You are on page 1of 5

Journal of Alloys and Compounds 453 (2008) 138–142

Hydrogen sorption kinetics of MgH2 catalyzed with NbF5


Ye Luo, Ping Wang, Lai-Peng Ma, Hui-Ming Cheng ∗
Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences,
72 Wenhua Road, Shenyang 110016, China
Received 28 September 2006; received in revised form 18 November 2006; accepted 20 November 2006
Available online 26 December 2006

Abstract
The influence of NbF5 as an additive on the H-sorption kinetics of MgH2 was investigated. We found that the hydrogen sorption capacity and
sorption kinetics of the composites were dependent on the addition amount of NbF5 as well as the milling time. Fast kinetics are obtained using
2 mol% NbF5 . At 573 K, the MgH2 + 2 mol% NbF5 milled for 5 h can absorb 5 wt.% hydrogen in 12 s and 6 wt.% in 60 min, and desorb 4.4 wt.%
in 10 min and 5 wt.% in 60 min. DSC examination indicates that the desorption temperature is reduced by 64 K, and the activation energy for
hydrogen desorption is estimated to be ∼90 kJ/mol H2 from the Kissinger Plot, indicating that MgH2 is significantly activated due to the catalytic
effect of NbF5 . Furthermore, X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS) were performed to characterize the structural
evolution upon milling and dehydrogenation.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Hydrogen storage; Magnesium hydride; High-energy ball milling; Niobium fluoride; XPS

1. Introduction perature within 15 s [13]. Regarding the catalytic mechanism,


Barkhordarian et al. proposed that high valence state of transi-
Magnesium is considered as a promising candidate for hydro- tion metal and the tailoring effect of anion should be responsible
gen storage due to its high storage capacity with a theoretical for the observed catalytic enhancement in Mg–Nb2 O5 system
value of 7.6 wt.%, high abundance and low cost [1]. However, [15].
because of its high dissociation enthalpy (H = −74.5 kJ/mol) Fluorination treatment in an aqueous solution containing F−
and slow sorption kinetics, it cannot be widely adopted com- was developed to improve the H-storage properties of Mg-based
mercially [2]. As demonstrated experimentally, adding catalysts alloys. It was claimed that the formation of fluorinated surface
into MgH2 by mechanical milling can accelerate H-sorption, layer has a highly reactive and protective effect for H uptake
but within the thermodynamic limit [3–11]. The catalysts used [16]. Quite recently, we found that, as a source of both transition
include transition metals and/or their oxides, hydrogen stor- metal and fluorine anion, transition metal fluorides also offered
age alloys and carbon materials. Among them, transition metal a promising choice in constituting Mg-based hydrogen stor-
oxides such as Nb2 O5 or V2 O5 have recently attracted consider- age system. Their pronounced catalytic effect on the reversible
able interests due to their pronounced catalytic effect [1,12–15]. absorption/desorption processes of Mg has been demonstrated
Barkhordarian et al. reported that the 0.5 mol% Nb2 O5 catalyzed in the Mg–TiF3 system [17]. In the present paper, we further
MgH2 milled for 100 h could absorb about 7 wt.% hydrogen present the results of another system, Mg–NbF5 , which aims at
within 60 s and desorb almost all the hydrogen gas within 90 s combining the functions of both high-valence niobium and fluo-
in vacuum at 573 K [12]. Recently, Hanada et al. investigated rine anion. Again, the addition of fluoride results in a markedly
the ball-milled MgH2 doped with 1 mol% Nb2 O5 (particle size : improved H-capacity and sorption kinetics of Mg.
∼3.2 nm), and showed that the composite after dehydrogenation
at 473 K absorbed hydrogen of 4.5 wt.% even at room tem-
2. Experimental details

The initial magnesium hydride powder was prepared by ball milling mag-
∗ Corresponding author. Tel.: +86 24 2397 1788; fax: +86 24 2390 3126. nesium powder (purity > 99.9%, 300 mesh) under hydrogen atmosphere with an
E-mail address: cheng@imr.ac.cn (H.-M. Cheng). initial pressure of ∼1 MPa, followed by a long-period hydrogenation at 573 K.

0925-8388/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2006.11.113
Y. Luo et al. / Journal of Alloys and Compounds 453 (2008) 138–142 139

The process was repeated for three times to achieve a hydrogenation ratio of
∼80%, which was determined by volumetric method. Niobium fluoride (NbF5 ,
98% purity) was purchased from Sigma–Aldrich. Thus-prepared MgH2 was
mechanically milled with different amounts of NbF5 for different periods under
argon atmosphere using a SPEX 8000 mill. Two hardened steel balls of 12.75 mm
and four hardened steel balls of 6.35 mm in diameter were used for milling, and
the ball-to-powder weight ratio was about 40:1. All the sample handling was
performed in an Ar-filled glove box (MBraun), in which water/oxygen levels
were below 1 ppm.
The samples were characterized using X-ray diffraction (XRD, Rigaku
D/max 2400) with Cu K␣ radiation and X-ray photoelectron spectroscopy (XPS)
in a VG Escalab 250 spectrometer with a six channel hemispheric analyzer
photoelectron registration, whereby the data acquisition takes place in constant
energy analyzer mode with pass energy of E = 50 eV and Al K␣ radiation from
a monochromatic aluminum X-ray source, the binding energy calibration was
performed on the carbon C1s peak with energy of 284.6 eV. The binding energy
spectra was given after smoothing and fitting by XPSPEAK software, followed
by normalization for comparison.
Hydrogen storage properties of the MgH2 –NbF5 composites were examined
using an automatic Sieverts’ apparatus, in which the pressure and temperature Fig. 2. Hydrogen desorption profiles of the composite MgH2 + 2 mol% NbF5
data were simultaneously collected to allow a correction of the pressure change ball-milled for different periods and pure MgH2 milled for 5 h, at 573 K.
caused by the temperature variation. The leakage rate of the system in pressure
holding test (12 MPa hydrogen for 120 h) was lower than 3.1 × 10−8 mmol/s. absorption kinetics. An initial increase from 0.2 to 2 mol% NbF5
The mass of the sample for every measurement run was around 200 mg. Before
measurements, the system was degassed for 1.5 h with the sample cell being
leads to an increased capacity, as well as enhanced absorption
heated to 623 K. Absorption/desorption measurements were performed at var- kinetics. For the optimal composite containing 2 mol% NbF5 ,
ious temperatures with an initial pressure of 2 MPa and <10 Pa, respectively. about 5 wt.% hydrogen can be absorbed in about 12 s. Further
The desorption performance of the re-hydrogenated samples was also examined increasing the NbF5 amount to 3 mol%, however, results in a
by diffraction scanning calorimetry (DSC, Netzsch STA 409C). In the thermal decreased capacity. Clearly, the main hydrogen absorption phase
analysis measurements, high purity Ar (purity > 99.999%) was used as purge
gas. The heating rate was 5 K/min.
in the composites is Mg, and the added NbF5 acts as an effec-
tive catalyst. Too much addition of NbF5 , on the contrary, will
degrade the hydrogen storage capacity of the composite, as seen
3. Results and discussion
from the profile of the sample containing 3 mol% NbF5 . The
MgH2 + 2 mol% NbF5 composite, therefore, was selected for
3.1. Influence of the addition amount of NbF5
further investigation.
Fig. 1 shows the absorption kinetics at 573 K of the compos-
ites MgH2 –X mol% NbF5 (X = 0.2, 1, 2 and 3) and pure MgH2 3.2. Optimization of milling time
milled for 5 h. The hydrogen storage capacity was obtained after
taking all the components into account. It is observed that in Fig. 2 shows the desorption profiles at 573 K of the samples
all cases, the addition of NbF5 results in markedly improved prepared by milling MgH2 with 2 mol% NbF5 for a period rang-
ing from 1.5 to 10 h as well as the pure MgH2 milled for 5 h.
Apparently, a longer milling time leads to a more pronounced
kinetic enhancement, in particular during the initial periods.
However, it is also observed that increasing milling time to 10 h
causes a substantial decrease in the hydrogen capacity, about
1 wt.% lower than that of the 5 h sample. The hydrogen des-
orption properties of all composites are higher than that of pure
MgH2 , which almost does not desorb hydrogen at 573 K. Addi-
tionally, it is found that changing the milling time from 1.5 to
10 h has slight influence on the absorption property.

3.3. Hydrogen storage performance

Fig. 3 shows the hydrogen absorption behaviors of


MgH2 + 2 mol% NbF5 milled for 5 h at room temperature (RT,
∼293 K), 373 and 473 K under 2 MPa hydrogen atmosphere.
It should be pointed out that the sample was directly mea-
sured after initial desorption at 623 K for 1.5 h without any
Fig. 1. Hydrogen absorption profiles of pure MgH2 and the composite
MgH2 –X mol% NbF5 (X = 0.2, 1, 2 and 3) ball-milled for 5 h, at 573 K with pre-activation. At 473 K, the composite can absorb 4.5 wt.%
an initial hydrogen pressure of about 2 MPa. The hydrogen storage capacity was hydrogen within 20 s, and 5.2 wt.% hydrogen within 1 h. At
calculated after taking all components into account. 373 K, about 4 wt.% hydrogen can be absorbed within 4 min.
140 Y. Luo et al. / Journal of Alloys and Compounds 453 (2008) 138–142

Fig. 3. Hydrogen absorption profiles of room temperature (RT, ∼293 K), 373
and 473 K under 2 MPa for the composite MgH2 + 2 mol% NbF5 .
Fig. 5. XRD patterns for MgH2 + 2 mol% NbF5 (a) milled for 5 h and (b) dehy-
drogenated after milled for 5 h.
Even at RT, the composite can absorb 3.4 wt.% hydrogen
after 1 h hydrogenation treatment. It is noteworthy that, when
the magnesium hydride powder was re-prepared to achieve 3.4. Structural characterization and discussion
a hydrogen ratio of 85%, the hydrogen storage property was
much better than the results reported here. Fig. 5 shows the XRD patterns of MgH2 + 2 mol% NbF5 in the
It has been claimed by Hanada et al. that the behavior of first as-milled and dehydrogenated states. In the case of the as-milled
order reaction is a universal phenomenon for H-desorption of powders, the most intense Bragg peaks are identified as tetrag-
the ball-milled MgH2 with some suitable catalysts [18]. In our onal ␤-MgH2 (PDF 12-0697), and a metastable orthorhombic
efforts to understand the fast H-sorption kinetics arising upon phase ␥-MgH2 (PDF 35-1184) also can be observed, which has
the addition of 2 mol% NbF5 , the activation energies of samples been previously reported to form upon milling MgH2 [19]. After
MgH2 + 2 mol% NbF5 and pure MgH2 in the desorption process 5 h milling, the grain size of host MgH2 was estimated to be
were estimated from the Kissinger plot, which is deduced from 8 nm from Sherrer equation. In the dehydrogenated sample, the
the first order reaction [18]. From the Kissinger plot shown in most intense peaks are from Mg (PDF 35-0821) transformed
Fig. 4, the activation energy Edes for H-desorption for 2 mol% from ␤-MgH2 and ␥-MgH2 . Reflections from MgF2 (PDF 41-
NbF5 catalyzed and non-catalyzed MgH2 are estimated to be 1443) can also be observed, suggesting that MgH2 reacts with
90 and 120 kJ/mol, respectively. This clearly demonstrates the NbF5 to produce MgF2 . Additionally, MgO (PDF 45-0946) was
pronounced catalytic effect of NbF5 on the desorption process detected in both samples, which is likely due to the slight oxygen
of MgH2 . contamination.
Fig. 6a and b show the XPS spectra of MgH2 + 2 mol% NbF5
powders in the as-milled and dehydrogenated states. For com-
parison, the spectra of the commercial NbF5 powder are also
included. It is found that the milling results in the disappearance
of the Nb3d3/2 , Nb3d5/2 and F1s from NbF5 and appearance of
Nb3d3/2 , Nb3d5/2 from Nb and the F1s peak from MgF2 . Addi-
tionally, Nb3d3/2 , Nb3d5/2 peaks from niobium fluoride with
intermediate valence states between 0 and +5 were also detected
in the milled sample, which remained after desorption reaction.
However, the definite valence states cannot be identified due
to the lack of standard references. Without definite evidence
for local structure, we attribute these signals of intermediate
valence states to Nbx+ -containing compounds (0 < x < 5). Sim-
ilarly, it appears that there are other fluorides, such as residual
NbF5 , newly formed Nbx+ -containing fluorides (which cannot
be identified due to the lack of standard references) besides the
MgF2 in the sample on the basis of F1s spectra. These fluorides
Fig. 4. The Kissinger plot for the hydrogen desorption reaction for (a) pure
will affect the peak locations of F1s, as shown in Fig. 6 b. More-
MgH2 and (b) MgH2 + 2 mol% NbF5 milled for 5 h. The heating rate used in over, it is noteworthy that for the 5 h-milled MgH2 + 2 mol%
DSC measurements was 2, 5, 10 and 20 K/min. NbF5 sample, there should be numerous defects and strain for
Y. Luo et al. / Journal of Alloys and Compounds 453 (2008) 138–142 141

[23]. However, when MgH2 , 2 mol% Nb and 5 mol% MgF2 were


milled simultaneously under identical processing conditions, the
H-sorption kinetics enhancement was not as marked as that of
MgH2 + 2 mol% NbF5 (results not shown here). Nb and MgF2
are not essentially responsible for the improved overall sorption
behavior. Therefore, Nbx+ –containing compounds may play a
major role as a catalyst.
It has been well established that transition metal may facili-
tate the dissociation of hydrogen molecules and recombination
of hydrogen atoms toward the molecular state. As a H2 molecule
reaches a catalyst surface, the charge is donated from the H2 s-
orbital to the surface s- and d-states, followed by a back-donation
from the surface s- and d-states to the H2 anti-bonding state. The
interaction of the surface s- and d-electrons with H2 molecular
orbital facilitates the H2 dissociation on the surface [24]. When
H-desorption happens, the Mg–H bond of MgH2 is activated by
the catalyst, where the electrons of the bonding orbitals of MgH2
are donated to the unoccupied orbitals of the catalyst, accom-
panied by a back-donation from the electrons of the occupied
orbitals of the catalyst to the anti-bonding orbitals of MgH2 .
The electronic exchange reactions result in easier Mg–H dis-
sociation, thus accelerating recombination of hydrogen atoms
[25]. The valence electron configurations of Nbx+ is 5s1 4d4−x ,
compared with 5s1 4d4 of Nb. Clearly, the lower 4d orbitals occu-
pancy of Nbx+ allows an increased electron exchange with H2
molecules and Mg–H bond of MgH2 . As a result, Nbx+ -species
may possess a more pronounced catalytic effect than metallic
Nb. This is supported by the DSC examination of the compos-
ites, as shown in Fig. 7. At a heating rate of 5 K/min, the peak
temperature of the endothermic effect associated to hydrogen
desorption is observed to shift downward from 649 K for pure
MgH2 to 638 K for MgH2 + 2 mol% Nb, and further to 585 K
for MgH2 + 2 mol% NbF5 . It strengthens our speculation that
Fig. 6. X-ray photoelectron spectroscopy signal evolution showing the Nbx+ -containing compounds instead of Nb mainly contribute
occurrence of the partial fluorine transfer reaction 5MgH2 + 2NbF5 → to the overall sorption behaviors. Similarly, in their study on
5MgF2 + 2Nb + 5H2 . MgH2 + 2 mol% NbF5 powders (a) milled for 5 h and (b) MgH2 –Nb2 O5 system, Friedrichs et al. suggested that the par-
dehydrogenated after milled for 5 h as compared with commercial NbF5 powder. tially reduced Nb species with a wide range of valence might
play a major role as a catalyst [26].
the as-formed MgF2 , which might lead to the observed binding
energy shift of the F1s peak between the as-formed MgF2 and
that in dehydrogenated state (MgF2 can be identified in XRD
for the dehydrogenated sample).
The combined XRD and XPS characterizations suggest that
the redox reaction Eq. (1) may occur partially during milling
process.
5MgH2 + 2NbF5 → 5MgF2 + 2Nb + 5H2 (1)
The standard Gibbs free energy, G◦ f of MgH2 , MgF2 [20] and
NbF5 [21] are −85.42, −1140.47 and −1861.54 kJ/mol, respec-
tively, thus the total change G associated with reaction (1) will
be −1552.17 kJ/mol of MgH2 . This confirms the possibility of
reaction (1) from the thermodynamic potentials.
One may argue that the enhancement in H-sorption perfor-
mance for MgH2 + 2 mol% NbF5 results from Nb, formed either
upon milling or dehydriding, which has been reported to serve
as an effective catalyst [22]. The presence of MgF2 may also Fig. 7. DSC profiles of pure MgH2 , MgH2 + 2 mol% NbF5 , MgH2 + 2 mol% Nb
catalyze the H-sorption, as has been claimed by Deledda et al. milled for 5 h followed by hydrogenation. The heating rate was 5 K/min.
142 Y. Luo et al. / Journal of Alloys and Compounds 453 (2008) 138–142

Other than the speculated catalytic effect of Nb species, the [3] A. Zaluska, L. Zaluski, J.O. Ström-Olsen, J. Alloys Compd. 289 (1999)
function of F− anion may also deserve considerable attention 197.
in the Mg–NbF5 system. As proposed, in situ formed MgF2 [4] A. Zaluska, L. Zaluski, J.O. Ström-Olsen, J. Alloys Compd. 288 (1999)
217.
has a highly reactive and protective effect for H uptake, which [5] J.V. John, M. Florian, C.A. Channing, C.B. Robert Jr., F. Brent, J. Phys.
may further combines with the catalytic function of Nb species Chem. B 108 (2004) 13977.
to produce a “synergetic” effect. Alternatively, F− anion may [6] N. Hanada, T. Ichikawa, H. Fujii, J. Phys. Chem. B 109 (2005) 7188.
also directly participate in the generation of the catalytically [7] T. Ichikawa, N. Hanada, S. Isobe, H.Y. Leng, H. Fujii, Mater. Trans. 46
active species. Further investigations are required to check these (2005) 1.
[8] P. Wang, A.M. Wang, H.F. Zhang, B.Z. Ding, Z.Q. Hu, J. Alloys Compd.
possibilities. 313 (2000) 218.
[9] P. Wang, A.M. Wang, H.F. Zhang, B.Z. Ding, Z.Q. Hu, Acta Mater. 49
4. Conclusions (2001) 921.
[10] H. Imamura, M. Kusuhara, S. Minami, M. Matsumoto, K. Masanari, Y.
Sakata, K. Itoh, T. Fukunaga, Acta Mater. 51 (2003) 6407.
The hydrogen capacity and absortion/desorption kinetics of [11] W.F. Luo, J. Alloys Compd. 381 (2004) 284.
Mg are markedly improved by adding NbF5 . At 573 K, the [12] G. Barkhordarian, T. Klassen, R. Bormann, Scr. Mater. 49 (2003) 213.
composite MgH2 + 2 mol% NbF5 milled for 5 h can absorb [13] N. Hanada, T. Ichikawa, S. Hino, H. Fujii, J. Alloys Compd. 420 (2006)
5 wt.% hydrogen in about 12 s and 6 wt.% in 60 min, and des- 46.
[14] O. Friedrichs, T. Klassen, J.C. Sánchez-López, R. Bormann, A. Fernández,
orb 4.4 wt.% in about 10 min and 5 wt.% in 60 min. From
Scr. Mater. 54 (2006) 1293.
the Kissinger plot, the activation energy for the H-desorption [15] G. Barkhordarian, T. Klassen, R. Bormann, J. Phys. Chem. B 110 (2006)
was reduced from 120 kJ/mol for the non-catalyzed MgH2 to 11020.
90 kJ/mol after adding 2 mol% NbF5 . Combined XRD and XPS [16] F.J. Liu, S. Suda, J Alloy Compd. 231 (1995) 742.
examinations suggest that Nb species with a range of high [17] L.P. Ma, P. Wang, H.M. Cheng, J. Alloys Compd., in print.
[18] N. Hanada, T. Ichikawa, H. Fujii, J. Phys. Chem. B 109 (2005) 7192.
valence state may contribute to kinetic improvement. To gain
[19] C.Z. Wu, P. Wang, X.D. Yao, C. Liu, D.M. Chen, G.Q. Lu, H.M. Cheng,
a better understanding of MgH2 + NbF5 system, further investi- J. Phys. Chem. B 109 (2005) 22217.
gations are still required to clarify the role of F− anion as well [20] Y.J. Liang, Y.C. Che, X.X. Liu, Thermodynamics Data Handbook of Inor-
as the nature of active species. ganic Substances, 1st ed., Northeastern University Press, Shenyang, China,
1993, p. 220.
[21] Y.J. Liang, Y.C. Che, X.X. Liu, Thermodynamics Data Handbook of Inor-
Acknowledgment ganic Substances, 1st ed., Northeastern University Press, Shenyang, China,
1993, p. 260.
The financial support for this research from the Hundred [22] J. Huot, J.F. Pelletier, L.B. Lurio, M. Sutton, R. Schulz, J. Alloys Compd.
Talents Project of Chinese Academy of Sciences is gratefully 348 (2003) 319.
[23] S. Deledda, A. Borissova, C. Poinsignon, W.J. Botta, M. Dornheim, T.
acknowledged.
Klassen, J. Alloys Compd. 404–406 (2005) 411.
[24] M. Tsuda, W.A. Diño, H. Kasai, H. Nakanishi, H. Aikawa, Thin Solid Films
References 509 (2006) 157.
[25] K. Nobuhara, H. Kasai, W.A. Diño, H. Nakanishi, Surf. Sci. 566–568 (2004)
[1] G. Liang, J. Huot, S. Boily, A. Van Neste, R. Schulz, J. Alloys Compd. 292 703.
(1999) 247. [26] O. Friedrichs, J.C. Sanchez-Lopez, C. Lopez-Cartes, T. Klassen, R. Bor-
[2] J.F. Stampfer, C.E. Holley, J.F. Suttle, J. Am. Chem. Soc. 82 (1960) 3504. mann, A. Fernandez, J. Phys. Chem. B 110 (2006) 7845.

You might also like