atoms that are
mostly surrounded by Mg atoms, which we call lowcoordina-
tion (lc). Note that in the lc (hc) structure, the Ti atoms that
remain in Mg
(1-2y)
Ti
y
X
y
all have a large (small) N. The hc and
lc structures represent extreme cases, and we have checked
that other scenarios of replacingTi by Xatoms give results that
are intermediate between these two cases.
The formation energies with respect to the elements of
these Mg
(1-2y)
Ti
y
X
y
structures are given in Table 2. They can
be compared to the formation energies of the corresponding
Mg
(1-2y)
Ti
2y
structures. Al and Si both have a stabilizing effect
on the Mg-Ti compounds, irrespective of where these atoms
are positioned, with Si having a larger effect than Al. In
addition, the stabilityof Mg
(1-2y)
Ti
y
X
y
decreases withdecreas-
ing y. On an absolute scale, only the structures with y = 0.25
for Al and y =0.25 and 0.1875 for Si are thermodynamically
stable with respect to decomposition into the elements. By
comparing the formation energies of the lc and hc structures
in Table 2, one observes that the relative ordering of the metal
atoms can have a sizable effect on the formation energies.
Perhaps somewhat counterintuitive, the lc case yields the
lowest energy. In the lc structures, the Al or Si atoms are
mainly surrounded by Mg atoms, and as discussed above,
Mg-Al and Mg-Si bonds are not particularly strong. How-
ever, the Ti atoms that remain inthe lc structure all are mainly
surrounded by Ti atoms. This minimizes the number of direct
Mg-Ti contacts that severely decrease the stability.
Comparing the formation energies of the layered struc-
tures, Table 1, to those of the disordered structures, Table 2,
one observes that the layered structures are more stable. This
can provide new ideas for structures with interesting proper-
ties, formed by immiscible elements.
We now turn to the energetics of the Mg-Ti-Al,Si hy-
drides. A fluorite-type cubic structure of Mg
(1-x)
Ti
x
H
2
is stable
for x J0.2.
5,6
In the following, we assume that replacing half
of the Ti atoms by Al or Si maintains the fluorite structure. We
use the structural models of the Mg-Ti-X alloys discussed
above and position hydrogen atoms in the tetrahedral inter-
stitial sites. For the disordered systems, we use the lc struc-
tures since these give the lowest energies for the alloys.
Subsequently, all cell parameters and geometries of the
hydrides are optimized.
The hydrogenation energies E
h
are given in Table 3,
calculated by subtracting the total energies of the alloy
Mg
(1-2y)
Ti
y
X
y
and the hydrogen molecule from that of the
hydride Mg
(1-2y)
Ti
y
X
y
H
2
.
The results clearly showthat the hydrides Mg
(1-2y)
Ti
y
X
y
H
2
are much less stable than their corresponding Mg
(1-2y)
Ti
2y
H
2
,
whether in the layered or in the disordered structures. In
contrast, the alloys Mg
(1-2y)
Ti
y
X
y
are more stable than the
corresponding Mg
(1-2y)
Ti
2y
; see Tables 1 and 2. This is an
example of van Mal's or Miedema's rule of reversed stability,
which states that stabilization of an alloy phase is accompa-
nied by a destabilization of its hydride.
20
Apparently, this rule
also holds if one compares the layered to the disordered
structures. The layered structures of the Mg
(1-2y)
Ti
y
X
y
alloys
are more stable then their corresponding disordered struc-
tures (compare Tables 1 and 2), whereas the hydrides
Mg
(1-2y)
Ti
y
X
y
H
2
are less stable in the layered structures than
those in the disordered structures; see Table 3. Moreover, as
Mg-Ti-Si alloys are more stable than Mg-Ti-Al alloys, one
may expect that Mg-Ti-Si hydrides are less stable than
Mg-Ti-Al hydrides. The results in Table 3 confirm this.
The hydrides of the layered alloys have a well-ordered
structure, with hydrogen atoms occupying positions at or
close to the tetrahedral interstitial sites, as one would expect
of a fluorite-type structure. In a fluorite MH
2
structure, each
metal atom M is cubically coordinated by eight hydrogen
atoms. Indeed, the radial distribution functions (RDFs) of the
M-H distances in layered Mg
(1-2y)
Ti
y
X
y
H
2
clearly show this
eight-fold coordination; see Figure 2. Such an atomic arrange-
ment is thought to be advantageous for fast hydrogen
kinetics.
4
The hydrides of the disordered lc structures of
Mg
(1-2y)
Ti
y
X
y
H
2
show a larger degree of disorder in the
positions of the H atoms, in particular, for compounds with
a high Mg content, as can be clearly observed in the RDF. One
may expect that H atoms close to Mg atoms like to assume
positions that are more typical for the six-fold octahedral
coordination observed in MgH
2
.
6
Around Al or Si atoms, one
expects to find a preference for a tetrahedral or an octahedral
coordination. Such competing preferential coordinations can
Table 2. Calculated Formation Energies, E
f
(eV/atom), of
Mg
(1-2y)
Ti
y
X
y
in the Disordered Structures with Respect to the
Elements
a
X = Ti X = Al X = Si
compound lc hc lc hc
Mg
0.5
Ti
0.25
X
0.25
0.156 -0.008 0.052 -0.066 0.006
Mg
0.625
Ti
0.1875
X
0.1875
0.153 0.025 0.091 -0.020 0.064
Mg
0.75
Ti
0.125
X
0.125
0.110 0.035 0.058 0.038 0.058
a
The construction of the lc and hc structures is discussed in the text.
Table 3. Calculated Hydrogenation Energies, E
h
(eV/H
2
), of
Mg
(1-2y)
Ti
y
X
y
H
2
in the Layered and the Disordered lc Structures
a
layered disordered lc
compound X = Ti X = Al X = Si X = Ti X = Al X = Si
Mg
0.5
Ti
0.25
X
0.25
H
2
-1.118 -0.178 0.335 -1.120 -0.358 -0.156
Mg
0.625
Ti
0.1875
X
0.1875
H
2
-0.906 -0.278 0.093 -0.981 -0.468 -0.240
Mg
0.75
Ti
0.125
X
0.125
H
2
-0.711 -0.317 -0.086 -0.807 -0.501 -0.434
a
The second and the fifth columns represent the formation energies
of the corresponding Mg
(1-2y)
Ti
2y
H
2
structures.
r2010 American Chemical Society 1985 DOI: 10.1021/jz100386j |J. Phys. Chem. Lett. 2010, 1, 19821986
pubs.acs.org/JPCL
be observed in the RDF of the disordered structures.
The layered structures enforce cubic coordination of all metal
atoms by hydrogens, thereby reducing the hydrogenation
energy.
A hydrogenation energy of -0.4 eV/H
2
gives an equilib-
riumhydrogenpressure of 1 bar at roomtemperature. Even
with this hydrogenation energy, the amount of heat released
in a very short time when a bulk material is loaded with
hydrogen can be a problem in some applications. Therefore,
depending on the application, a hydrogenation energy bet-
ween -0.15 and -0.4 eV/H
2
is required.
21
The hydro-
genation energies of pure Mg-Ti alloys are clearly outside of
this range. They are in fact larger or comparable to the
hydrogenation energy of pure Mg.
5,6
The addition of Al or Si
markedly destabilizes the hydrides and brings the hydrogena-
tionenergies into the range required for possible applications,
where Si has a larger effect than Al.
The effects are particularly large in the layered struc-
tures. Mg
(1-2y)
Ti
y
Si
y
H
2
in a layered structure is either
unstable or marginally stable for the compositions that
we studied. In contrast, the hydrogenation energy of
Mg
(1-2y)
Ti
y
Al
y
is -0.18 to -0.32 eV/H
2
. Moreover, the alloy
Mg
(1-2y)
Ti
y
Al
y
is stable with respect to decomposition into
the elements for y = 0.1875 and 0.25 and only marginally
unstable for y = 0.125; see Table 1. These properties
indicate that Mg
(1-2y)
Ti
y
Al
y
in layered form is a potential
material for hydrogen storage. Also, in the disordered
structures, the addition of Al or Si does destabilize the
hydrides, although the effects are not as large as those in
the layered structures. Again, Si has a larger effect than Al,
but the hydrogenation energies of all compositions studied
are markedly better than those of pure Mg or Mg-Ti. If one
demands that disordered Mg
(1-2y)
Ti
y
X
y
alloys are stable
with respect to decomposition into the elements, then
for X = Al and Si, one needs y = 0.25 and y = 0.1875,
0.25, respectively (see Table 2), which gives hydrogenation
energies ranging from -0.16 to -0.36 eV/H
2
.
COMPUTATIONAL METHODS
DFTcalculations are performed at the level of the general-
ized gradient approximation (GGA) using the PW91
functional.
22
We employ the projector augmented wave
(PAW) technique and a plane wave basis set,
23,24
as imple-
mented in the Vienna ab initio simulation package
(VASP).
25,26
Standard frozen core PAW potentials are used,
where the H 1s, Mg 3s, Al 3s3p, Si 3s3p, andTi 3p4s3d shells
are treated as valence shells. The plane wave kinetic energy
cutoff is set at 518 eV, and the Brillouin zone is integrated
using a regular k-point mesh with a spacing of at most
0.02
-1
. The volumes and shapes of the unit cells are
optimized, as well as the positions of the atoms. Total energies
of the optimized structures are finally calculated using the
tetrahedron method.
27
Details of the computational pro-
cedure can be found in ref 6.
AUTHOR INFORMATION
Corresponding Author:
*To whomcorrespondence should be addressed. E-mail: g.brocks@
tnw.utwente.nl.
ACKNOWLEDGMENT The authors acknowledge P. J. Kelly for
stimulating discussions. This work is part of the research
programs of Advanced Chemical Technologies for Sustainability
(ACTS) and the Stichting voor Fundamenteel Onderzoek der
Materie (FOM). The use of supercomputer facilities was
sponsored by the Stichting Nationale Computerfaciliteiten (NCF).
These institutions are financially supported by the Nederlandse
Organisatie voor Wetenschappelijk Onderzoek (NWO).
REFERENCES
(1) Buschow, K. H. J.; Bouten, P. C. P.; Miedema, A. R. Hydrides
Formed from Intermetallic Compounds of Two Transition
Metals ;A Special Class of Ternary Alloys. Rep. Prog. Phys.
1982, 45, 9371039.
(2) Niessen, R. A. H.; Notten, P. H. L. Electrochemical Hydrogen
Storage Characteristics of Thin Film MgX (X = Sc, Ti, V, Cr)
Compounds. Electrochem. Solid-State Lett. 2005, 8, A534A538.
(3) Niessen, R. A. H.; Notten, P. H. L. Hydrogen Storage in Thin
Film Magnesium-Scandium Alloys. J. Alloys Compd. 2005,
404, 457460.
(4) Vermeulen, P.; Niessen, R. A. H.; Notten, P. H. L. Hydrogen
Storage in Metastable Mg
y
Ti
1-y
Thin Films. Electrochem.
Commun. 2006, 8, 2732.
(5) Er, S.; Tiwari, D.; de Wijs, G. A.; Brocks, G. Tunable Hydrogen
Storage in Magnesium-Transition Metal Compounds: First-
Principles Calculations. Phys. Rev. B 2009, 79, 024105.
(6) Er, S.; van Setten, M. J.; de Wijs, G. A.; Brocks, G. First-
Principles Modelling of Magnesium Titanium Hydrides.
J. Phys.: Condens. Matter 2010, 22, 074208.
(7) Liang, G.; Schulz, R. Synthesis of Mg-Ti Alloy by Mechanical
Alloying. J. Mater. Sci. 2003, 38, 11791184.
(8) Rousselot, S.; Bichat, M. P.; Guay, D.; Roue, L. Structure and
Electrochemical Behaviour of Metastable Mg
50
Ti
50
Prepared
by Ball Milling. J. Power Sources 2008, 175, 621624.
(9) Borsa, D. M.; Baldi, A.; Pasturel, M.; Schreuders, H.; Dam, B.;
Griessen, R.; Vermeulen, P.; Notten, P. H. L. Mg-Ti-H Thin
Films for Smart Solar Collectors. Appl. Phys. Lett. 2006, 88,
241910.
Figure 2. Optimized structures of Mg
0.75
Ti
0.125
Al
0.125
H
2
in the
(a) layered and (b) disordered lc structures. The dashed and solid
lines give the RDFs and the integrated RDFs of metal-hydrogen
pairs, respectively.
r2010 American Chemical Society 1986 DOI: 10.1021/jz100386j |J. Phys. Chem. Lett. 2010, 1, 19821986
pubs.acs.org/JPCL
(10) Vermeulen, P.; van Thiel, E. F. M. J.; Notten, P. H. L. Ternary
MgTiX-Alloys: A Promising Route towards Low-Temperature,
High-Capacity, Hydrogen-Storage Materials. Chem.-Eur. J.
2007, 13, 98929898.
(11) Ekman, M.; Ozolin-s, V. Electronic Structure and Bonding
Properties of Titanium Silicides. Phys. Rev. B 1998, 57,
44194424.
(12) Ohnuma, I.; Fujita, Y.; Mitsui, H.; Ishikawa, K.; Kainuma, R.;
Ishida, K. Phase Equilibria in the Ti-Al Binary System. Acta
Mater. 2000, 48, 31133123.
(13) Kerimov, K. M.; Dunaev, S. F.; Sljusarenko, E. M. Investigation
of the Structure of Ternary Phases in Al-Mg-Ti, Al-Mg-V
and Al-Mg-Cr Systems. J. Less-Common Met. 1987, 133,
297302.
(14) Vajo, J. J.; Mertens, F.; Ahn, C. C.; Bowman, R. C., Jr.; Fultz, B.
Altering Hydrogen Storage Properties by Hydride Destabili-
zation through Alloy Formation: LiH and MgH
2
Destabilized
with Si. J. Phys. Chem. B 2004, 108, 1397713983.
(15) Wang, N.; Yu, W.-Y.; Tang, B.-Y.; Peng, L.-M.; Ding, W.-J.
Structural and Mechanical Properties of Mg
17
Al
12
and
Mg
24
Y
5
from First-Principles Calculations. J. Phys. D: Appl.
Phys. 2008, 41, 195408.
(16) Momma, K.; Izumi, F. VESTA: AThree-dimensional Visualiza-
tion System for Electronic and Structural Analysis. J. Appl.
Crystallogr. 2008, 41, 653658.
(17) Zunger, A.; Wei, S. H.; Ferreira, L. G.; Bernard, J. E. Special
QuasirandomStructures. Phys. Rev. Lett. 1990, 65, 353356.
(18) Ruban, A. V.; Simak, S. I.; Shallcross, S.; Skriver, H. L. Local
Lattice Relaxations in Random Metallic Alloys: Effective
Tetrahedron Model and Supercell Approach. Phys. Rev. B
2003, 67, 214302.
(19) van Setten, M. J.; Er, S.; Brocks, G.; de Groot, R. A.; de Wijs,
G. A. First-Principles Study of the Optical Properties of
Mg
x
Ti
(1-x)
H
2
. Phys. Rev. B 2009, 79, 125117.
(20) van Mal, H. H.; Buschow, K. H. C.; Miedema, A. R. Hydrogen
Absorption in LaNi
5
and Related Compounds: Experimental
Observations and Their Explanation. J. Less-Common Met.
1974, 35, 6576.
(21) Er, S.; de Wijs, G. A.; Brocks, G. Hydrogen Storage by Poly-
lithiated Molecules and Nanostructures. J. Phys. Chem. C
2009, 113, 89979002.
(22) Perdew, J. P.; Chevary, J. A.; Vosko, S. H.; Jackson, K. A.;
Pederson, M. R.; Singh, D. J.; Fiolhais, C. Atoms, Molecules,
Solids, and Surfaces: Applications of the Generalized Gradi-
ent Approximation for Exchange and Correlation. Phys. Rev.
B 1992, 46, 66716687.
(23) Bl ochl, P. E. Projector Augmented-Wave Method. Phys. Rev. B
1994, 50, 1795317979.
(24) Kresse, G.; Joubert, D. FromUltrasoft Pseudopotentials to the
Projector Augmented-Wave Method. Phys. Rev. B 1999, 59,
17581775.
(25) Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid
Metals. Phys. Rev. B 1993, 47, 558561.
(26) Kresse, G.; Furthm uller, J. Efficient Iterative Schemes for Ab
Initio Total-Energy Calculations Using a Plane-Wave Basis Set.
Phys. Rev. B 1996, 54, 1116911186.
(27) Bl ochl, P. E.; Jepsen, O.; Andersen, O. K. Improved Tetrahe-
dron Method for Brillouin-Zone Integrations. Phys. Rev. B
1994, 49, 1622316233.

Você também pode gostar