Vous êtes sur la page 1sur 11

Author's personal copy

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4

Available online at www.sciencedirect.com

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

Non-isothermal kinetics and in situ SR XRD studies


of hydrogen desorption from dihydrides of binary
TieV alloys

S. Suwarno a,*, J.K. Solberg a, J.P. Mæhlen b, R.V. Denys b, B. Krogh c,


E. Ochoa-Fernández c, B.T. Børresen c, E. Rytter c, I.E. Gabis d,
V.A. Yartys a,b,**
a
Department of Materials Science and Engineering, NTNU, NO-7491 Trondheim, Norway
b
Institute for Energy Technology, P.O. Box 40, NO-2027 Kjeller, Norway
c
Statoil ASA Research Centre, Rotvoll, NO-7005 Trondheim, Norway
d
Department of Physics, Saint-Petersburg State University, St. Petersburg, Russia

article info abstract

Article history: Phase transformations during dynamic dehydrogenation of Ti1xVxH2 (x ¼ 0.1; 0.2; 0.3) were
Received 5 July 2013 studied using in situ Synchrotron X-Ray Diffraction (SR XRD) and non-isothermal kinetics
Received in revised form experiments. The main dehydrogenation path for g-Ti1xVxH2 was found to be
21 August 2013 g / d / b / balloy. Body-centred tetragonal d-hydride was found to be an intermediate
Accepted 23 August 2013 phase of the g / b transformation in Ti0.8e0.9V0.1e0.2H2. TDS, in situ SR XRD and iso-
Available online 1 October 2013 conversional kinetics studies showed that hydrogen desorption from Ti1xVxH2 is
composed of simultaneous reactions taking place between 300 and 600  C. The effective
Keywords: activation energy of hydrogen desorption depends on the vanadium contents and the re-
Titanium action pathway, increasing from 21 kJ/mol H2 (g / d) to 60e110 kJ/mol H2 (d / b).
Vanadium Copyright ª 2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
Metal hydride reserved.
In situ synchrotron X-ray diffraction
Thermal desorption spectroscopy
Isoconversional kinetics

1. Introduction 150 kg H/m3) [2]. Ti1xVx alloys form a variety of hydrides


containing up to 2 at. H/Me. Their crystal structures depend on
Titanium and vanadium form disordered BCC-type solid so- the alloy composition and the amount of absorbed hydrogen.
lution alloys existing in a broad range of vanadium content Hagi et al. [3] have shown that at low hydrogen content, TieV
between 2.7 at.% to 80 at.% V [1]. TieV-based hydrides are able alloys form a monohydride with a BCC crystal structure, b-
to reach high densities of hydrogen (w4 wt.% H or about Ti1xVxHy, y ¼ 0 to w1.4. Its lattice parameter increases

* Corresponding author. Present address: Uttrecht University, 3511 AC Utrecht, Netherlands. Tel.: þ31 649290657.
** Corresponding author. Department of Materials Science and Engineering, NTNU, NO-7491 Trondheim, Norway. Tel.: þ47 63 80 64 53;
fax: þ47 63 81 29 05.
E-mail addresses: S.Suwarno@uu.nl, warnoise@yahoo.com (S. Suwarno), volodymyr.yartys@ife.no (V.A. Yartys).
0360-3199/$ e see front matter Copyright ª 2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijhydene.2013.08.103
Author's personal copy

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4 14705

linearly with the hydrogen content [3]. At high hydrogen In-situ Synchrotron X-ray diffraction (SR XRD) experiments
densities (H/M ¼ 1.45e2), the crystal structure of the formed g- were done in an experimental setup identical to that used in
dihydride is of the CaF2 type. There have been several studies the work by Denys et al. [11]. A fine powder sample was put
on the properties of titanium vanadium hydrides including an into a double capillary arrangement, including a 0.5 mm
X-ray structural study [4], studies of hydrogen diffusion [5e7] diameter quartz glass capillary of 0.01 mm wall thickness
and thermodynamics [6]. However, much less attention has placed inside a 0.7 mm diameter capillary of 0.02 mm wall
been addressing studies of the mechanisms of the hydroge- thickness. The capillary setup was attached to a goniometer
nation and dehydrogenation processes [8,9]. Such studies are head and was connected to a vacuum pump with flexible
important as Ti-rich TieV alloys are relevant for hydrogen tubing that allowed the sample to wobble during the X-ray
storage and show a potential in applications for hydrogen exposure. Data was collected using a MAR2300 image plate
storage and separation from gaseous mixtures at medium to detector with a sample to detector distance of 199.54 mm. The
high temperatures [10]. General Structure Analysis System (GSAS) [12] software was
In the present work, we used in situ synchrotron X-ray used for Rietveld refinement of the powder patterns [13] using
diffraction (SR-XRD) and thermal desorption spectroscopy profiles of the pseudo-Voigt function [14,15]. In situ SR-XRD
experiments to study hydrogen desorption from Ti1xVxH2, studies were performed with a synchrotron radiation wave-
x ¼ 0.1e0.3. Correlations between the hydrogen desorption length of 0.72085  A at a station BM01A, SwisseNorwegian
behaviour during thermal desorption spectroscopy studies Beam Lines (SNBL), at the European Synchrotron Radiation
and the structural transformations observed by SR XRD are Facility (ESRF), Grenoble, France.
reported and discussed.

3. Results
2. Materials and experimental methods
3.1. Alloy microstructure and compositions
Titanium sponge (99.7%) and vanadium turning (99.7% metal
basis) were melted together in argon environment using an The microstructure of all Ti1xVx alloys showed the presence
Edmund Bühler arc melter. Melting and remelting were done of very large grains, w200e800 mm, as observed in the light
at least three times to ensure chemical homogeneity. During optical microscope studies. Quantitative chemical composi-
the melting, zirconium was used as a getter material to tion analyses were performed by EPMA. Point by point mea-
reduce oxygen contamination during the alloy synthesis. surement was done along a line with a step length of 1 mm. A
Then the alloy buttons were cut into pieces for the hydroge- statistical average was calculated from 20 points. The Ti and V
nation process. The alloy microstructure and chemical contents in the Ti0.7V0.3 sample was 71  1.5 at.% and
composition were examined using a Leica MeF4 light optical 29  1 at.%, respectively. The Ti0.8V0.2 alloy was composed of
microscope and a JEOL JXA-8500F electron probe micro- 79  2 at.% Ti and 20  2 at.% V. For Ti0.9V0.1, the composition
analyzer (EPMA). was 89  2 at.% Ti and 10  0.5 at.% V, respectively. Thus, the
Hydrogenation was done in a Sievert’s-type thermal prepared samples of Ti1xVx were chemically homogenous
desorption spectroscopy (TDS) setup. Approximately 3e4 g of alloys well within the nominal composition.
alloy pieces were placed in an autoclave connected to a vac-
uum pump. When the pressure reached w5$105 mbar, the 3.2. Thermal desorption spectroscopy
autoclave was placed into a furnace. The sample was heated
to 600  C with a heating rate of 20 K/min and held at 600  C for A fully hydrogenated sample Ti1xVxH2 containing about
30 min. The furnace was then switched off, de-attached and 3.9 wt.% H was used in each TDS measurement. Fig. 1a shows
the autoclave was cooled down in air. When the sample hydrogen desorption spectra during dehydrogenation of
temperature reached 500  C, 20 bar of hydrogen gas was Ti1xVxH2. The hydrogen desorption traces of TiH1.9 are
introduced into the autoclave for hydrogenation. This process included for comparison. The hydrogen desorption spectra of
was repeated two times to get fully hydrogenated samples. Ti1xVxH2 showed that desorption started at w100  C and that
Thermal desorption spectroscopy studies were done in vac- the hydrogen flow rate increased as the sample temperature
uum, w1  105 mbar, at 2, 5 or 10 K/min heating rates; the increased. Further heating led to a maximum hydrogen flow
experiments span from room temperature to 800  C. rate reached at a certain temperature (the TDS main peak
Hydrogen desorption experiments were also performed in temperature). At a later stage, the hydrogen flow rate
a differential scanning calorimeter coupled with thermal decreased down to almost zero at 800  C, which indicates that
gravimetric analysis (DSC/TGA) and in a temperature pro- hydrogen was completely desorbed from the sample at that
grammed reaction (TPR) experiment. During the DSC/TGA temperature. The TDS main peak temperature was a unique
experiment, 30 ml/min argon flow was used during the heat- characteristic for each alloy composition, with the most stable
ing at a rate of 5 K/min from room temperature to 800  C. alloy hydride having the highest TDS temperature. Therefore,
During the TPR measurements the flow rate was 400 ml/min, the TDS main peak temperature was used as a measure of the
and the heating rate was 5 K/min. A Thermostar Mass Spec- thermal stability of the hydrides.
trometer (MS) from Pfeiffer Vacuum GmbH was connected to The amount of desorbed hydrogen during the dehydroge-
the TPR outgoing gas flow for measuring the gas composition. nation was calculated by integrating the hydrogen flow rate
The MS was calibrated to determine concentrations of H2 in over the desorption time, Fig. 1b. Based on the integrated TDS
the outgoing gas. graphs, hydrogen desorption from the TieV hydrides could be
Author's personal copy

14706 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4

6
Experimental data
Fitted data
5
Individual peaks fitting

H flow (cm3/g.mim)
2
3

2
3
1
1

100 200 300 400 500 600 700


o
Temperature ( C)

Fig. 2 e Hydrogen desorption spectrum of Ti0.8V0.2H2


during the heating (heating rate 2 K/min) in vacuum at
temperatures from 30  C to 800  C. The hydrogen
desorption spectrum consists of 3 curves related to three
individual hydride phase transformations during the
dehydrogenation.

The amount of hydrogen desorbed during the second step was


1.55 wt.%, or 0.77 H/M. The estimated amount of desorbed
hydrogen in the remaining hydride (the 3rd step) was 1.3 wt.%,
H/M ¼ 0.67.

3.3. Kinetic analysis


Fig. 1 e Thermal desorption spectra of Ti1LxVxH2: (a)
Hydrogen flow rate during non-isothermal The kinetics of a phase transformation can be expressed by an
dehydrogenation at 1 3 10L5 mbar. The thermal stability equation describing the reaction rate. In case of reactions
characterized by main peak temperature depends on the V involving solid and gas phases, it can be presented as [16]
contents; (b) Integrated amounts of hydrogen desorbed.
Three segments can be identified and related to the dx
¼ KT f ðxÞKP ðPÞ (1)
dehydrogenation steps. dt
where x is the extent of conversion, i.e. the fraction of con-
verted phase at time t, KT and Kp are temperature and
divided into 3 segments. As can be seen from Fig. 1b, the 1st pressure dependent kinetic rate constants, respectively, and
segment correlated with the first step of dehydrogenation, i.e. f(x) is a reaction model equation describing the trans-
low temperature hydrogen desorption from the FCC g-hydride. formation reaction mechanism. KT can be expressed using
The hydrogen desorption mode then changed, starting at an Arrhenius relation, i.e. KT ¼ A exp (E/RT), where A is a
about 300e400  C depending on the alloy composition. This pre-exponential factor, E is the activation energy, and T is
second mode of hydrogen desorption (the 2nd segment) was the temperature. For desorption, Kp is approximately equal
the fastest one (large curve gradients). In agreement with the to 1  P/Peq where P is the partial pressure of the hydrogen
TDS experiments, results from DSC/TGA as well as from the product during reactions, and Peq is the equilibrium pressure
TPR experiments showed that hydrogen desorption of of the hydrides. When the experiment is performed in vac-
Ti0.8V0.2H2 was composed of three steps (see Supplementary uum, the pressure dependency can be eliminated since the
information, Fig. S1). partial pressures of gas as a reaction product is very low and
To further analyse the dehydrogenation steps of g- assumed to be negligible. Thus, for the non-isothermal case
T1xVxH2, the hydrogen desorption spectrum curve from g- with a linear heating rate b ¼ dT/dt, by setting Kp ¼ 1 and by
T0.8V0.2H2 was deconvoluted into three individuals curves, replacing KT with the Arrhenius term, equation (1) can be
marked 1e3 in Fig. 2 using the pseudo-Voigt function. A written as
Pearson correlation coefficient, R2, of 0.99 indicated a good
dx A
quality of the fit. By simple calculations of the areas under the ¼ expðE=RTÞf ðxÞ (2)
dT b
curves, the amount of desorbed hydrogen in each step was
estimated. The amount of desorbed hydrogen during the 1st By rearranging equation (2) and by integrating over the
step (1st curve) was calculated to be 1.1 wt.% or H/M ¼ 0.55. transformation process, the reaction model, g(x) can be
Author's personal copy

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4 14707

expressed in terms of temperature, activation energy and composition, indicating H content-related changes in the
heating rate, as follows: desorption mechanism during the dehydrogenation process.
For Ti0.9V0.1H2 the Ex graph followed a slightly different trend
Zx ZT
dx A at high x values as compared to the graphs for Ti0.8V0.2 and
gðxÞ ¼ ¼ expðE=RTÞdT (3)
f ðxÞ b Ti0.7V0.3 indicating that the last step of hydrogen desorption
0 0
from the Ti0.9V0.1Hx might follow a different path and possibly
The integral on the right hand side of equation (3) is occurs by a different mechanism than in the other two cases.
generally called the temperature integral and can be solved by During the determination of Ex by the isoconversional
using approximations or by numerical integration. Such so- method, a particular reaction model f(x) was not adopted, but
lutions are the so-called isoconversional methods [16,17]. One even though the values of the activation energy might be used
solution of equation (3) is the following expression proposed to predict a reaction mechanism. A phase transformation can
by Kissinger, Akira, Sunose (KAS) [18,19]: be assumed to follow a single reaction model when Ex remains
relatively constant at increasing extent of conversion [16].
bi Ex
ln ¼ CK ðxÞ  (4) Thus, by observing Fig. 3, it can be seen that reaction mech-
T2xi RTxi
anisms of hydrogen desorption within each segment for all
where C is a constant, and the index i identifies an experiment hydrides studied were composed of more than a single pro-
done at a heating rate i. Therefore, by performing a series of cess, except for Ti0.9V0.1Hx within the 3rd segment. In other
non-isothermal experiments at different linear heating rates, words, only for x > x2 the dehydrogenation of Ti0.9V0.1Hx
one can find the dependency of the activation energy on occurred by a single reaction mechanism.
temperature or on extent of conversion. This procedure is the To determine possible reaction mechanisms, model fitting
so-called model-free method by which one can find the acti- has been performed within each segment for every deconvo-
vation energy without previous knowledge of the reaction luted hydrogen spectrum. The extent of conversion value was
model [20,21]. inserted into the reaction model, and a plot of g(x) versus x was
In the present work, a series of non-isothermal dehydro- obtained for the various reaction models (Supplementary
genation tests were performed for each Ti1xVxH2 alloy using information, Table S1). By using equation (3) these g(x)
linear heating rates of 2, 5, and 10 K/min. By employing the graphs were fitted under the condition that the activation
KAS relation, Ex versus x during the transformation g energy value (E) should converge towards the values obtained
(hydride) / BCC (metal) was obtained as shown in Fig. 3. It can by the isoconversional method. This means that the activa-
be seen that Ex changed with the extent of conversion. Also tion energy value obtained from the KAS methods will not
from Fig. 1b, it can be deduced that the Ex dependency on x can always be the same as that obtained from the fitting since a
be divided into 3 segments; the 1st segment is for x < x1, the single reaction mechanism was assumed even though two or
2nd segment is for x1 < x < x2, and the 3rd segment is for x > x2. more mechanisms may have operated. The integral temper-
It is worth noticing that the heating rates at the beginning of ature was estimated using an approximation by Gorbachev
experiment were not stable so that reliable Ex was not ach- [22], i.e.
ieved for x < x1. Thus, only Ex values obtained for x > x1 will be
discussed further. ZT    
E RT2 E
As can be seen from Fig. 3, in the 2nd segment, Ex decreased exp dT ¼ exp (5)
RT E þ 2RT RT
0
for Ti0.8V0.2, but fluctuated for Ti0.7V0.3 and Ti0.9V0.1. This
behaviour was different in the 3rd segment within which Ex Fig. 4 shows plots of the reaction models g(x) for different re-
increased as the transformation progressed. Furthermore, the action mechanisms. As shown in Fig. 4, for Ti0.8V0.2Hx the re-
values of x for which the Ex graph changes depend on action mechanism within the 2nd segment was best fitted with
the Avrami-Erofeev nuclei growth model A1, [ln(1  x)]2/3.
However, the numerical value of the activation energy received
segment 1: < 1 Ti0.9V0.1H2 during the fitting, 80 kJ/mol H2, is lower than the average value
175 segment 2: 1< < 2 Ti0.8V0.2H2 obtained by the isoconversional method. Thus, we assume that
segment 3: > 2 simultaneous reactions must be occurring within the 2nd
Ti0.7V0.3H2
segment. In contrast, for Ti0.8V0.2Hx within the 3rd segment of
150 1
the desorption traces, the experimental data fitted well with the
E (kJ/mol)

2 three-dimensional contracting volume model with decelerated


1
125 interface D4 (see Supplementary information, Fig. S2a) with an
activation energy of desorption of 102 kJ/mol H2.
100 2 Fitting of the experimental data for Ti0.7V0.3Hx (not shown)
1
yielded similar reaction mechanisms, i.e. the Avrami-Erofeev
2 nuclei growth model for the 2nd segment and the three-
75 dimensional contracting volume model for 3rd segment. As
also shown in Fig. 4, the 2nd segment for Ti0.9V0.1Hx is well
0.2 0.4 0.6 0.8
described by the three-dimensional contracting volume
model with decelerated interface D4, g(1  2/3x)  (1  x)2/3.
Fig. 3 e Dependency of the activation energy, Ex on the The 3rd segment of Ti0.9V0.1Hx also follows the D4 model (see
extent of conversion for the binary TieV hydrides. Supplementary information, Fig. S2b). However, the fitted Ex
Author's personal copy

14708 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4

with H/M  2 crystallize with an FCC CaF2 type crystal struc-


ture, space groupFm3m, and are called g hydrides. After
completion of the TDS process, the dehydrogenated powders
contain only a BCC phase, except of richest with titanium
Ti0.9V0.1, which contained a mixture of BCC and HCP a-phase,
space group P63/mmc.
Table 1 gives the crystallographic data of the fully hydro-
genated alloys, the g hydrides, and the corresponding dehy-
drogenated alloys. g-Ti0.9V0.1H2 was fitted with a single g-
Ti0.9V0.1H2 phase even though the hydride powder could be
composed of two types of g-(TieV)H2. For Ti0.8V0.2 and Ti0.7V0.3,
it is observed that the unit cell lattice parameter depend on
the vanadium contents: it decreases as the vanadium content
increases. The unit cell volume expansion per dissolved
hydrogen atom (DnH) is equal to the unit cell volume of the
metal hydride subtracted by the unit cell volume of the
metallic phase (same crystal structure) divided by the number
of hydrogen atoms in the unit cell. Since there exist no stable
Fig. 4 e Fitting of the experimental data of segment 2 for FCC TieV alloys, the unit cell volume is calculated based on
Ti0.8V0.2Hx and Ti0.9V0.1Hx using various reaction models. electronic structure calculations by Wolverton [23]. The re-
The heating rate of the TDS experiment was 5 K/min. sults are given in Table 1, which shows that the obtained
A1 [ Avrami-Erofeev one-dimensional nuclei growth; values are close to the typical value observed for the metal
A2 [ Avrami-Erofeev two-dimensional nuclei growth; hydrides (according to Fukai, i.e. 2.8  A3/at.H [24]), which is
A3 [ Avrami-Erofeev three-dimensional nuclei growth; consistent with occupancy of the tetrahedral sites.
CV1 [ One-dimensional contracting volume; CV2 [ Two-
dimensional contracting volume; CV3 [ Three-
3.5. In situ SR-XRD study of Ti0.8V0.2H2
dimensional contracting volume; D1 [ One-dimensional
diffusion controlled; D2 [ Two-dimensional diffusion
A g-Ti0.8V0.2H2 powder sample containing 3.96 wt.% H was
controlled; D3 [ Three-dimensional diffusion controlled;
used in the in situ experiment. The phase evolution during the
D4 [ Three-dimensional contracting volume with
hydrogen desorption from g-Ti0.8V0.2H2 while the sample was
decelerated interface.
heated in vacuum at 3 K/min is shown in Fig. 5. The diffraction
pattern at each time-step was fitted using Rietveld refinement
methods in the GSAS. During the fitting, Rwp and Rp were in
values of 112 kJ/mol H2 within 2nd segment and 118 kJ/mol H2 the ranges of 4.6e11.7% and 3.4e9.7%, respectively. Selected
within the 3rd segment are both higher for the last sample results of the Rietveld refinements are given in Table 2. It can
with the highest Ti content. The activation energy within 3rd be seen that at 40  C the powder was composed of nearly
segment was equal to the value obtained from the iso- 100 wt.% g with a unit cell lattice parameter of 4.4332(5)  A.
conversion method of KAS that showed that a single reaction Diffraction patterns above 320  C were fitted as two-phase
mechanism had been operating. hydride samples, i.e. g FCC and d BCT (base centred tetrag-
onal) structure type, space group I4/mmm. A strong indication
3.4. Phase-structural composition that the sample was composed of a mixture of g and d can be
seen by observing the progress of the overlapping peaks of g-
The alloys Ti1xVx (x > 0.1) have a BCC structure, space (111) and d-(101) from 450  C to 497  C, as shown in the inset in
groupIm3m [1]. The corresponding hydrides of these alloys Fig. 5. Asymmetric profiles were also observed for the (200),

Table 1 e Crystallographic data for the BCC Ti1LxVx alloys and their corresponding hydrides.
x Ti1xVx g-(Ti1xVx)Hw2 DnH (
A3/at.H) Comments

a (
A) V (
A3) 
A V (
A3)

0.1 3.20770(2) 33.005(5) 4.4390(7) 87.47(4) e Two phase alloy


a ¼ 2.94272(2) 35.097(5)
c ¼ 4.6798(4)
0.2 3.2322(2) 33.76(1) 4.4341(5) 87.17(9) 2.81 This work
0.3 3.2113(3) 33.11(6) 4.40168(5) 85.28(9) 2.74 This work
0 a ¼ 2.95111 35.33 4.454 88.35 2.62 [2,25] a
c ¼ 4.68433
a
1 3.03 27.8 4.27 77.5 2.93 [26]
a
Calculated from crystal structure data given in the references.
Author's personal copy

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4 14709

Fig. 5 e Phase evolution during in-situ SR-XRD of g-Ti0.8V0.2H2. The figure inset shows overlapping g-(111) and d-(101) peaks.
Small fraction (w0.5 wt.%) of stable g hydride was observed up to 580  C. The experiment was performed by increasing
temperature from 30  C to 800  C in vacuum.

(220), and (311) g peaks. Such asymmetric profiles are an temperatures the powder was composed of g-Ti0.8V0.2H2 and
indication that the sample has been partly transformed to a this phase was preserved up to 320  C. Above 320  C, some g
tetragonally deformed BCT structure. Some fitted SR-XRD hydride transformed to d-Ti0.8V0.2H1.5. The fraction of g
spectra at high temperatures are given in Supplementary decreased very fast in the temperature range 320  Ce390  C.
information (Fig S3). As a result, the d-Ti0.8V0.2Hw1.5 fraction increased fast, and it
Fig. 6 shows the phase fraction changes derived from the reached a maximum of 55 wt.% at 390  C. The d phase fraction
Rietveld refinements. As can be seen from the figure, at low decreased above 390  C because d readily transformed to the b

Table 2 e Crystallographic data at selected temperatures during the dehydrogenation of g-Ti0.8V0.2H2.


Temperature ( C) Phase-space group Unit cell lattice Abundance (wt.%) Rietveld refinements parameter
parameter ( A)
Rwp (%) Rp (%)
a
40 g-Ti0.8V0.2H2(Fm3m) a ¼ 4.4332(5) 100 6 5.3
310 g-Ti0.8V0.2H2(Fm3m) a ¼ 4.4478(5) 100a 5.8 4.7
360 g-Ti0.8V0.2H1.8(Fm3m) a ¼ 4.415(1) 52(1) 9.1 7.2
d-Ti0.8V0.2H1.4(I4/mmm) a ¼ 3.161(1) 47(1)
c ¼ 4.308(1)
420 g-Ti0.8V0.2H1.6(Fm3m) a ¼ 4.394(1) 36(1) 11.7 9.7
d-Ti0.8V0.2H1.2(I4/mmm) a ¼ 3.152(1) 27(1)
c ¼ 4.282(2)
b-Ti0.8V0.2H0.7(Im3m) a ¼ 3.381(1) 35.8(7)
530 g-Ti0.8V0.2H1.7(Fm3m) a ¼ 4.410(1) 5.9(3) 5.2 3.4
b-Ti0.8V0.2H0.4(Im3m) a ¼ 3.3467(3) 94.07(4)
780 b-Ti0.8V0.2H0.02(Im3m) a ¼ 3.28189(7) 100a 4.6 3.8
800 b-Ti0.8þxV0.2x(Im3m) a ¼ 3.27738(1) 98.8(1) 8.4 6.2
a-Ti(V)-(P63/mmc) a ¼ 2.995(1) 1.1(2)
c ¼ 4.823(3)
a
Phase fraction parameter was not refined during Rietveld refinements.
Author's personal copy

14710 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4

Fig. 6 e Evolution of phase composition calculated from


Fig. 7 e (a) Changes of the g and d hydride lattice
powder diffraction data of Ti0.8V0.2Hx by Rietveld
parameters of Ti0.8V0.2Hx during non-isothermal
refinement method with Rwp [ 3.76e10.28% and
transformation. The d lattice parameters are plotted in a
Rp [ 2.61e7.19%. d-hydride formed from g-hydride at
deformed FCC setting. The g lattice parameter increased up
320  C and existed up to about 500  C. The b hydride
to 320  C and started to decrease rapidly when the
appeared at about 390  C.
d hydride formed at 320  C. Above 450  C. The lattice
parameter of the remaining g hydride slightly increased.

hydride. The amount of b phase increased upon increasing


temperature. In other words, the transformation of g / d and 0.94. The unit cell lattice parameter of the b hydride at 400  C
d / b occurred simultaneously at temperatures above 390  C. was 3.3796(2) 
A. Upon further heating, this unit cell parameter
Thus, there was a three-phase region g þ d þ b in the tem- linearly decreased as hydrogen was desorbed from the
perature range 390  Ce500  C. sample.
The co-existence of three phases during hydrogen trans- Fig. 8 shows changes of the unit cell volume per metal atom
formation has previously been observed in the AB5-H2 system due to hydrogen desorption from the hydride,
[27]. However, it can be observed in Fig. 6 that the direct DVH ¼ VMH  VM =nM , where VMH is the unit cell volume of the
transformation of g / b also occurred at high temperatures hydride, VM is unit cell volume of the metal of the same crystal
when the amount of remaining g hydride was about 10 wt.% or structure, and nM is number of metal atoms in the unit cell. In
less. Above 580  C, the hydride became a single phase b hy- order to account for the thermal expansion effect, the unit cell
dride. At 790  C, small amounts of HCP a-Ti were observed and volume changes of the g and d hydrides were corrected by
reached maximum of 1.1 wt.% at 800  C (not shown in the linear extrapolation of the unit cell volume changes to the
figure). temperatures below 100  C, where hydrogen desorption does
The changes of the unit cell lattice parameters during non- not occur. The effect of temperature on the unit cell volume
isothermal dehydrogenation are shown in Fig. 7. It can be seen change for the b phase was corrected by the difference in b unit
that, when only the single g phase is present, the unit cell cell volume at 850  C and at room temperature. As can be seen
lattice parameter increased with temperature. A decrease of
the g lattice parameter started at 310  C. At 320  C the decrease
in the g lattice parameter was significant and was accompa- 5
nied by a formation of the d hydride. The unit cell lattice
parameter of the remaining g hydride did not change signifi- 4
cantly from 390  C to 450  C. However, above 450  C, the lattice
VH/M (Å3)

parameter of g increased, which indicates that this g hydride 3


must be the Ti-rich g (TieV)Hx.
As can be seen in Fig. 7 where the changes in the lattice 2
parameters of the d hydride are plotted in a distorted FCC
setting (BCT unit cell). As the transformation proceeds, there 1
is an anisotropic contraction of the cubic unit cell, which
0
changes c/a from 1 to 0.96 in a deformed FCC (BCT) cell. Prior to
the formation of b phase, both the a- and the c-axes continu- 0 100 200 300 400 500 600 700 800 900
ously decreased upon the heating indicating that hydrogen Temperature ( C) o
was desorbed as d formed. This is in agreement with the phase
fraction evolution as discussed earlier. However, when the b Fig. 8 e Evolution of unit cell volume change per metal
phase formed, the a-axis of the d hydride increased while the atom due to hydrogen desorption from Ti0.8V0.2Hx. For FCC
c-axis continued to decrease so that c/a finally decreased to TieV, VM is based on the calculated value of 64.67 A3.
Author's personal copy

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4 14711

from Fig. 8, the specific volume of the unit cell decreased lin- while hydrogen atoms diffuse in opposite direction, from the
early during the heating of the sample due to hydrogen bulk of the hydride to the surface of the particles. As the initial
desorption. The linearly fitted lines in Fig. 8 indicate that the hydride has metallic properties, we assume that desorption
rates of volume decrease of the g and d unit cell volumes are takes place from the whole surface area of the powder
equal, i.e. 0.0086 
A3/K, and are higher than that of the b-hy- particles.
dride, being 0.0044 A3/K. A simple mechanism of hydrogen desorption from a
The slopes of the unit cell volume changes of the g and single-phase hydride and a characteristic form of the initial
d phases are similar, which indicates a tetragonal site occu- parts (1st segment) of the thermal desorption spectra makes
pation. The slope of the unit cell volume changes of the b fitting of the experimental data possible. An Arrhenius type of
hydride is approximately half of that for the g and d phases. dependence dx=dt ¼ SC2 k0des expðE=RTÞ with fitting parame-
Considering that the stoichiometric concentration of H in the ters A and E has been used during the fitting of the initial part
dihydrides g and d are approximately 2 times higher than for (red line in Fig. 9) of the TDS graphs for all samples, having
the monohydride b, the lattice volume changes due to a different content of vanadium. The rate of change of hydrogen
desorption of one hydrogen atom must be approximately the content in the samples can be presented as
same, i.e. 0.004 A/K. This is not surprising since we assume
dx
that H atoms fill similar interstitial sites in these structures. ¼ SC2 k0des expðE=RTÞ (6)
dT
Dividing DVH by DnH (the atomic volume of hydrogen given
in Table 1) gives the amount of desorbed hydrogen in unit H/M where
during progress of dehydrogenation progress (not shown). S is the specific surface area of the hydride sample (we
Such calculation was performed using DnH value of 2.82  A3 for assume that it does not change during the experiment),
3
g and d and 2.17 A for b, taken from a previous investigation C is the hydrogen content in the subsurface part of the
[3]. At the beginning of the experiment, the g hydride con- sample,
tained w2 H/M, and this amount decreased upon the heating. k0des is a desorption constant accounting for a shape factor,
The d hydride contained 1.58 H/M when it first formed at E is the activation energy of hydrogen desorption.
320  C, and then this value continuously decreased to 1.15 H/M An analysis of the expression above leads to the conclusion
before the hydride transformed to b. The first formed b (at that C remains practically constant during the first step of
400  C) contained 0.78 H/M, and this value was reduced upon hydrogen desorption when g-hydride starts to decompose. In
further heating. Thus, the Ti0.8V0.2H1.99 transformation upon the last expression for the desorption flow the hydrogen
dehydrogenation as observed by in situ SR-XRD is g- content C represents either “lattice” or “diffusion-active”
Ti0.8V0.2H2 / d-Ti0.8V0.2H1.58 / b-Ti0.8V0.2H0.78 / BCC- hydrogen. The very successful fitting and the fact that
Ti0.8V0.2. “diffusion-active” hydrogen changes its concentration in a
broad interval are in favour of the first option, i.e. that the
hydrogen is present as “lattice” hydrogen. The fitting was
4. Discussion performed for curves obtained with the lowest heating rate, as
in their case influence of the diffusion-driven changes in the
The TDS, DSC/TGA and TPD hydrogen desorption spectra
presented above have shown that hydrogen desorption from
g-Ti1xVxH2 dihydrides is composed of several stages.
Different steps of hydrogen desorption, corresponding to the
phase diagram of the system and in situ SR-XRD data, can be
identified as:

(1) nucleation of d hydride;


(2) nucleation of b hydride;
(3) complete decomposition of d hydride;
(4) complete decomposition of g hydride.

Because hydrogen evacuation rates are different for the SR


XRD and TDS experimental setups, some variation in the
temperature of the events was observed.
From the diagram shown in Fig. 6 we conclude that below
w320  C hydrogen evolution takes place from a single-phase
hydride. The hydride does not change its FCC CaF2 type of
crystal structure; however, the unit cell parameter slightly Fig. 9 e Hydrogen vacuum desorption traces from
increases during the H desorption (see Table 1). This means Ti0.8V0.2H2 at a heating rate of 2 K/min. Red line represents
that vacancies are formed in the lattice when the hydrogen calculated data. The numbers show different steps of the
stoichiometry becomes lower than H/Ti0.8V0.2 ¼ 2.0. Hydrogen desorption process and are used for description in the text.
desorption takes place from the surface of the hydride and is (For interpretation of the references to colour in this figure
related to the increase in concentration of the vacancies and legend, the reader is referred to the web version of this
their diffusion from the surface to the core of the particles article.)
Author's personal copy

14712 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4

subsurface concentration should be maximum, and the con- during the further stages hydrogen is released from the single-
centration C should change within the narrow limits. For the phase b hydride, which is a solid solution of hydrogen in a
alloys Ti0.8V0.2 and Ti0.7V0.3 the heating rate was 2 K/min, and metal alloy. As the temperatures become higher, the hydrogen
for Ti0.9V0.1 is was 5 K/min. A good agreement between the diffusion coefficient increases i and despite of the presence of
fitted and the experimental data is evident from Fig. 9. In the rather large particles of the powdered hydride, it seems that
initial part of hydrogen desorption spectra (1st segment), the hydrogen diffusion is no longer a limiting step of hydrogen
activation energy remains approximately the same for all desorption. Consequently, the rate of hydrogen transfer from
studied alloys, 21 kJ/mol H2. E has a rather small value, in the alpha solid solution into the gaseous phase becomes pre-
agreement with the data of the calorimetric study which dominant, pointing on a strong metal-hydrogen binding in the
showed absence of significant heat effects at working tem- solid solution.
peratures below 300  C (see Supplementary information, Thus, from both TDS and in situ SR-XRD, the path of g-
Fig. S1a). A certain difference in the rate of hydrogen release Ti0.8V0.2H2.0 non-isothermal dehydrogenation can be proposed
from the alloys containing different amounts of vanadium is as follows: g / d / b / balloy, noting that the transformations
due to the differences in the values of A, which again result g / d and d / b were simultaneously occurring within the
from variations in the surface area, S. material. The direct transformation g / b also happened at
An increase in temperature leads an increase the desorp- high temperatures above 500  C when the amount of g hydride
tion rate; at some moment, which decreases the content of was about 10 wt.%. This direct transformation g / b origi-
hydrogen in the subsurface layers, as the amount of hydrogen nates from the Ti-rich phase having the higher thermal sta-
coming via diffusion from the bulk does not compensate for bility than the major (Ti,V)-based gamma hydride.
the desorbed H2. This leads to the bending of the TDS curve at Simultaneous decomposition reactions were occurring
260  C. When the hydrogen content drops below the critical during hydrogen desorption. This is in agreement with in situ
value at 320  C (point 1 in the TDS traces), a phase transition SR-XRD results that showed that at intermediate tempera-
takes place. After nucleation of the d-hydride, during a modest tures, reactions g / d and d / b took place in parallel during
temperature increase from 320  C to 390  C, this hydride be- the transformation of g-Ti0.8V0.2H2. At higher temperatures,
comes a major component in the sample as its abundance interaction between the transformations d / b and b / balloy
surpasses 55%. Thus, we assume that a thick shell of d-hydride also took place during the dehydrogenation of the b-
is formed which grows inside the particle leading to the Ti0.8V0.2H0.7. This indicates that, most likely the intermediate
contraction of the central spherical core of the g-hydride. As d- phase d was participating in the transformation of g to b hy-
hydride is rather rich in hydrogen, a contribution of the ged dride of Ti0.9V0.1Hx. For TiH2 a direct transformation of g to b
transformation to the overall desorption is rather limited. has been shown.
Nucleation of the b-hydride taking place at 390  C (point 2 Based on the KAS method presented in the previous sec-
in the TDS traces) correlates with the maximum rates of tion, the activation energy for the d / b transformation was
hydrogen desorption; thus, the hydrogen flow rate signifi- 79 kJ/mol H2 for Ti0.8V0.2Hx and 60 kJ/mol H2, i.e. the slightly
cantly increases. According to the DSC data (Supplementary lower, for Ti0.7V0.3Hx. During the dynamic transformation
information, Fig. S1a), hydrogen desorption at such a tem- d / b another reaction also proceeded, i.e. dehydrogenation of
perature is accompanied by a significant heat effect. This the b hydride. The transformation g / b in the case of Ti0.9V0.1
observation well agrees with a stronger metal-hydrogen bond was fitted to the three-dimensional contracting volume model
in the b-hydride compared to in the g-hydride. Decreasing with a decelerated interface. The g / b transformation during
abundance of the d-hydride (taking place at temperatures dehydrogenation of Ti0.9V0.1H2 followed the TiH2 trans-
between 420 and 500  C) is accompanied by a decrease in the formation during which a direct transformation of g to b was
hydrogen flow rate. observed [28]. The activation energy of the transformation
From Fig. 6 it is evident that both g and d hydrides are in g / b was about 112 kJ/mol H2. The hydrogen desorption
equilibrium with b-hydride in this three phase region. In other mechanism from the g hydride of Ti0.9V0.1 is rather compa-
words, in this region there are simultaneous transformations rable to the transformation of FCC to BCC hydrides of titanium
g / d and d / b. It is worth mentioning that a similar which is limited by the interface reaction [29]. The activation
behaviour has also been observed in the AB5eH2 system [14]. energy of the FCC / BCC TiH2 transformation is in the range
The DSC traces (Supplementary information, Fig. S1a), how- of 123e137 kJ/mol H2 [29,30] which is significantly higher than
ever, show that the heat effect of the transformations reaches that of Ti0.9V0.1H2 due to the vanadium alloying in the latter.
maximum in the temperature interval 450e500  C, which in- The main difference is that for the in Ti0.9V0.1 hydride the
dicates that the energy level of hydrogen is higher in the d- interface was decelerated upon the progress of dehydroge-
phase than in the b-hydride. nation. The reason for a decelerated interface must be a
At the moment when the d-hydride vanishes (w500  Cdpoint stronger hydrogen binding of hydrogen to Ti, or slow diffusion
3), the remaining amount of g-hydride is rather low (see Fig. 6). of hydrogen within the transformed particle. Since the
However, decomposition of the remaining g hydride in the hydrogen diffusion within b hydride is fast at high tempera-
temperature range 500 < T < 580  C keeps the hydrogen tures, the first reason is probably the most reasonable one.
desorption flow at a rather high level. We believe that H2 The high stability of the g hydride up to 580  C, is proposed
desorption at these temperatures takes place from the surface of to be associated with the presence of a Ti-rich TieV hydride,
the b-hydride, and that g-hydride exists as a contracting area i.e. Ti0.8þxV0.2xHy. As mentioned earlier, Ti and V form a
located inside the b-hydride. At T > 580  C (point 4) no further disordered BCC solid solution. Furthermore, TieV alloys have
hydrogen is available from the hydrogen-rich g-hydride, and a positive energy of mixing, so a miscibility gap exists in the
Author's personal copy

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4 14713

phase diagram of the TieV system. Ti1xVx alloys tend to form


clusters and other related types of short range ordering. The Acknowledgements
occurrence of an a-Ti phase at about 790  C during the in situ
experiment therefore most likely originates from a Ti-rich The authors thank the Norwegian Research Council Norway
TieVeH phase. This separation of a Ti phase from the main and Statoil for their financial support. We appreciate a skilful
BCC hydride phase was not observed in the dehydrogenated assistance from the personnel of Swiss-Norwegian Beam
samples (ex-situ studies) of Ti0.7V0.3. This behaviour can also Lines during the SR XRD experiments.
be explained from the equilibrium TieV phase diagram. The
balloy / a þ balloy transformation temperature decreases with
increasing vanadium fraction. As a result, at lower vanadium Nomenclature
fractions, the temperature range for the a-Ti phase formation
is broad, so the driving force for the formation is also high. No
a-Ti (HCP) was observed in the dehydrogenated Ti0.7V0.3 alloys. E activation energy, kJ/mol H2
Therefore, it is evident that formation of a-Ti from the BCC A pre-exponential factor, 1/min
phase will be dominant in low vanadium alloys. x extent of conversion
The existence of the Ti-rich TieVeH hydrides mentioned b heating rate, K/min
above is also in agreement with the TDS result shown in Fig. 1, k kinetics constant, 1/min
in which one can observe that the stability of the TieV hydride C hydrogen content in the surface
depends on the alloy composition, i.e. atomic ratio between Ti S sample surface of hydride particle, cm2/g
and V. Thus, there must be an inhomogeneity in the sample of VH unit cell volume of metal hydride, 
A3
g-Ti0.8V0.2H2.0.. This kind of inhomogeneity in the TieVeH2 g FCC hydride sp. gr. Fm3m
system was suggested to be a result of a high-pressure syn- d BCT hydride sp. gr. I4/mmm
thesis and also of hydrogenationedehydrogenation cycling. b BCC hydride sp. gr. Im3m
Since our sample of g-Ti0.8V0.2H2.0 was prepared using 20 bar
hydrogenation pressure, such inhomogeneity is rather likely.

Appendix A. Supplementary data


5. Conclusions
Supplementary data related to this article can be found online
at http://dx.doi.org/10.1016/j.ijhydene.2013.08.103.
Hydrogen absorptionedesorption properties of Ti1xVxH2
have been studied using thermal desorption spectroscopy and
in situ synchrotron X-ray diffraction studies. The following
conclusions can be drawn:
references

(a) The Ti1xVx alloys form a saturated g-Ti1xVxH2 dihydride


[1] Predel B. Ti-V (Titanium-vanadium). In: Madelung O, editor.
(sp.gr.Fm3m) upon hydrogenation. The changes in the
Springer Materials e the Landolt-Börnstein database (http://
lattice parameters of the dihydrides correlate with the
www.springermaterials.com). http://dx.doi.org/10.1007/
trends observed for the lattice parameters of the initial 10551312_2842.
alloys. [2] San-Martin A, Manchester F. The HeTi (hydrogenetitanium)
(b) Isoconversional kinetics studies indicated that dynamic system. J Phase Equilib 1987;8:30e42.
dehydrogenation of Ti1xVxH2 is composed of simulta- [3] Hagi T, Sato Y, Yasuda M, Tanaka K. Structure and phase
neous reactions that are in agreement with TDS and in situ diagram of TieVeH system at room temperature. Mater
Trans 1987;28:198e204.
SR XRD studies.
[4] Hayashi S, Hayamizu K. X-ray diffraction and 1H and 51V
(c) The main dehydrogenation path of g-Ti1xVxH2 is NMR study of the TieVeH system. J Less-Common Met
g / (d) / b / balloy. The BCT hydride d (sp.gr. I4/mmm), 1990;161:61e75.
has been observed in hydride phases containing 20 and [5] Ueda T, Hayashi S, Hayamizu K. Hydrogen motion and local
30 at.% V. For 10 at.% V, transformation of g hydride to b structure of metals in Ti1yVyHx as studied by 1H NMR. Phys
hydride follows a direct g / b route, which is similar to the Rev B 1993;48:5837e43.
transformation of TiH2, i.e. without formation of d hydride [6] Brouwer R, Rector J, Koeman N, Griessen R. Hydrogen as a
local probe: diffusion and short-range order in Ti1yVy alloys.
phase as an intermediate step.
Phys Rev B 1989;40:3546e59.
(d) A simple model of hydrogen desorption from single phase [7] Sevilla EH, Cotts RM. Hydrogen diffusion in bcc TiHx and
hydrides nicely describes its release from g-Ti1xVxH2. Ti1yVyHx. Phys Rev B 1988;37:6813e20.
This allowed accurate determination of the activation [8] Nowak B, Hayashi S, Hayamizu K, Yamamoto O. X-ray
energy of hydrogen desorption for all the investigated diffraction and 1H and 51V NMR study of the structure of a
TieV hydrides. TieVeH alloy in relation to preparation conditions. J Less-
Common Met 1986;123:75e84.
(e) At high vanadium content, 20e30 at.% V, the kinetics of the
[9] Ono S, Nomura K, Ikeda Y. The reaction of hydrogen with
transformation of d to b hydride was well fitted by a one
alloys of vanadium and titanium. J Less-Common Met
dimensional Avrami-Erofeev nucleation and growth model, 1980;72:159e65.
while at high titanium/vanadium ratios hydride decompo- [10] Suwarno S, Gosselin Y, Solberg JK, Maehlen JP, Williams M,
sition is affected by formation of a Ti-rich hydride. Krogh B, et al. Selective hydrogen absorption from gaseous
Author's personal copy

14714 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 7 0 4 e1 4 7 1 4

mixtures by BCC TieV alloys. Int J Hydrogen Energy [21] Ortega A. The kinetics of solid-state reactions toward
2012;37:4127e38. consensus, part 2: fitting kinetics data in dynamic
[11] Denys RV, Riabov AB, Maehlen JP, Lototsky MV, Solberg JK, conventional thermal analysis. Int J Chem Kinet
Yartys VA. In situ synchrotron X-ray diffraction studies of 2002;34:193e208.
hydrogen desorption and absorption properties of Mg and [22] Gorbachev VM. A solution of the exponential integral in the
MgeMmeNi after reactive ball milling in hydrogen. Acta non-isothermal kinetics for linear heating. J Therm Anal
Mater 2009;57:3989e4000. 1975;8:349e50.
[12] Larson AC, Dreele RBV. Los Alamos National Laboratory [23] Wolverton C, Ceder G, de Fontaine D, Dreyssé H. Ab initio
Report LAUR; 2004. determination of structural stability in fcc-based transition-
[13] Rietveld H. A profile refinement method for nuclear and metal alloys. Phys Rev B 1993;48:726e47.
magnetic structures. J Appl Crystallogr 1969;2:65e71. [24] Fukai Y. Site preference of interstitial hydrogen in metals. J
[14] Howard C. The approximation of asymmetric neutron Less-Common Met 1984;101:1e16.
powder diffraction peaks by sums of Gaussians. J Appl [25] Yakel Jr H. Thermocrystallography of higher hydrides of
Crystallogr 1982;15:615e20. titanium and zirconium. Acta Crystallogr 1958;11:46e51.
[15] Thompson P, Cox DE, Hastings JB. Rietveld refinement of [26] Müller H, Weymann K. Investigation of the ternary systems
DebyeeScherrer synchrotron X-ray data from Al2O3. J Appl NbeVeH and TaeVeH. J Less-Common Met
Crystallogr 1987;20:79e83. 1986;119:115e26.
[16] Vyazovkin S, Burnham AK, Criado JM, Pérez-Maqueda LA, [27] Riabov AB, Denys RV, Maehlen JP, Yartys VA. Synchrotron
Popescu C, Sbirrazzuoli N. ICTAC Kinetics Committee diffraction studies and thermodynamics of hydrogen
recommendations for performing kinetic computations on absorption-desorption processes in La0.5Ce0.5Ni4Co. J Alloy
thermal analysis data. Thermochim Acta 2011;520:1e19. Compd 2011;509:S844e8.
[17] Vyazovkin S. Isoconversional kinetics. In: Michael EB, [28] Borchers C, Khomenko TI, Leonov AV, Morozova OS.
Patrick KG, editors. Handbook of thermal analysis and Interrupted thermal desorption of TiH2. Thermochim Acta
calorimetry. Elsevier Science B.V.; 2008. p. 503e38. 2009;493:80e4.
[18] Kissinger HE. Reaction kinetics in differential thermal [29] Liu H, He P, Feng JC, Cao J. Kinetic study on nonisothermal
analysis. Anal Chem 1957;29:1702e6. dehydrogenation of TiH2 powders. Int J Hydrogen Energy
[19] Akahira T, Sunose T. Method of determining activation 2009;34:3018e25.
deterioration constant of electrical insulating materials. [30] Evard EA, Gabis IE, Voyt AP. Study of the kinetics of hydrogen
Research report. Chiba Inst Technol Sci 1971;16:22e31. sorption and desorption from titanium. J Alloy Compd
[20] Vyazovkin S, Wight CA. Model-free and model-fitting 2005;404e406:335e8.
approaches to kinetic analysis of isothermal and
nonisothermal data. Thermochim Acta 1999;340e341:53e68.

Vous aimerez peut-être aussi