You are on page 1of 12

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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0

Available online at www.sciencedirect.com

ScienceDirect

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

High temperature hydrogenation of TieV alloys:


The effect of cycling and carbon monoxide on the
bulk and surface properties

S. Suwarno a,*, J.K. Solberg a, B. Krogh c, S. Raaen 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, NTNU, NO-7491, Trondheim, Norway

article info abstract

Article history: In the present work, the high temperature (425e575  C) hydrogen storage properties of
Received 4 September 2015 TieV alloys have been studied, both at static conditions and in a flow of hydrogen gas.
Received in revised form The selected isothermal temperature range is considered as an optimal condition for
10 November 2015 hydrogen sorption enhanced steam reforming. When hydrogenation and dehydrogena-
Accepted 10 November 2015 tion were performed in pure hydrogen gas and in vacuum, a large reversible hydrogen
Available online 24 December 2015 capacity of 3.95 wt. % H was obtained, demonstrating completeness of the formation and
decomposition of (Ti,V)-dihydrides. However, when cycling was performed in a flow of
Keywords: hydrogen gas, the reversible hydrogen capacity decreased to ~2 wt. % H caused by the
Titaniumevanadium alloys formation of stable lower (Ti,V)-hydrides. A further decrease in the reversible hydrogen
Metal hydrides storage capacity took place when pure hydrogen gas was replaced by a mixture of
High-temperature hydrogen storage hydrogen and carbon monoxide CO. This decrease was caused by the formation of an
Hydrogen production oxygen rich layer on the surface of the alloy, which was partially blocking the hydrogen
Carbon monoxide exchange between the surface and the bulk of the sample.
Copyright © 2015, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.

utilisation [1]. Today hydrogen is mostly produced by


Introduction reforming of hydrocarbon feedstocks via a steam reforming
process. The steam reforming is an established process for
Hydrogen is an important industrial gas that is used mostly hydrogen production at large industrial scale. Nevertheless,
for ammonia synthesis and refining oil products. Other in- further improvements are needed, particularly in the effi-
dustrial processes such as production of metals and food ciency of the reforming and in decreasing the emissions.
preparation also use large amounts of hydrogen. An emerging Several new technologies are now widely studied as having
application is hydrogen as a fuel for the FC hydrogen vehicles, a potential of improvement of the conventional steam
because of its high energy density and zero emission reforming process [2,3]. The working temperature of steam

* Corresponding author. Present address. Department of Mechanical Engineering, ITS, Kampus ITS, Keputih Sukolilo Surabaya, 60111,
Indonesia. Tel./fax: þ62 315946230.
E-mail addresses: warno@me.its.ac.id, warnoise@yahoo.com (S. Suwarno).
http://dx.doi.org/10.1016/j.ijhydene.2015.11.077
0360-3199/Copyright © 2015, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
1700 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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0

reforming can be reduced while maintaining or increasing the amount of CO. However, significant degradation was observed
hydrogen yield by removing one or more of the reaction leading to a decreased reversible hydrogen capacity during the
products. Such a strategy is employed for example by allowing cycling. The purpose of the present study was to investigate in
hydrogen to pass a membrane [4]; another approach is an in detail microstructural changes in the bulk and surface prop-
situ absorption of the reaction products by absorbent mate- erties of the Ti0.8-0.9V0.2-0.1 H storage alloys during hydrogen
rials. In such a way, by absorbing CO2 as one of the products by absorption-desorption cycling at high temperatures. The goal
absorbent materials, e.g. CaO [5] or LieZr mixed oxides [6,7], was to understand the controlling factor for degradation. The
achievement of a higher degree of hydrocarbon conversion is effects of pressure, temperature and gas composition on the
anticipated. performance of the alloys as well as the kinetics of the
One recent approach is hydrogen sorption enhanced steam chemical transformations were studied.
reforming (HSER) [8]. Thermodynamic modeling confirmed
that HSER allows increasing energy efficiency and reducing
the operation temperatures. HSER utilizes metal hydride
Materials and experimental methods
forming alloys that absorb hydrogen and form hydrides. In the
HSER process, the properties of the hydrogen sorbent are
Synthesis of alloys and hydrides
crucial for its realization. Ideal metal hydrides reversibly store
large amounts of hydrogen and maintain unchanged the
Alloy buttons with compositions Ti0.8V0.2 and Ti0.9V0.1 were
reversible hydrogen capacity during long term cycling. An
prepared from titanium sponge (99.7%) and vanadium turn-
important issue is that, in the HSER reactors, the metal hy-
ings (99.7%) using an Edmund Bühler arc melter. During the
dride should be able to form even at very harsh operating
melting, zirconium was used as a getter material to reduce
conditions. Hydrogen should be absorbed from a mixture of
oxygen contamination of the prepared alloy buttons. The al-
the reactants/products containing significant amounts of the
loys were remelted three times to ensure their homogeneity.
chemically reactive gases, CO, CO2 and H2O. Properly
Activation of the alloys and synthesis of their hydrides
addressing the effect of these impurity gases is very chal-
were performed following the procedures described in our
lenging as it is well-known that they suppress hydrogen ex-
earlier publication [17]. After the hydrogenation, the initial as
change in hydrogen storage alloys [9e11].
cast millimeter-size particles were pulverized to the powders
Another requirement is that the hydride must be formed/
of the dihydrides (Ti,V)H2 with average particles size of
decomposed at rather high operating temperatures,
200 mm.
450e800  C. One of a potentially promising from this view-
point systems is TieH system. The exceptionally high volu-
metric density of hydrogen in TiH2, being 150 kg/m3, is more Hydrogenation and dehydrogenation experiments
than two times higher than in liquid hydrogen. The high for-
mation enthalpy of TiH2 of 123.4 kJ/mol H2 at 298 K [12] Hydrogen absorption-desorption experiments were per-
causes a significant heat release during the hydrogenation formed in a Sievert's volumetric set-up. Hydrogen absorption
process. This makes titanium and its alloys promising candi- was done at quasi-isothermal conditions while dehydroge-
dates for applications in hydrogen enhanced steam reform- nation was performed non-isothermally by heating the sam-
ing. However the hydride formation enthalpy and the kinetics ple from the hydrogenation temperatures to 800e850  C with a
of hydrogen exchange of the TiH2 must be tuned to obtain rate of 5 deg/min, with a starting pressure of appr.
favorable properties suitable at HSER working temperature. ~1  105 mbar. Hydrogenation-dehydrogenation experi-
One of the alloying elements to combine with titanium is ments were also performed in a flow of gas using the
vanadium. Interaction of hydrogen with elemental vanadium temperature-programmed desorption (TPD) equipment by
is rather complex and leads to the formation of several hy- Altamira Instrument Inc. The TPD equipment consisted of two
drides with different H/V ratios reaching 2.0 at maximum. parts that could be used for the experiments performed both
Stable vanadium hydrides are formed at temperatures below at atmospheric conditions and at high pressures. A Thermo-
appr. 200  C. The formation enthalpy of FCC vanadium dihy- star Mass Spectrometer (MS) from Pfeiffer Vacuum GmBH was
dride VH2 is much lower than that of TiH2, only 40 kJ/mol H2 connected to the TPR outgoing gas flow for measuring the
[13,14]. As the equilibrium pressure of 1 bar H2 is reached at a composition of the outgoing gases. The MS was calibrated to
rather modest temperature, ~20  C, vanadium dihydride is determine the contents of H2, Ar, CH4, CO, CO2, O2 and H2O.
classified as a low temperature hydride [15]. Alloying titanium The experiment control and data acquisition were conducted
with vanadium results in the formation of solid solutions at all using the LabView based software program developed by the
Ti/V ratios, which in turn form dihydrides (Ti,V)H2. An in- provider of the equipment. Two kinds of the experiments
crease in the vanadium content decreases the enthalpy of were conducted in the gaseous flow; the first was hydroge-
formation and thus allows adjustment of the stability of the nation and dehydrogenation cycling at 1 bar. The hydroge-
hydrides and, consequently, optimization of the working nation experiments were performed at various temperatures
temperature required for a particular application. (425, 475, 525 and 575  C) in a mixture Ar þ 10% H2 using a rate
In a previous publication [16], hydrogen sorption properties of gas flow of 50 ml/min while the dehydrogenation were done
and in situ synchrotron X-ray diffraction study of TieV alloys in pure Ar at 1 bar. The cycling experiments of second type
interacting with hydrogen gas mixed with CO were reported. were done in flows of pressurized gaseous mixtures Ar þ 25%
The study [16] showed that hydrogenation-dehydrogenation H2 and Ar þ 25%H2 þ 1%CO. The hydrogenation was per-
occurred even when the gas mixture contained a substantial formed at 425  C at a total pressure of 20 bar while the
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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0 1701

dehydrogenation was done by heating the sample in pure Ar


to 760  C at the same pressure.

Materials characterizations

A Zeiss Ultra 55LE field emission scanning electron micro-


scope (FE-SEM) equipped with a Bruker energy dispersive
spectrometry (EDS) was used for the studies of the micro-
structure, morphology and chemical analyses. In addition,
some chemical micro analyses were performed using a JEOL
JXA-8500F electron probe micro-analyzer (EPMA). Samples for
the SEM investigations were prepared by dispersing small
amounts of the powders on a carbon tape attached to the
sample holder. To examine the powder particle cross-sections
in the EPMA, powder particles were embedded into a carbon
filled resin, grinded and mechanically polished down to 1 mm.
During the SEM investigations, the surface morphology of the
powders was investigated using secondary electrons at a
typical accelerating voltage of 5 kV. EDS area analyses were
performed at 15e20 kV accelerating voltage and at 200 s
acquisition time. EPMA point analyses were done at 15 kV
accelerating voltage and 10 s acquisition time.
Laboratory XRD was performed using a Bruker D8 Focus
diffractometer with a Cu Ka X-ray source. A crushed powder
sample was dispersed on a single crystal silicon sample holder
using ethanol as dispersing agent. Powder diffraction patterns
were collected using a square slit of 0.6e1 mm side length. The
data was collected in a 2q angle range from 10 to 90 deg. using
a step size of 0.01 deg., while the sample holder was contin-
uously rotated. Some diffraction patterns were obtained using
synchrotron X-ray diffraction (SR-XRD) at the Swiss Norwe-
gian Beam Lines (BM01A, SNBL), Grenoble, France.

Surface characterization: AES and XPS Fig. 1 e Quasi-isothermal hydrogenation of Ti0.9V0.1 at


dynamic pressure conditions using volumetric methods:
Auger electron spectroscopy (AES) analyses were performed (a) Ti0.9V0.1; (b) Ti0.8V0.2. Hydrogen absorption experiments
in a JAMP 9500F field emission Auger electron microscope. in pure H2 were performed at 450  C and 550  C and at 2
Some amounts of the powders were dispersed on the and 4 bar initial hydrogen pressures. Dehydrogenation
aluminum foils attached to the AES microscope sample was done in vacuum by heating the sample from the
holder. The specimens were analyzed applying 10 kV accel- hydrogenation temperature to 800e850  C.
erating voltage. Ion sputtering was employed to analyze sub-
surface compositions. X-ray photoelectron spectroscopy
(XPS) measurements were done using a monochromatized Al relative rates of hydrogen uptake and reaching the saturation
Ka (hn ¼ 1487 eV) X-ray source (Gammadata-Scienta) and a hydrogen content in the formed hydrides appear to be
SES2002 electron energy analyzer. Ti2p, V2p, and O1s spectra affected by a) Ti/V ratio; b) temperature of interaction; c) hy-
were recorded. drogenation pressure. The hydrogenation rates increase with
the Ti content of the alloys (being the highest for Ti0.9V0.1) and
with decreasing interaction temperature (being the highest at
Results and discussion 450  C). In contrast, increasing the hydrogenation pressure
from 2 bar to 4 bar caused only a marginal increase in the
Hydrogen absorption properties amount of absorbed hydrogen and in the rates of
hydrogenation.
Hydrogenation at static conditions
When the driving force for the hydride formation was high, at Hydrogenation properties in a flow of hydrogen gas
30  C and at hydrogen pressure 4 bar H2, both Ti0.9V0.1 and In order to simulate the conditions applied during the steam
Ti0.8V0.2 alloys instantly, in less than 1 min, formed dihydrides reforming process, the hydrogenation properties of the sam-
containing up to 3.95 wt. % H (see Figs. 1, a and b). ples were studied in a gas flow. This allowed performing ki-
At higher interaction temperatures of 450  C and 550  C, the netics studies of hydrogen absorption at close to isothermal
process of hydrogen absorption slows down; however, for conditions. Small amounts of the completely hydrogenated
both alloys it still completes rather fast, in less than 5 min. The samples, around 20 mg each, were used in the measurements
1702 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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0

runs to achieve stable, close to isothermal, conditions. The reversible hydrogenation took place at the end of the 4th cycle
experiments were performed using the TPD equipment. The and led to formation of g-(TiV)H1.45 with a body centered
hydrogenation experiments were performed at various tem- tetragonal unit cell (distorted face centered cubic lattice).
peratures (425, 475, 525 and 575  C) in a mixture Ar þ 10% H2 During the consecutive four hydrogen desorption experi-
(1 bar) using a rate of gas flow of 50 ml/min. ments, hydrogen release took place from the partially hydro-
As expected, the hydrogen saturation content of the genated samples obtained during hydrogenation at 425  C
Ti0.9V0.1 alloy depends on the processing temperature (see (des. 2), 475  C (des. 3), 525  C (des. 4), and 575  C (des. 5) (see
Fig. 2a). The highest hydrogen content was reached when the Fig. 2b). Regular changes in the desorption traces are observed
hydrogenation was performed at the lowest temperature of and show a shift of the desorption peaks to higher tempera-
425  C. A progressive decrease of H content took place tures when the temperature of the synthesis increases.
following a temperature increase up to 575  C. Furthermore, a gradual transformation of the double-peak
After the hydrogenation process was completed, the desorption traces into a single-event decomposition during
formed hydrides were cooled to room temperature and des.5 is observed. This indicates that the initial hydride syn-
dehydrogenated by heating the samples from RT to 800  C at a thesized at room temperature was g-dihydride Ti0.9V0.1H2,
rate of 5 deg/min. The hydrogen desorption spectra acquired while the hydride synthesized at the highest synthesis tem-
during the heating are shown in Fig. 2b. During the 1st perature of 575  C was b-hydride Ti0.9V0.1H1-1.5, and the in-
desorption, the fully hydrogenated sample desorbed 3.7 wt.% termediate samples (des.2-des.4) were the samples with
H, which indicates a complete decomposition of the starting gradually decreasing content of g-dihydride and, corre-
g-dihydride. spondingly, increasing content of b-hydride.
Non-isothermal cycling of hydrogen absorption-desorption Similarly, also for the Ti0.8V0.2 alloy, the reversible
in a gas flow mixture of H2 þ CO has been reported by the hydrogen absorption behavior was found to be temperature-
present authors earlier [16]. The experiments showed that depended. Dehydrogenation spectra of the Ti0.8V0.2 hydrides
showed that only high temperature peaks associated with
decomposition of b-hydrides were observed. In all, the
hydrogen capacity of Ti0.8V0.2 was found to be lower than that
of Ti0.9V0.1 since only b-hydrides were formed in the former
alloy at the applied experimental conditions (see supple-
mentary information, Fig. S1, for further details).

Hydrogenation kinetics
Since the hydrogenation experiments were performed at a
constant hydrogen partial pressure of 0.1 bar H2, the ther-
modynamic driving force remained constant, and the kinetics
of the hydrogen absorption can be expressed as

gðxÞ ¼ KT $dt (1)

where gðxÞ is a dimensionless quantity being characteristic of


the particular selected kinetics model, x is a normalized extent
of conversion, and KT is a temperature-dependent kinetics
constant. Consequently, a linear time-dependent gðxÞ is
anticipated.
The experimental data were fitted using the common ap-
proaches described in the reference data (Supplementary in-
formation, Table S1). Following the described procedure, the
kinetic hydrogenation data at 425  C can be fitted using two
related nuclei-growth models, a three-dimensional Avrami-
Erofeev model for Ti0.9V0.1 and a two-dimensional Avrami-
Erofeev model for Ti0.8V0.2 (Supplementary Information Fig. S2).
By using above presented linear relationship for gðxÞ, the KT
value for each hydrogenation temperature was obtained. The
activation energies of hydrogenation were then calculated
from plots of KT vs. 1/T (Supplementary Information Fig. S3).
In the case of Ti0.8V0.2, the activation energy was rather low,
17 kJ/mol, being comparable with the activation energy of
Fig. 2 e Ti0.9V0.1 hydrogenation and dehydrogenation in a hydrogenation for LaNi5, 16e37 kJ/mol [18,19]. For Ti0.9V0.1, the
gas flow of 1 bar at 50 ml/min: a) isothermal hydrogen activation energy was found to be slightly higher, 35 kJ/mol.
absorption in a flow of 10%H2 þ Ar, b) non-isothermal This value is similar to the activation energy for hydrogena-
dehydrogenation during heating from RT to 800  C at 5 K/ tion of beta titanium, 38e45 kJ/mol [20]. One reason why the
min in a pure Ar flow after hydrogenation at temperatures activation energy was changed from Ti0.9V0.1 to Ti0.8V0.2 may
given in the parentheses. be in the difference of growth dimensionality.
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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0 1703

Cycle life and stability of the hydrogen absorption-


desorption performance

For both Ti0.9V0.1 and Ti0.8V0.2, cycling experiments performed


at conditions giving complete hydrogenation and dehydroge-
nation, resulted in high values of the reversible hydrogen
storage capacity, which remained constant for nine cycles
without any degradation (see supplementary information,
Fig. S4).
The second cycling test was done in flows of pressurized
gaseous mixtures Ar þ 25%H2 and Ar þ 25%H2 þ 1%CO. The
hydrogenation was performed at 425  C at a total pressure of
20 bar, while dehydrogenation was done by heating the
sample in pure Ar to 760  C. Fig. 3 presents changes in the
amounts of hydrogen desorbed during the cycling of Ti0.8V0.2
and Ti0.9V0.1.
For both alloys, the reversible hydrogen storage capacity Fig. 4 e Reversible hydrogen storage capacity of Ti0.9V0.1
decreased during the cycling in Ar þ 25%H2. While the initial during hydrogenation and dehydrogenation cycling in the
Ti0.9V0.l based hydride released about 3.7 wt.% H, this value temperature range 150e780  C, 5 deg/min cooling and
was reduced to 2.9 wt.% H during the second cycle, and it heating rate. Hydrogenation was done at different
decreased further to 2 wt.% H during the 11th cycle. The hydrogen pressures, and dehydrogenation was done in
reversible H capacity of the initial sample of Ti0.8V0.2H2.0, pure Ar, 100 ml/min gas flow. CO was introduced into the
3.8 wt.% H, was even stronger affected by cycling. During the gas flow before cycle 16 and removed after cycle 18.
second cycle this sample desorbed only 1.7 wt.% H, and after
the 5th cycle the amount of desorbed hydrogen continuously
decreased during the cycling. The third series of the cycling tests was performed at var-
As can be seen from Fig. 3, the reversible H storage capacity iable hydrogen pressures and at a higher content of CO in the
of Ti0.9V0.1 decreases faster than that of Ti0.8V0.2. This is gas stream. Fig. 4 shows the cycle dependence of the hydrogen
probably a consequence of a higher stability of the Ti0.9V0.1 desorption capacity during the experiments performed in a
hydride so that the amount of “trapped” hydrogen in Ti0.9V0.1 mixture of Ar þ H2. As can be seen from the figure, the effect of
becomes higher than in Ti0.8V0.2. It appears that the amount of pressure was marginal. The reversible hydrogen storage ca-
un-desorbed hydrogen accumulatively increased during each pacity was around 3 wt.% H, and was only slightly dependent
consecutive cycle so that a continuous decrease in reversibly on the pressure changes. When 2% CO was introduced into the
released hydrogen took place during the experiments. gas mixture at the 16th cycle, the hydrogen absorption and
The detrimental effect of CO on the reversibility of the H desorption significantly decreased to just ~1 wt% H. However,
storage capacity is evident from the data presented in Fig. 3. when CO was removed from the gas flow, 80% (2.3 wt.% H) of
The alloy only absorbed about 0.5 wt. % H in presence of CO. the initial H storage capacity was recovered.

Phase-structural studies

The data of the phase-structural analyses of the samples was


related to the testing conditions in order to get insight into the
reasons of the variations of their cycling performance. The
data for the dehydrogenated samples after the 1st and the 8th
cycle performed in pure H2 are compared in Table 1. The
samples contained a mixture of the BCC solid solutions (Ti,V)
and HCP type alloys based on b-Ti. The ratio between the BCC
and HCP compounds appeared to be related to the test history
of the samples.
After the 1st cycle, the two-phase Ti0.9V0.1 was found to
contain HCP a-phase as the main phase and the BCC phase as
the secondary phase. Interestingly, the fraction of the BCC
phase had increased to 48 wt.% after the 8th cycle.
The situation was quite different for the Ti0.8V0.2 alloy. In
Fig. 3 e Amount of hydrogen desorbed in each cycle for this alloy, the BCC phase constituent was initially the vast
TieV alloys hydrogenated at 15e20 bar Ar þ 25%H2 or majority, and it remained so upon cycling (94e99%), while the
Ar þ 25%H2 þ 1%CO, flow rate 300 ml/min. Absorption was amount of the HCP a-phase slightly increased from 1 wt.%
done isothermally at 425  C and desorption was performed after the 1st cycle to 3 wt.% after the 8th cycle. A small amount
non-isothermally from 425 to 760  C, heating rate 5 K/min. of rutile related oxygen-hydrogen containing compound
1704 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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0

Table 1 e Phase-structural data for dehydrogenated samples after 1st and 8th cycle.
Sample Dehydro- Phase Phase content (wt.%) Lattice parameter (
A) Unit cell volume (
A3)
genation cycle#
Ti0.9V0.1 1st b  ðTiVÞðIm3mÞ 33 (5) a ¼ 3.20770(2) 33.005(5)
a  TiðVÞðP63 =mmcÞ 67 (3) a ¼ 2.94277(21) 35.097(8)
c ¼ 4.6798(4)
8th b  ðTiVÞ ðIm3mÞ 48 (4) a ¼ 3.2460(6) 34.200(1)
a  TiðVÞ ðP63 =mmcÞ 51 (3) a ¼ 2.9444(6) 34.999(7)
c ¼ 4.6615(9)
Ti0.8V0.2 1st b  Ti0:8 V0:2 ðIm3mÞ 98.55(3) a ¼ 3.2322 (4) 33.76(1)
a  TiðH=OÞ ðP63 =mmcÞ 1.4 (1) a ¼ 2.956(2) 35.55(8)
c ¼ 4.69(1)
8th b  Ti0:8x V0:2þx ðIm3mÞ 94.1(3) a ¼ 3.239(1) 33.98(4)
a  TiðH=OÞ ðP63 =mmcÞ 3.0(5) a ¼ 2.947(2) 34.99(7)
TiðO; HÞ2 ðP42=mnmÞ 2.8(3) c ¼ 4.652(8) 68.0(1)
a ¼ 4.740(4)
c ¼ 3.027(6)

TiOHx [21] appeared after the cycling (XRD pattern can be seen are shown in Fig. 6. As can be seen in Fig. 6a, the Ti0.9V0.1 alloy
in Supplementary Information Fig. S5). that was cycled in Ar þ H2, is composed of aHCP and BCC
The results of the SEM investigations are in agreement with phases. In addition to that, a small amount of monohydride
the XRD study. Small oxygen-rich areas within the microstruc- (TiV)H~1 was observed. This indicates that some hydrogen was
ture of alloy Ti0.8V0.2 were observed for the cycled sample after not completely released during the dehydrogenation when
the 8th cycle. As can be seen from the EPMA results in Fig. 5, a the sample was heated to 760  C in Ar flow.
local segregation of oxygen is observed in the area crossed by the After the cycling of Ti0.9V0.1 in CO-containing mixtures,
line AB, with the highest concentration of oxygen in the center. XRD pattern revealed the presence of the HCP aTi and FCC d-
In addition, since the vanadium content is decreasing when the TiHx phases (see Fig. 6c). During the cycling, interaction with
oxygen content increases, the oxygen-rich area is most likely an CO resulted in enrichment by oxygen and/or carbon of the
island of an HCP Ti-based a-phase that dissolves large amounts surface layers, preventing hydrogen desorption at lower
of oxygen. Oxygen segregation was observed also for temperatures. Consequently, decomposition of d-TiHx (FCC)
TieVeCreFe alloys, where high oxygen concentration in the hydride was eliminated, in contrast with the behavior of the
alloy leads to the formation of aTi phase [22]. samples cycled in the mixtures of hydrogen that did not
A similar segregation did not occur in the Ti0.9V0.1 alloy since contain CO.
it contains high amounts of the HCP a-phase. However, it ap- After dehydrogenation, the Ti0.8V0.2 alloy cycled in Ar þ H2
pears that the oxide/oxygen rich phase formation did not give was composed of mainly BCC and only small amounts of HCP
significant impact on the hydrogenation behavior, probably a-phase. As can be seen from Fig. 6b, the composition of the
because the amount of these phases in Ti0.8V0.2 was rather dehydrogenated alloy was similar to that of the alloy having
small. been cycled in pure H2 and vacuum, except for the presence of
The X-ray diffraction patterns collected for Ti0.9V0.1 and a small amount of (TiV)Hx. This confirms that hydrides of the
Ti0.8V0.2 after cycling in Ar þ 25%H2 and Ar þ 25%H2 þ 1%CO Ti0.9V0.1 and Ti0.8V0.2 alloys were not completely decomposed

Fig. 5 e a) Metallographic cross section of a Ti0.8V0.2 powder particle after 8 cycles; b) Oxygen and vanadium point analyses
along the line AB in (a). Formation of an oxygen rich phase was observed.
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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0 1705

Fig. 6 e XRD diffraction patterns (Cu Ka) of dehydrogenated alloys after 11 cycles: a) Ti0.9V0.1 hydrogenated in Ar þ 25% H2; b)
Ti0.8V0.2 hydrogenated in Ar þ 25% H2; c) Ti0.9V0.1 hydrogenated in Ar þ 25% H2 þ 1%CO; d) Ti0.8V0.2 hydrogenated in
Ar þ 25% H2 þ 1%CO.

during the desorption step. These effects were gradually H2 þ 1% CO, there was no change in the bulk behavior and,
accumulated and led to the decreasing H capacity during the therefore, no other phases than BCC were observed after the
cycling. cycling.
For the Ti0.8V0.2 alloy, only BCC phase was observed after Fig. 7 shows results of an EPMA analysis that was per-
the sample had been cycled in Ar þ 25% H2 þ 1% CO. A formed within the microstructure of the Ti0.8V0.2 alloy after 8
competitive adsorption of H2 and CO [23] explains the small cycles in the gas mixture containing CO. A line analysis was
amount of hydrogen absorbed by this alloy at the applied performed along the line AB across the interface between two
conditions. The adsorption of CO at the surface is known to grains. The interface is seen as the weak curved line through
block the active sites for hydrogen adsorption from mixtures the center of Fig. 7a. As shown in Fig.7b, the oxygen content
of hydrogen with CO [24]. As a result, only limited amounts of was highest at the boundary, which is seen to be the interface
hydrogen can be absorbed by the Ti0.8V0.2 alloy during the between two Ti0.8V0.2 particles. This result is different from
absorption cycles. Bulk hydrogen storage properties can only the sample that was cycled in pure H2 (Fig. 5), in which aTieO
be altered at conditions when the alloy experiences phase was concentrated in small areas within the bulk. Thus,
hydrogenation-dehydrogenating cycling. Since no significant it can be concluded that CO-containing gas increased the ox-
hydride formation occurred during the cycling in Ar þ 25% ygen content at the surface of the TieV particles.

Fig. 7 e a) Metallographic cross section of a Ti0.8V0.2 powder particle after 8 cycles. Hydrogenation was done in Ar þ 25%
H2 þ 1%CO. b) Oxygen and vanadium point analyses performed along the AB line.
1706 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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0

Fig. 8 e Ti0.9V0.1Hx after hydrogenation and dehydrogenation at the conditions given in Fig. 4: a) Particle morphology; b) EDS
spectrum; c) high magnification view of a hydride particle surface; d) synchrotron XRD pattern (l ¼ 0.72085 A)  measured ex
situ.

Fig. 8 shows SEM images, EDS spectrum and SR XRD original sample was only partially recovered after the expo-
spectra from a Ti0.9V0.1H1.5 hydride surface after 21 hydroge- sure to CO.
nation/dehydrogenation cycles tested at varying conditions
given in Fig. 4. Fig. 8a shows typical particle morphologies at Surface studies
low magnification, and Fig. 8c gives a high magnification view
of a hydride particle surface. As can be seen from the EDS Surface studies were performed on the dehydrogenated alloys
spectrum in Fig. 8b, the surface of the hydride contained sig- which were cycled in pure hydrogen gas and in gaseous flows
nificant amounts of oxygen. In Fig. 8c it can be observed that in using the mixtures Ar þ 25% H2 and Ar þ 25% H2 þ 1% CO.
addition to the oxygen, deposits of carbon materials were also
present at the hydride surface. The carbon deposits must XPS studies
originate from the dissociation of CO, as frequently has been An XPS study was performed on the Ti0.8V0.2 surface before
observed in steam reforming processes performed at and after cycling under ideal conditions, i.e. hydrogenation in
300e1000  C [25]. Rietveld refinement of the X-ray powder pure H2 and dehydrogenation in vacuum. Ti2p, V2p, and O1s
diffraction patterns presented in Fig. 8d showed that the spectra were recorded. As can be seen from Fig. 9, the Ti2p
sample was composed of tetragonal BCT hydride (61 wt.%) and spectrum can be deconvoluted into four individual spectra
HCP a-Ti (39 wt.%). The unit cell lattice parameters of the a-Ti with the main peaks located at 458.8 eV, 457.5 eV, 455.1 eV and
phase were larger than that of pure aTi, which indicates that 454.1 eV, which corresponds to titanium in the four oxidation
the HCP phase contained hydrogen and/or oxygen. This result states Tiþ4, Tiþ3, Tiþ2 and Ti0, respectively [26]. As shown in
is in agreement with in situ SR-XRD data obtained in our pre- Fig. 9a, after the 1st dehydrogenation, the intensity of the Tiþ4
vious work [16], in which we reported that high amounts of energy state (458.8 eV) was the highest one, showing that the
the HCP a-phase were not converted to hydride during rehy- sample surface contained mainly TiO2. As can be seen in
drogenation in a gas mixture containing CO. Furthermore, the Fig. 9b, after the 8th dehydrogenation, the surface contained
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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0 1707

Fig. 9 e Typical XPS Ti2p spectra of Ti0.8V0.2 vacuum-dehydrogenated powder after: a) 1st dehydrogenation; b) 8th
dehydrogenation; c) 1st dehydrogenation after 5 min sputtering; d) 8th dehydrogenation after 5 min sputtering.

less oxygen than after the 1st dehydrogenation. In other phase within the bulk, resulting in a reduction in the oxygen
words, a thinner oxide layer was observed after 8 de- content at the surface.
hydrogenations than after the first dehydrogenation. The V2p
and O1s intensities were very low, the highest intensity AES studies
observed in the V2p spectrum was only 1/3 of the maximum AES spectra of Ti0.8V0.2 after 8 vacuum dehydrogenations are
intensity of the Ti2p spectrum, so deconvolution was not shown in Fig. 10a. As can be seen from the figure, oxygen and
performed (see supplementary data). Nevertheless, the V2p carbon were observed in addition to titanium and vanadium.
spectrum consisted of 3 individual spectra related to Vþ5, Vþ3 The oxygen peak at ~513 eV appears to be rather dominant,
and 2V0, which indicates the presence of the V2O5, VO2 and and oxygen is naturally bound to Ti since the prominent Ti
metallic vanadium, respectively [27]. The highest contribution peaks at ~388 eV and ~419 eV are very strong. Previous ob-
was observed at 517.19 eV, which corresponds to V2O5. servations have shown that possible compounds giving such
After 5 min of ion sputtering inside the XPS chamber, spectra are TiO and TiO2 [28,29]. A weak vanadium peak at
which resulted in removal of about ~5e~10 nm thickness of 473 eV was observed in addition, most likely in the form of a
the surface layer, it became possible to collect data on the mixture of three various vanadium oxides. After sputtering for
composition of the sub-surface layers. As can be seen in 210 s, which resulted in removal of about 7 nm of the surface
Fig. 9c, after sputtering, the normalized intensity of metallic layer, the level of oxygen was reduced, as can be concluded
titanium, Ti0, and the contribution from the lower oxidation from the reduced intensities of its peaks in the spectrum. The
state of titanium increased, which indicates that only a thin level of oxygen was decreased even further after a longer
surface layer, not thicker than 10 nm, was enriched with C/O. sputtering time. Ion sputtering for 690 s, corresponding to
The Ti2p spectrum that was recorded from the sample after 23 nm surface removal, resulted in an increase in the relative
completing the 8th dehydrogenation is shown in Fig. 9d, and it intensity of the Ti peaks, which means that the oxygen con-
can be observed that the amount of Ti0 has become higher centration decreased even further when moving into the bulk.
than the other oxidation states in this condition. From these After 690 s ion sputtering, the carbon peak at ~272 eV dis-
results, it can be inferred that cycling in pure hydrogen and appeared, indicating that its presence at the sample surface
dehydrogenation in vacuum leads to removal of oxygen from originated from handling of the sample. AES spectra of
the alloy surface. Alternatively, EPMA data indicated that ox- Ti0.8V0.2 cycled in a mixture of Ar þ 25H2 þ 1% CO are pre-
ygen concentrated in specific locations as an oxygen-rich sented in Fig. 10b. It can be seen that the surface composition
1708 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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0

hydrogen and in less than 60 s formed saturated hydrides


containing FCC g(Ti,V)H2 as the majority constituent, with
an H storage capacity reaching 3.7 wt.% H for Ti0.9V0.1H~2. At
higher temperatures, 450e550  C, the hydrogenation process
led to increased formation of the lower hydride BCC b-(Ti,V)
H~0.7 [17] resulting in decreased reversible hydrogen storage
capacities.
Hydrogenation in an Ar þ 10% H2 gas flow reduced the
reversible H storage capacities (down to 1e3 wt. % H); the
extent of the decay was related to the alloy composition and
the hydrogenation temperature. In most cases the hydride
formed during the hydrogenation at 425e575  C was found to
be a lower bmonohydride with a BCC type structure. This
correlated well with the dehydrogenation properties of the
formed hydrides, as their desorption spectra contained only
one high temperature peak of hydrogen evolution.
Kinetic analysis of the process of hydrogen absorption
showed that hydrogenation of the two TieV alloys followed
the, respectively, two- and three-dimensional Avrami-Erofeev
nuclei growth model. Differences in the dimensionality of the
nuclei growth are likely to be caused by the microstructural
and phase-structural state of the alloys. Indeed, Ti0.9V0.1 alloy
is composed of a mixture of HCP and BCC phases (three-
dimensional model), while a BCC type Ti0.8V0.2 has close to a
single phase microstructure (two-dimensional model). How-
ever, in common to both studied alloys, the growth of the
nuclei of the hydride phases was found to be the rate-limiting
step of the hydrogenation process.
Even though cycling in pure hydrogen/vacuum resulted in
slight changes of the phase composition of the alloys, e.g.
Fig. 10 e AES spectra of dehydrogenated Ti0.8V0.2 after 9 increased content of HCP aphase and formation of oxygen-
hydrogenation and dehydrogenation cycles: a) rich TieO phase in the Ti0.8V0.2 alloy, these changes did not
hydrogenation in pure H2 and dehydrogenation in affect the values of the reversible hydrogen capacity.
vacuum; b) hydrogenation in Ar þ 25%H2 þ 1%CO and In contrast, hydrogen cycling in a flow of Ar þ H2 resulted
dehydrogenation in pure Ar. in reduced amount of hydrogen in the hydrides, down to
~2e2.5 wt.% H, already during the first testing cycles. The
hydrogen content was found to continuously decrease upon
cycling. This was caused by the presence of undesorbed
hydrogen that was trapped in the remnant (TiV)Hx observed in
was similar to that of the sample cycled at ideal conditions,
the cycled alloys. The amount of remnant (TiV)Hx increased
which was dominated by titanium and oxygen. The difference
upon cycling, which in part was caused by a tendency of for-
was that a carbon peak was observed after 690 s sputtering
mation of Ti- and V-rich constituent phases during the cycling
after the cycling in the presence of CO. This means that CO
as well as hydride separation during the hydrogenation. Thus,
enriched the carbon content on the surface of the sample
at some stage, formation of Ti-rich high stability TieVeH
cycled in the gas mixture.
phase increases and negatively affects the cycling behavior.
In presence of carbon monoxide that was added to the gas
flow (Arþ25%H2 þ 1%CO), a limited amount of hydrogen, i.e.
Summary and final remarks ~0.6 wt.% H, was absorbed during the hydrogen absorption-
desorption cycling at 425  C. This amount of hydrogen
Ti0.9V0.1 and Ti0.8V0.2 were tested as candidate hydrogen remained relatively constant during the cycling.
absorbent alloys to be employed in Hydrogen Sorption Interestingly, when the hydrogenation was performed
Enhanced Reforming (HSER) e a novel process allowing an non-isothermally in the experiments covering the tempera-
improved efficiency of natural gas reforming [8]. ture range RT-800  C, the amount of the absorbed hydrogen
Hydrogen absorption experiments were performed in two slightly increased to ~1 wt.% H, despite of the higher CO
different modes; (a) quasi-isothermally at static conditions content in the mixture (see Fig. 7). In situ SR-XRD experiments
using pure hydrogen gas, and (b) isothermally at 425 C-575  C showed that d-(TiV)H~1.5 is formed during non-isothermal
using a flow of gaseous mixture containing argon, 1e2% CO hydrogenation in a gaseous flow of 90% H2 þ 10% CO [9]. In
and 25% H2. that case, after the 3rd non-isothermal cycle in the tempera-
When the hydrogenation was performed at 450  C and ture range RT-800  C the sample was composed of 70 wt.% the
4 bar H2, both alloys instantly started to interact with dhydride. Differences in the hydrogenation mechanism have
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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0 1709

roots in variations of the employed experimental conditions, ability of the (Ti,V) alloys to absorb hydrogen was significantly
including the time/temperature for introducing CO in the gas suppressed in the presence of CO. Two complementary
mixture and, furthermore, conditions of the hydrogenation mechanisms stand as the reasons for that. At lower temper-
processing - isothermal or non-isothermal. atures, CO blocked the active sites of hydrogen adsorption at
As an illustration of the effect of the conditions of the al- the surface. It was possible to revert this condition since the
loys processing in hydrogen, HCP aphase was not observed hydrogenation capabilities of the alloys were partially recov-
in the present work in the powder XRD pattern of the Ti0.8V0.2 ered when CO was removed from the gas flow. At high pro-
sample cycled isothermally in Ar þ 25%H2 þ 1%CO, in contrast cessing temperatures above 425  C, chemical decomposition
with the significant amounts of HCP aphase that was re- of CO took place at the sample surface. This resulted in (a)
ported by Børresen at al. [8] to form during non-isothermal carbon deposition on the surface, blocking the active surface
cycling. AES analysis of the surface of a cycled specimen sites for hydrogen adsorption, and (b) oxygen diffusion into
(Fig. 10) showed that the surface layers contained significant the bulk alloys forming solid solutions of oxygen within the
amounts of carbon and oxygen. This agrees well with the re- alloy as well as oxide films at the grain boundaries, affecting
sults of an investigation on LaNi5 [10], showing that, at high both kinetics and thermodynamics of the hydrogen exchange.
interaction temperatures above 300  C, CO decomposes at the The most viable advantage obtained by using hydrogen
metal surface, leading to deposition of carbon and formation absorbent alloys in the HSER process lies in the possibility to
of methane and a surface oxide. Even at low interaction limit the process temperature below the value where unre-
temperatures, CO has been shown to dissociate on a LaNi5 coverable changes of the alloy (surface and bulk) take place
surface, but at those conditions the reaction products were due to interaction with carbon monoxide, and at the same
methane and H2O [30]. The effect of carbon contamination time to adjust the reversible hydrogen storage capacity of the
appears to be related to the gas composition and temperature, formed hydrides to the operating temperatures, leading to an
and it has been shown in previous works that gas mixtures overall improvement of the efficiency of the process.
containing carbon monoxide reduce the number of active
surface sites for metalehydrogen interaction [23,24,31].
The oxygen that was detected within the HCP a-phase after
Acknowledgments
non-isothermal hydrogenation, must originate from the
interaction of the alloy surface with CO. When the content of
Financial support from the Norwegian Research Council and
CO in the gaseous mixture is high, non-isothermal cycling at
Statoil ASA are gratefully acknowledged. The thank staff of
high temperatures will create conditions that allow oxygen
SNBL/ESRF for the skillful assistance during SR XRD
atoms to diffuse into the bulk material, leading to a shift in
experiments.
equilibrium towards formation of HCP Ti a-phase which is
able to dissolve large amounts of interstitial oxygen [32].
Nevertheless, it was demonstrated that the reversible
hydrogen storage capacity was recovered up to 80% of its Appendix A. Supplementary data
initial value when CO was removed from the gas mixture
(Fig. 7). A similar recovery of the hydrogenation behavior has Supplementary data related to this article can be found at
also been observed to occur during low temperature cycling of http://dx.doi.org/10.1016/j.ijhydene.2015.11.077.
LaNi5 [30] and TieMn alloys [9] in hydrogen gas containing,
respectively, 0.1 and 0.5% CO. The reason for this recover-
references
ability is probably associated with removal of carbon/carbon
monoxide from the alloy surface and thus reactivation of the
active sites for hydrogen adsorption.
[1] Ball M, Wietschel M. The future of hydrogen e opportunities
and challenges. Int J Hydrogen Energy 2009;34:615e27.
[2] Angeli SD, Monteleone G, Giaconia A, Lemonidou AA. State-
Conclusions of-the-art catalysts for CH4 steam reforming at low
temperature. Int J Hydrogen Energy 2014;39:1979e97.
In the present work, hydrogen storage properties of Ti0.9V0.1 [3] Barelli L, Bidini G, Gallorini F, Servili S. Hydrogen production
and Ti0.8V0.2 alloys were studied using various experimental through sorption-enhanced steam methane reforming and
and characterization techniques. Ti-rich Ti0.9-0.8V0.1-0.2 solid membrane technology: a review. Energy 2008;33:554e70.
[4] Uemiya S, Sato N, Ando H, Matsuda T, Kikuchi E. Steam
solution alloys demonstrated excellent hydrogenation ki-
reforming of methane in a hydrogen-permeable membrane
netics, forming saturated g-dihydrides with high reversible reactor. Appl Catal 1990;67:223e30.
hydrogen storage capacity (up to 3.7 wt.% H) in less than 60 s. [5] Balasubramanian B, Lopez Ortiz A, Kaytakoglu S,
A complete reversibility of the hydrogen absorption- Harrison DP. Hydrogen from methane in a single-step
desorption capacity was achieved during the cycling experi- process. Chem Eng Sci 1999;54:3543e52.
ments (up to 9 cycles) when the hydrogenation was performed [6] Ochoa-Ferna  ndez E, Rønning M, Grande T, Chen D. Synthesis
and CO2 Capture properties of Nanocrystalline Lithium
in pure H2 gas and the hydride decomposition was accom-
Zirconate. Chem Mater 2006;18:6037e46.
plished by in a vacuum. However, when cycling was per-
[7] Park J, Yi KB. Effects of preparation method on cyclic stability
formed in a flow of H2 mixed with Ar, decreased values of the and CO2 absorption capacity of synthetic CaOeMgO
reversible hydrogen capacity were observed as a consequence absorbent for sorption-enhanced hydrogen production. Int J
of the formation of stable b-(Ti,V)H1-1.2 lower hydrides. The Hydrogen Energy 2012;37:95e102.
1710 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 4 1 ( 2 0 1 6 ) 1 6 9 9 e1 7 1 0

[8] Borresen B, Rytter E, Aartun I, Krogh B, Ronnekleiv M. 2010. 2005;404e406:335e8. http://dx.doi.org/10.1016/


Method and reactor for production of hydrogen. j.jallcom.2005.05.001.
US20100047158A1. [21] Swope RJ, Smyth JR, Larson AC. H in rutile-type compounds:
[9] Block FR, Bahs H-J. Selective absorption of hydrogen by Ti- I. Single-crystal neutron and X-ray diffraction study of H in
Mn-based alloys from gas mixtures containing CO or CH4. J rutile. Am Mineral 1995;80:448e53.
Common Met 1984;104:223e30. http://dx.doi.org/10.1016/ [22] Ulmer U, Asano K, Bergfeldt T, Chakravadhanula VSK,
0022-5088(84)90407-7. Dittmeyer R, Enoki H, et al. Effect of oxygen on the
[10] Sandrock GD, Goodell PD. Surface poisoning of LaNi5, FeTi and microstructure and hydrogen storage properties of
(Fe,Mn)Ti by O2, Co and H2O. J Common Met 1980;73:161e8. VeTieCreFe quaternary solid solutions. Int J Hydrogen
http://dx.doi.org/10.1016/0022-5088(80)90355-0. Energy 2014;39:20000e8.
[11] Sandrock GD, Goodell PD. Cyclic life of metal hydrides with [23] Hou K, Hughes R. The effect of external mass transfer,
impure hydrogen: overview and engineering considerations. competitive adsorption and coking on hydrogen permeation
J Common Met 1984;104:159e73. through thin Pd/Ag membranes. J Membr Sci
[12] San-Martin A, Manchester FD. The HTi (Hydrogen- 2002;206:119e30.
Titanium) system. Bull Alloy Phase Diagr 1987;8:30e42. [24] Li A, Liang W, Hughes R. The effect of carbon monoxide and
[13] Reilly JJ, Wiswall RH. Higher hydrides of vanadium and steam on the hydrogen permeability of a Pd/stainless steel
niobium. Inorg Chem 1970;9:1678e82. membrane. J Membr Sci 2000;165:135e41.
[14] Maeland AJ. Investigation of the VanadiumdHydrogen [25] Bartholomew CH. Carbon deposition in steam reforming and
system by X-Ray diffraction Techniques1,2. J Phys Chem Methanation. Catal Rev 1982;24:67e112.
1964;68:2197e200. [26] Lu G, Bernasek SL, Schwartz J. Oxidation of a polycrystalline
[15] Yartys VA, Lototsky MV. An Overview of hydrogen storage titanium surface by oxygen and water. Surf Sci
methods. In: Veziroglu TN, Zaginaichenko SY, Schur DV, 2000;458:80e90.
Baranowski B, Shpak AP, Skorokhod VV, editors. Hydrog. [27] Silversmit G, Depla D, Poelman H, Marin GB, De Gryse R.
Mater. Sci. Chem. Carbon Nanomaterial. Netherlands: Determination of the V2p XPS binding energies for different
Springer; 2004. p. 75e104. vanadium oxidation states (V5þ to V0þ). J Electron Spectrosc
[16] Suwarno S, Gosselin Y, Solberg JK, Maehlen JP, Williams M, Relat Phenom 2004;135:167e75.
Krogh B, et al. Selective hydrogen absorption from gaseous [28] Davis GD, Natan M, Anderson KA. Study of titanium
mixtures by BCC Ti-V alloys. Int J Hydrogen Energy oxides using Auger line shapes. Appl Surf Sci
2012;37:4127e38. 1983;15:321e33.
[17] Suwarno S, Solberg JK, Mæhlen JP, Denys RV, Krogh B, Ochoa- [29] Shih HD, Jona F. Low-energy electron diffraction and auger
Ferna ndez E, et al. Non-isothermal kinetics and in situ SR electron spectroscopy study of the oxidation of Ti {0001} at
XRD studies of hydrogen desorption from dihydrides of room temperature. Appl Phys 1977;12:311e5.
binary TieV alloys. Int J Hydrogen Energy 2013;38:14704e14. [30] Han S, Zhang X, Shi S, Tanaka H, Kuriyama N, Taoka N, et al.
[18] Førde T, Maehlen JP, Yartys VA, Lototsky MV, Uchida H. Experimental and theoretical investigation of the cycle
Influence of intrinsic hydrogenation/dehydrogenation durability against CO and degradation mechanism of the
kinetics on the dynamic behaviour of metal hydrides: a LaNi5 hydrogen storage alloy. J Alloys Compd
semi-empirical model and its verification. Int J Hydrogen 2007;446e447:208e11.
Energy 2007;32:1041e9. [31] Iyoha O, Howard B, Morreale B, Killmeyer R, Enick R. The
[19] Andreasen A, Vegge T, Pedersen AS. Compensation effect in effects of H2O, CO and CO2 on the H2 Permeance and surface
the Hydrogenation/Dehydrogenation kinetics of metal characteristics of 1 mm Thick Pd80wt%Cu membranes. Top
hydrides. J Phys Chem B 2005;109:3340e4. Catal 2008;49:97e107.
[20] Evard EA, Gabis IE, Voyt AP. Study of the kinetics of hydrogen [32] Murray JL, Wriedt HA. The OTi (Oxygen-Titanium) system.
sorption and desorption from titanium. J Alloys Compd J Phase Equilibria 1987;8:148e65.

You might also like