Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Journal of Alloys and Compounds 617 (2014) 633–638

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Investigation of the structural and hydrogenation properties


of disordered Ti-V-Cr-Mo BCC solid solutions
C. Raufast, D. Planté, S. Miraglia ⇑
Univ. Grenoble Alpes, Inst. NEEL, F-38042 Grenoble, France
CNRS, Inst. NEEL, F-38042 Grenoble, France

a r t i c l e i n f o a b s t r a c t

Article history: Selected compositions in the Ti-Cr-V-Mo system (with the BCC structure-type) have been synthesized
Received 12 May 2014 and characterized for structural (crystalline structure, solidification microstructure) and thermodynamic
Received in revised form 9 July 2014 properties (equilibrium and reversible hydrogen storage capacity). We present as well the effect of
Accepted 10 July 2014
co-melting with a so-called activating phase that results in a secondary phase development and a
Available online 19 July 2014
subsequent enhancement of the hydrogen sorption kinetics. Ageing properties and applicability of such
materials for hybrid hydrogen storage systems are also discussed.
Keywords:
Ó 2014 Elsevier B.V. All rights reserved.
Hydrogen absorbing materials
Microstructure

1. Introduction reported as the hydrogen storage materials for hybrid hydrogen


storage vessels [4–7].
A hybrid hydrogen storage system, that is, a system containing a The work of Matsunaga et al. [6] teaches us that Mo substitution
metal hydride with high dissociation pressure together with com- may drastically enhance the dissociation pressure of BCC alloys.
pressed hydrogen (filling the empty space), displays an improved This paper reports on a microstructural study of selected composi-
volumetric hydrogen storage density. In addition, when operating tions in the Ti-Cr-V-Mo system which have been further character-
such a system hydrogen can be easily supplied even at low temper- ized for their reversible hydrogen storage properties. We present
atures, this allows to start the system on demand. In the frame of a as well the effect of co-melting with a so-called activating phase
national program (ANR funding) aimed at developing mobile appli- that results in a secondary phase precipitation and a subsequent
cations with hybrid hydrogen storage, we proposed TiCrV alloys enhancement of the hydrogen sorption kinetics.
with the BCC structure which have been studied for years since
the pioneering work of Akiba and Iba [1]. Based on analogy with 2. Experimental details
vanadium (a single metal and an isostructural system), the aim is
to monitor the relative stabilities of the hydrides and thereby mod- Intermetallic samples were prepared by induction melting of the pure elements
(Neyco, Sigma Aldrich products with a purity >99.95%) under 100 mbar of argon
ify the thermodynamic properties of these solid solutions of early
pressure using a High Frequency (HF) furnace with a water-cooled copper crucible
transition metals. A general feature of these materials is their and then molten one to three times each.
multi-plateau pressures, where only the upper plateau pressure The ingot obtained weighed 10–15 g. Mo addition in TiVCr sample requires spe-
area (i.e. that of the FCC dihydride or c phase) can be used for cial attention since Mo or un-melted additives are likely to remain in the alloys and
reversible hydrogen storage in normal condition [2,3]. Whereas because of molybdenum’s comparatively high melting point (2883 K). Therefore
Cr–Mo alloys were melted first in order to lower the alloy melting point, and then
most applications of BCC materials for hydrogen storage lie in Ti-Cr-V-Mo alloys were prepared. Sample ingots were re-melted three times in
the low to medium pressure range, these materials are proposed order to ensure their homogeneity. No special annealing procedure was performed
for high pressure applications (hybrid hydrogen storage vessels) except for the purpose of studying the effect of heat treatment. Regarding the so-
since their dissociation pressure can be tailored through transition called activating phase, the Zr7Ni10 doped samples were annealed at temperatures
from 1100 K to 1300 K for 10mn in the HF furnace. The batches of samples, obtained
metal substitution. In recent studies, Ti–Cr–Mn (C14 Laves phase:
after melting, were hand-crushed in air using a steel mortar. The rough powders
hexagonal MgZn2 structure) alloys, Ti-Cr-V-Mo (BCC phase) alloys, were used for further co-melting or PCT analysis, and the finer powders (200 lm)
and Ti–V–Mn (C14-BCC composite phase) alloys have been for XRD analysis. The structural characterizations of the samples were done using
a Siemens D-5000 X-ray diffractometer with Cu Ka radiation. A standard crystal cell
parameters refinement program based on least squares refinement methods, was
⇑ Corresponding author at: Univ. Grenoble Alpes, Inst. NEEL, F-38042 Grenoble, used for the calculations. Microstructural analysis was carried out using a JEOL-
France. JSM 840A scanning electron microscope. Sorption analyses were done on HERA
E-mail address: salvatore.miraglia@grenoble.cnrs.fr (S. Miraglia). PCT-PRO and HIDEN Isochema HTPS apparatus coupled with a thermal bath (Sievert

http://dx.doi.org/10.1016/j.jallcom.2014.07.089
0925-8388/Ó 2014 Elsevier B.V. All rights reserved.
634 C. Raufast et al. / Journal of Alloys and Compounds 617 (2014) 633–638

method). Prior to the Pressure–Composition–Isotherm (PCI) measurements, each


sample has been activated during 20 min under 18 MPa of hydrogen at ambient
temperature and then evacuated under secondary vacuum.

3. Results and discussion

3.1. Materials synthesis and structural analysis from selected


compositions of undoped samples

It has been shown by Akiba [8] that the hydrogen storage capac-
ities and the stability of the metal hydride in the range of hydrogen
pressure 0.003–5 MPa, are strongly dependent on the Ti/Cr ratio for
Ti–Cr–V solid solutions with a BCC structure. This is because the
affinity of hydrogen to titanium is largest and that to chromium Ti 0.66Cr1.31V0.42Mo0.51
is smallest among the three elements; in addition it is because
the lattice parameter increases linearly with Ti/Cr ratio. In the
preliminary investigations we chose to fix the Ti/Cr ratio at the
almost constant value of 0.5, while the Mo content increased at
the expense of the V concentration. The following nominal
compositions were therefore synthesized: Ti0.66Cr1.31V0.42Mo0.51,
Ti0.63Cr1.35V0.27Mo0.75, Ti0.6Cr1.29V0.09Mo0.96 and Ti0.57Cr1.26Mo1.17.
Standard X-ray powder diffraction analysis shows that the BCC
structure type is maintained over the investigated Mo substitu-
tional range. The diffraction diagram of the alloy with maximal
substituted Mo composition exhibited 2 BCC patterns with slightly
different lattice parameters which have been ascribed to cubic lat-
tices corresponding to different compositions. In the meantime, we Ti 0.63Cr1.35V0.27Mo0.75
found a variation of the BCC lattice parameter with Mo concentra-
tion that did not follow Vegard’s law. Since the atomic radii of Fig. 2. Evolution of the Mo dendritic growth with increasing Mo/V ratio.
vanadium and molybdenum are almost similar, we suspected that
this variation was related instead to variations of the Ti/Cr ratio
which is not strictly constant and lies in the range [0.4–0.5] (see – A good agreement between overall composition and nominal
Fig. 1). (target) composition is found at low Mo concentrations.
The above experimental observations prompted a composi- – At higher Mo concentrations, the dendritic flake-like micro-
tional and microstructural analysis. A complementary investiga- structure reveals Ti-rich zones (darker boundaries) that
tion was therefore undertaken by means of SEM coupled to EDX account for the variations of the Ti/Cr ratio with increasing
analysis. Fig. 2 shows the microstructures of the ingots prepared Mo substitution.
by HF melting. A dendritic crystal growth is clearly seen and seems
to be favored as the Mo/V content ratio increases, recalling that the This latter experimental observation corroborates the fact that
elaboration process (and therefore the cooling rate) is similar for the lattice parameter variation observed in Fig. 1 is related to a var-
all the samples. The two-phase character of the samples is thereby iation of the Ti/Cr ratio as already noted in the literature [1,8].
corroborated both by XRD and SEM.
The microstructural observations coupled with EDX analysis 3.2. Secondary phase precipitation (investigation of Zr7Ni10-doped
may be summarized as follows: samples)

Addition of a second phase (very liable to absorb hydrogen and


therefore called activating phase) often results in a significant
improvement of the hydrogen sorption kinetics [9]. The activating
phase can play a role of gate for hydrogen thus enhancing the
kinetics of formation of the metal hydride. The samples studied
in the previous section were thus subjected to Zr7Ni10 addition
and further co-melting.
X-ray diffraction patterns of the resulting materials (that have
not undergone any further annealing or heat-treatment) show that
the BCC peaks display a systematic shoulder attributed to the for-
mation of a binary Cr–Mo alloy on the base of EDX analysis and
pattern matching. It is then clear that the Cr migration in the sec-
ondary minor BCC phase leads to a change of the starting Ti/Cr
ratio.
Micrographs of TiVCr(Mo)–Zr7Ni10 composites display the pres-
ence of single Ti spots (Fig. 3). For all compositions, Mo is found to
be dissolved in the BCC major phase, however some Ti is accommo-
dated in the Zr/Ni rich intergranular phase.
The results of the microstructural observations on the TiVCrMo
series co-melted with Zr7Ni10 as an activating phase may therefore
Fig. 1. Lattice parameters of the Ti0.66Cr1.31V0.42Mo0.51 series. be summarized as follows:
C. Raufast et al. / Journal of Alloys and Compounds 617 (2014) 633–638 635

Mo,Cr,

Zr Ni +Ti

Ti

Fig. 3. Micrograph and corresponding chemical analysis of Ti0.6Cr1.29V0.09Mo0.96


co-melted with Zr7Ni10.

The major phase is a homogeneous Ti-V-Cr-Mo BCC solid solu-


tion with a composition fairly close to the target composition, the
second phase being a (Zr,Ni)-rich intergranular one with a rather
homogeneously distributed pseudo-cell microstructure. The typi-
cal mean size of the pseudo-cells is 50 lm. In most cases EDX anal-
ysis reveals that fractions of Ti, V and Cr have diffused in the
intergranular phases. In some cases the migrating species (Ti) is
not recombined.
It has been already pointed out by Itoh et al. [10] that a large
temperature gradient and the difference between liquidus and sol-
idus may cause formation of Ti-rich phases. It is our contention
that large temperature gradient exist in our HF melting set up
because of the geometry of the water-cooled copper crucible.

3.3. Solidification

As mentioned above, the intergranular phase can accommodate


a significant portion of transition metal species. This migration is
made at the expense of the initial stoichiometry of the disordered
BCC solid solution and may result in a lowering of the reversible
hydrogen capacity. A control of the solidification process of such
composites is therefore mandatory. Fig. 4. Micrographs of co-melted Ti25Cr50V25Mo5. From top to bottom: surface side,
We have investigated first the microstructure of as-cast materi- core, crucible side.
als. Three zones were selected namely the surface side (upper crust
prompted us to consider additional compositions at higher vana-
of the ingot), the core of the bulk material (inner part of the ingot)
dium stoichiometry. Fig. 6 shows the hydrogenation kinetics of
and finally the crucible side (outer part of the ingot in contact with
sample Ti0.3V1.7Cr0.7Mo0.3 performed under a H2 pressure of
the copper crucible). The micrographs (Fig. 4) show that surface
10 MPa. It is worth noting that the particle size of the hydroge-
side and core display the same microstructure. The presence of
nated powder is reduced as a result of hydrogen-induced
impurities at surface side reflects the fact that the surface acts as
decrepitation.
a getter for the furnace impurities. At the crucible side instead a
finer grain size is observed together with a finer microstructure.
This experimental observation is ascribed to the fact that, because 4.1. PCI measurements
of the crucible water-cooling, the crucible side undergoes faster
cooling rates once the HF heating has been switched off. The records of the PCI curves display significant slope and hys-
The above observations have been completed by a study of the teresis effects which are commonly observed in the case of disor-
effect of thermal annealing at 1000 °C on the pseudo-cellular dered solid solutions and/or composite materials.
microstructure. It is clearly seen on Fig. 5 that such a heat treat- A heat treatment at 1300 °C for one hour followed by quenching
ment no longer preserves the initial pseudo-cell microstructure. was undertaken and had a beneficial effect on the slope of the
equilibrium plateau. Note that in view of the foreseen applications
in our case such an annealing is not mandatory because of the large
4. Hydrogenation of the samples
pressure range used for the operation of hydride vessels. Fig. 7
shows the PCI characterization of three samples with the
As already reported for the Mo-free TiVCr system, the addition
respective compositions Ti0.5V1.9Cr0.6, Ti0.5V1.75Cr0.6Mo0.15 and
of Zr7Ni10 as a secondary phase facilitates the first hydrogenation
Ti0.75V1.75Cr1.35Mo0.15, the measurements were carried out at
process [9] that takes place at room temperature at moderate
293 K. The observed trends are briefly discussed.
hydrogen pressures (2.5–3 MPa). Hydrogenation was performed
on coarse pieces of ingot and this first activation resulted in a very
reactive powder that had to be handled in an inert atmosphere. We 4.1.1. Hydrogen uptake
chose then to perform hydrogenation at high pressures (10 MPa) in Fig. 7 shows a decrease of the hydrogen uptake from the lower to
order to select samples with the highest hydrogen uptake; this the upper curve. A correlation can be made with the corresponding
636 C. Raufast et al. / Journal of Alloys and Compounds 617 (2014) 633–638

Fig. 5. Smearing of the pseudo cellular microstructure (left) after a 1100 °C annealing. Heat-treated sample micrograph (right).

of the unit cell volume [12]. This picture has been successfully used
for the Ti–V–Cr systems where the plateau pressure increases with
decreasing lattice constants associated with increasing Cr content
[13,14]. The same holds for the Ti–V–Cr–Fe system and allows a
correlation between the Ti/(Cr + Fe) ratio and the enthalpy of the
hydrides [15]. However, for some BCC systems this picture has to
be refined. In the Ti–V–Cr–Mn system Matsunaga et al. have shown
that logarithm of the dissociation pressure increases with the bulk
modulus and decrease with the cell volume [7]. Actually, the
chemical effect of Mo substitution upon equilibrium of the
hydrides is not straightforward and the relationship of the dissoci-
ation pressures with the bulk moduli of Mo-substituted alloys have
Fig. 6. Hydrogenation kinetics of Ti0.3V1.7Cr0.7Mo0.3 at 20 °C and 10 MPa. been investigated by the same authors [6] who found a good cor-
relation between dissociation pressure and B/V0, where B and V0
are the bulk modulus and the equilibrium volume of an alloy,
respectively.

4.1.3. Plateau slope and hysteresis


The PC isotherms of samples Ti0.5V1.9Cr0.6 and Ti0. 5V1.75
Cr0.6Mo0.15 (with the same Ti/Cr ratio) display the same slope fac-
tor. When increasing the Ti/Cr ratio (sample Ti0.75V1.75Cr0.35Mo0.15),
a significant increase of the slope factor is observed. For instance
Coluzzi et al. [16] have shown that the slope of the pseudo-plateau
of PCT isotherms of TiVCrMn BCC alloys during hydrogen absorp-
tion turns out to be due to the occupation by H of tetrahedral sites
of progressively higher energy. It has been already quoted that dis-
Fig. 7. PC isotherms at 20 °C of TiVCr and TiVCrMo-based alloys. ordered alloys such as those of the Ti–Cr–V system tend to have a
sloping plateau because their equilibrium pressure depends very
much on their composition. This effect reflects a not complete
decrease of the vanadium content. Regarding the effect of Mo addi-
equilibrium referred to as para-equilibrium. A correct thermody-
tion, it is seen that a slight addition of Mo leads to a very small
namic treatment by Flanagan can be found in Ref. [17]. Contrarily
change of the hydrogen uptake (lower curve and middle curve),
to what has been observed in the TiCrMn–H system [7] where hys-
whereas a comparison of the middle curve and the upper curve
teresis decreased with increasing the Ti content and decreasing the
(compositions with the same Mo content) shows the effect of the
Mn content; no obvious trend here is worth of notice.
vanadium concentration on the hydrogen uptake. Note however
that, the PC isotherm of sample Ti0.75V0.75Cr1.35Mo0.15 is not much
informative since complete transformation could not be reached 4.2. Practicability
because of the limited available H2 commercial pressure (restricted
to 20 MPa). Actually the parabolic Sievert-like saturation curve Following an optimization of the formula, the most suitable
could not be observed under these experimental conditions. composition in regard of the reversible capacity (2 wt%) and
operating pressure (between 0.1 and 10 MPa) was found to be
Ti0.3V1.7Cr0.7Mo0.3. A series of isotherms (Fig. 8) was then recorded
4.1.2. Equilibrium pressure in order to access the thermodynamic parameters by means of
A comparison of the isotherms of Ti–V–Cr and Ti-V-Cr-Mo Van’t Hoff plots. The effective storage capacity is obtained from
alloys shows the expected increase of the hydride equilibrium the length of the plateau region. The hysteresis is an important
pressure due to Mo addition. A further comparison between Ti0.5- characteristic that determines the applicability of the alloy as it
V1.75Cr0.6Mo0.15 and Ti0.75V1.75Cr0.35Mo0.15 (2 samples with equal generally reduces the efficiency of the system. In addition, a small
Mo concentration) shows that an increase of the Ti/Cr leads to an slope factor is required and the change of the equilibrium pressure
increase of the hydride equilibrium pressure. A correct treatment induced by this slope factor should be as small as possible. If a dis-
of the correlation between dissociation pressure and crystal chem- tinct desorption pressure is necessary, a sloping plateau reduces
istry requires the knowledge of the radius of hydrogen insertion either the useful hydrogen storage capacity or increases the neces-
site [11]. It has been known however that, in BCC intermetallics, sary metal hydride temperature. Hysteresis and slope factor both
the decrease of the dissociation pressure agrees with the increase reduce the efficiency of the sorption system.
C. Raufast et al. / Journal of Alloys and Compounds 617 (2014) 633–638 637

Fig. 8. PC isotherms of Ti0.3V1.7Cr0.7Mo0.3. Fig. 10. Cycling test of Ti0.3V1.7Cr0.7Mo0.3 at 20 °C.

plateau. This allowed us to extract enthalpy and entropy terms


(see Table 1) for the extreme cases (i.e. more and less stable
hydride states).

4.3. Cyclability

Intermetallic hydrides are sometimes quoted as being poor


cyclable materials and may be subject to gradual disproportion-
ation upon hydrogen sorption cycling. For Ti–Cr–V BCC alloys it
was found that vanadium content and homogeneity have impor-
tant roles for cyclic durability [18]. Release of strain effects was
also found to have a beneficial effect on cyclability. Hydrogenated
alloy Ti0.3V1.7Cr0.7Mo0.3 was therefore investigated for its ageing
properties and cyclability. The absorption/desorption loops were
conducted at 293 K with a pressure under 10 MPa. This test
showed the good reversible capacity of the sample even after 10
charging/discharging cycles (see Fig. 10). Note a loss of 0.15 wt%
of the maximum capacity upon the first absorption/desorption
Fig. 9. Schematic description of the relevant parameters used to describe slope and
hysteresis effects in PC isotherms.
cycle. This result is promising in view of hydride tank applications.
The thermodynamics aspects remain the same after the test which
confirms the relative good stability of the compound. A further
X-ray diffraction check did not reveal any disproportionation
For the ideal case (no hysteresis, no plateau slope), the logarith- effect.
mic equilibrium pressure can be drawn against 1/T giving way to
the so-called Van’t Hoff plot whereby DH and DS are determined. 5. Conclusions
These latter values being determined, an equilibrium diagram
(Peq vs. T) can then be drawn. However, due to the hysteresis and The effect of Mo addition on a series of TiCrV BBC alloys with a
plateau slope effect, an accurate and realistic description of ‘‘real’’ fixed Ti/Cr ratio of 0.5 has been investigated. Varying the Mo/V
metal hydrides requires a distinct treatment of the absorption and ratio leads to significant changes of the hydrogen uptake and the
desorption isotherms respectively. We have chosen to define 5 equilibrium pressure. Among the investigated compositions, the
parameters in order to completely describe the PC isotherm. The Ti0.3V1.7Cr0.7Mo0.3 alloy exhibited good hydrogenation properties
first one is a slope factor: ad and af with respect to the formation with respect to effective hydrogen capacity and dehydrogenation
(absorption) and dissociation (desorption) state. Then come 3 pres- pressure under pressures reported as working pressures for hybrid
sure parameters (Pmin, Pmax and Pmid) taken respectively at both hydrogen storage vessels. It is important to keep in mind that the
ends and at the middle of the equilibrium plateau. Finally the investigated alloy was single phase. Thermodynamic characteriza-
width of the desorption and absorption equilibrium plateaus (Ld tion and ageing tests have been conducted on single phase materi-
or Lf). This description is summarized in Fig. 9. als knowing that there is room for kinetics improvement by means
In order to fully circumvent the operating conditions of the of Zr7Ni10 addition. A processing method for composite materials
selected alloy (i.e. to know the complete domain of formation or (which exhibit enhanced sorption kinetics) has been assessed
dissociation of hydride) we determined for both cases the mini- and validated.
mum and maximum pressure point. We defined the two points The alloy did not practically change structure and phase compo-
by the point where PCI curve deviate from 5% of the equilibrium sition after dehydrogenation. The maximum hydrogen capacity of

Table 1
Thermodynamics relevant parameters for hydride formation and dissociation state of Ti0.3V1.7Cr0.7Mo0.3 sample.

Formation Dissociation
DH (kJ mol 1) DS (kJ mol 1
K 1
) DH (kJ mol 1) DS (kJ mol 1
K 1
)
Minimum pressure (stable state) 32.29 132.47 45.34 163.06
Middle pressure 35.24 152.71 29.18 119.69
Maximum pressure (instable state) 36.01 159.63 29.61 126.09
638 C. Raufast et al. / Journal of Alloys and Compounds 617 (2014) 633–638

the alloy was 2.25 wt% at 273 K under 10 MPa H2. The plateau [7] Y. Kojima, Y. Kawai, S. Towata, T. Matsunaga, T. Shinozawa, M. Kimbara, J.
Alloys Comp. 419 (2006) 256–261.
slope in the PCI was moderate, and the hydrogen desorption pres-
[8] S.W. Cho, C.S. Han, C.N. Park, E. Akiba, J. Alloys Comp. 288 (1999) 294–298.
sure was in the range of 0.1–2 MPa at 273 K. [9] S. Miraglia, P. de Rango, S. Rivoirard, D. Fruchart, J. Charbonnier, N. Skryabina, J.
Alloys Comp. 536 (2012) 1–6.
Acknowledgement [10] H. Itoh, H. Arashima, K. Kubo, T. Kabutomori, J. Alloys Comp. 330–332 (2002)
287–291.
[11] K. Kubo, H. Itoh, T. Takahashi, T. Ebisawa, T. Kabutomori, Y. Nakamura, E.
The authors acknowledge funding from the Project ANR-08- Akiba, J. Alloys Comp. 356–357 (2003) 452–455.
PANH-013. [12] S. Fujitani, I. Yonezu, T. Saito, N. Furukawa, E. Akiba, H. Hayakawa, S. Ono, J.
Less-Common Met. 172–174 (1991) 220.
[13] M. Okada, T. Kuriiwa, T. Tamura, H. Takamura, A. Kamegawa, J. Alloys Comp.
References 330–332 (2002) 511–516.
[14] T. Tamura, T. Kazumi, A. Kamegawa, H. Takamura, M. Okada, J. Alloys Comp.
[1] E. Akiba, H. Iba, Intermetallics 6 (1998) 461–470. 356–357 (2003) 505–509.
[2] J.J. Reilly, R.H. Wiswall, BNL Report 1 (1972) nr 16546. [15] Y. Yan, Y. Chen, H. Liang, C. Wu, M. Tao, J. Alloys Comp. 427 (2007) 110–114.
[3] K. Fujita, Y.C. Huang, M. Tada, J. Jpn. Inst. Met. 43 (1979) 601–604. [16] B. Coluzzi, A. Biscarini, G. Mazzolai, F.M. Mazzolai, A. Tuissi, F. Agresti, S. Lo
[4] N. Takeichi, H. Senoh, T. Yokota, H. Tsukuta, K. Hamada, H.T. Takeshita, Int. J. Russo, A. Maddalena, P. Palade, G. Principi, J. Alloys Comp. 456 (2008) 118–
Hydrogen Energy 28 (2003) 11–21. 124.
[5] M. Shibuya, J. Nakamura, H. Enoki, E. Akiba, J. Alloys Comp. 475 (2009) 543– [17] T.B. Flanagan, W.A. Oates, J. Alloys Comp. 404–406 (2005) 16–23.
545. [18] H.C. Lin, K.M. Lin, K.C. Wu, H.H. Hsiung, H.K. Tsai, Int. J. Hydrogen Energy 32
[6] T. Matsunaga, M. Kon, K. Washio, T. Shinozawa, M. Ishikiriyama, Int. J. (2007) 4966–4972.
Hydrogen Energy 34 (2009) 1458–1462.

You might also like