Sie sind auf Seite 1von 9

Available online at www.sciencedirect.

com

ScienceDirect
Materials Today: Proceedings 5 (2018) 13726–13734
www.elsevier.com/locate/procedia

AEM 2016

Effect of -- cyclic phase transformations on the


hydrogenation/dehydrogenation kinetics of MgH2 binary system
M. Sherif El-Eskandarany, Ehab Shaban, and A. Alkandary*
Nanotechnology and Advanced Materials Program, Energy and Building Research Center, Kuwait Institute for Scientific Research,
Safat 13109, Kuwait

Abstract

Reactive ball milling (RBM) approach, using a high-energy ball mill operated at room temperature under 50 bar hydrogen gas
atmosphere was employed to synthesize MgH2 powders. A single low-pressure phase (tetragonal structure) of -MgH2 phase
with tetragonal structure was obtained after 6 h RBM time. This phase tended to transformed into less stable high-pressure -
phase (orthorhombic structure) upon milling for 25 h. However, increasing the RBM time to 75 h led to -to- phase
transformations. Cyclic -- phase transformations were observed several times with changing the RBM time between 25 – 200
h. The results of the present study showed that existence of metastablephase in the milled powders led to a significant
improvement on the hydrogenation/dehydrogenation kinetics. The decomposition temperature and the activation energy of the
powders obtained after 200 h of RBM time, which containing a large volume fraction of -phase, were 399 oC and 131 kJ/mol,
respectively. In addition, at 300 oC, MgH2 nanocrystalline obtained upon 200 h of RBM possessed excellent hydrogenation
properties for any pure MgH2 system, indexed by high hydrogen storage capacity (7.54 wt.%) with complete 600
absorption/desorption cycles.

© 2018 Elsevier Ltd. All rights reserved.


Selection and Peer-review under responsibility of 1st International Conference on Advanced Energy Materials and 8th
International Conference on Advanced Nanomaterials.

Keywords: Reactive ball milling; metal hydrides; lattice defects; kinetics; cycle-life-time;

* Corresponding author. Tel.: +965 69622622.


E-mail address: arkandary@kisr.edu.kw

2214-7853 © 2018 Elsevier Ltd. All rights reserved.


Selection and Peer-review under responsibility of 1st International Conference on Advanced Energy Materials and 8th International Conference
on Advanced Nanomaterials.
El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734 13727

1. Introduction

Reactive ball milling (RBM) technique was first investigate by Calka [1] and El-Eskandarany et al. [2] in the
beginning of 1990’s. This technique has offered practical methodology based on solid-state gas-solid reaction for
synthesizing wide verities of metal nitrides and metal hydrides at room temperature, using high-energy ball milling
approach [3]. MgH2 has been considered as the most promising hydrogen storage candidate for real applications in
light vehicles. This because Mg metal is a light, cheap metal, and can store hydrogen up to 7.60 wt.% [4, 5]. In spite
of these attractive properties, the slow kinetics of hydrogenation and dehydrogenation of MgH2 present the major
barriers for potential in fuel cell applications [6, 7]. On way used to improve the hydrogenation/dehydrogenation
kinetics of MgH2 powders is to subjecting the powders to long-term of ball milling that can be extended to several
hundred hours [8]. Such long-milling intervals lead to increases the influence of grain boundaries and enhance the
diffusion of hydrogen [9]. Moreover, it also leads to destabilize the stable phase of MgH2 upon introducing many
types of imperfections such as lattice and point defects, which are generated in the lattice of the ball milled powders
[8]. Such defects can be also created upon cold rolling of metal Mg, as was demonstrated by Floriano et al. [10].
Increasing the RBM time developed the grain size and creates nanocrystalline MgH2 powders, and this in turn leads
to increase both of the grain boundaries and the active surface area [7]. Thus, a shorten hydrogen diffusion path
through the bulk area is expected to be achieved [11].

In the present study we have demonstrated cyclic phase transformations that took place between  (most stable
phase) and  (metastable-high pressure phase) phases of MgH2 upon increasing the RBM time of pure Mg metal
under high pressure of a hydrogen gas atmosphere at room temperature. The effects of the cyclic phase
transformations and grain size refining on the thermal stability and hydrogenation/dehydrogenation kinetics of
MgH2 system are presented and discussed.

2. Experimental

Pure elemental Mg metal powders (<45m, 99.5 %) and hydrogen gas (99.999 %) were used as starting
materials. A certain amount of the powders (5 g) was balanced inside a helium gas atmosphere (99.99%) - glove box
(UNILAB Pro Glove Box Workstation, mBRAUN, Germany). The powders were then sealed together with fifty
hardened steel balls into a hardened steel vial (220 ml in volume), using a gas-temperature-monitoring system (GST;
supplied by evico magnetic, Germany). The ball-to-powder weight ratio was 40:1. The vial was then evacuated to
the level of 10-3 bar before introducing H2 gas to fill the vial with a pressure of 50 bar. The milling process was
carried out at room temperature using high energy ball mill (Planetary Mono Mill PULVERISETTE 6, Fritsch,
Germany). The ball-milling experiments were interrupted after selected milling times and a small amount of the
powder was taken from the vial in the glove box. The average crystal structure of all samples was investigated by X-
ray diffraction (XRD) with CuK radiation, using 9kW Intelligent X-ray diffraction system, SmartLab-Rigaku,
Japan. The local structure of the synthesized material powders at the nanoscale was studied by 200 kV-field
emission high resolution transmission electron microscope (FE-HRTEM) supplied by JEOL-2100F, Japan. The
hydrogenation properties; including absorption/desorption kinetics were investigated via Sievert’s method, using
PCTPro-2000, provided by Setaram Instrumentation. Shimadzu Thermal Analysis System /TA-60WS, using
differential scanning calorimeter (DSC) was employed to investigate the thermal stability indexed by the
decomposition temperatures of MgH2.

3. Results and discussions

XRD was used to monitor the structural changes performed in Mg powders upon RBM under a high pressure (50
bar) of hydrogen gas atmosphere for different milling time. The XRD patterns of the initial pure hcp-Mg powder (0
h) is displayed in Fig. 1(a). At this initial stage of RBM time (0 h), the structure of the powders was polycrystalline
hcp-Mg phase with large grains, suggested by the sharp diffraction Bragg’s peaks shown in Fig. 1(a). After 1 h of
RBM (Fig. 1(b)), new Bragg’s peaks appeared, suggesting the formation of a new crystalline phase. The major
13728 El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734

Bragg’s peaks appeared at 2 of 27.85o, 35.72o, 39.82o and 54.56o, as shown in Fig. 1(b). Comparing these values
with those reported for the tetragonal-MgH2 phase confirmed the formation of significant mole fractions of -
MgH2 phase after this early stage of RBM.

Fig. 1 XRD patterns of Mg powders after ball milling under hydrogen gas pressure for (a) 0, (b) 1, (c) 3, (d) 6, (e) 25 h, (f) 37.5,
(g) 50, (h) 100 and (i) 200h of the RBM time.

After 3 h of RBM time, the Bragg’s peaks corresponding to the -MgH2 phase become pronounced and sharp,
suggesting an increase in the mole fraction of the reacted phase (-MgH2) against the un-reacted phase of pure hcp-
Mg phase, as elucidated in Fig. 1(c). After 6 h of RBM time, the minor Bragg-peaks corresponding to hcp-Mg were
completely vanished, whereas the intensities major were dramatically decreased and can hardly be seen, as shown in
Fig. 1(d). Moreover, the intensities of the Bragg-peaks for -MgH2 were continuously increased, indicating a
monotonical increasing in the mole fractions of the formed reacted phase of -MgH2 (Fig. 1(d)). Toward the
completion of the RBM processing time (25 h), nanocrystalline -MgH2 phase coexisting with -MgH2 phase were
obtained, suggested by the obvious broadening in the Bragg’s lines shown in Fig. 1(e). In order to investigate the
effect of further RBM time on the stability of the formed -MgH2, the powders were continuously ball milled for
longer RBM time under the same experimental conditions. Figure 1(f) displays the XRD patterns at the low angle
scattering side (20o-37o) for the powders that were ball milled for 37.5 h. The -MgH2 phase tended to transform
into metastable -MgH2 phase upon RBM for 37.5 h, as indicated by the relative increase in the intensities of the
Bragg lines corresponding to -MgH2 phase (Fig. 1(a)).

In a cyclic phase transformation, the intensity of the major Bragg’s peak (110) of -MgH2 phase is drastically
decreased upon RBM for 75 h (Fig. 1(g)), suggesting the crystallization of the -phase into the most stable phase of
-MgH2. Again, this obtained -MgH2 phase could not withstand the mechanical deformations, combined with the
point and lattice defects that occurred in the ball-milled powders upon further RBM time to 100 h. This is indicated
by the obvious intensity increase on the Bragg peaks related to the -phase against those for the stable -phase, as
shown in Fig. 1(h). Continuous  to phase transformations took place upon RBM time to 200 h, indicated by the
dramatic increase on the (110) peak intensity corresponding to the -MgH2 phase (Fig. 4(d)). Simultaneously, the
El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734 13729

intensity of the Bragg’s peak (110) for -MgH2 crystals was hardly seen, indicating a significant decrease in the
mole fraction of -MgH2 phase, as presented in Fig. 1(i).
The bright field image (BFI), dark field image (DFI) and the corresponding selected area diffraction patterns
(SADP) of the powders obtained after 25 h of RBM time are shown together in Figs. 2(a), (b) and (c), respectively.
The powders had irregular grain-like morphology with an average grain size of 62 nm. The local structure of the
powder shows a -MgH2 phase (tetragonal) overlapped with metastable -phase (-PbO2 type structure) without any
evidences for the existence of unprocessed hcp-Mg, as displayed in the Miller-indexed SADP (Fig. 2(c)).

The FE-HRTEM-DFI for the powders obtained after 200 h of the RBM time is displayed in Fig. 3. Comparing
this image with that obtained in Fig. 2(b), it can be said that further milling (200 h) led to a significant refining in the
grain sizes for both phases of MgH2. The powders at this stage of milling are composed of nearly spherical fine
grains with average size of about 2.8 nm. It is worth noting in this figure that significant volume fractions of -MgH2
grains were detected, implying the continuity of the to cyclic phase transformations that took place upon
increasing the RBM time.

Fig. 2. (a) BFI, (b) DFI, and (b) the corresponding SADP of MgH2 obtained after 25 h of RBM time.

Fig. 3. FE-HRTEM/DFI of nanocrystalline MgH2 powders obtained after 25 h of RBM time.

DSC was performed with a constant heating rate of 20 K/min in order to investigate the thermal stabilities of
MgH2 powders obtained after different RBM time. The DSC traces of the powders obtained after different RBM
times are shown in Fig. 4. The onset temperature of the hydrogen desorption peak for -MgH2 phase obtained after
13730 El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734

6 h of RBM time (Fig. 4(a)) appeared at a temperature of 714 K (441 oC). The decomposition event took place
through a single-endothermic peak, suggesting the existence of a single phase of -MgH2 in the ball-milled powders.
The onset- temperature for this endothermic reaction tends to shift to the low temperature side (707 K) upon milling
the powders to 12.5 h of the RBM time (Fig. 4(b)), suggesting the destabilization of the formed -MgH2 phase. The
DSC curve of the MgH2 powders obtained after 25 h of the RBM time (Fig. 4(c)) revealed a shoulder-like
endothermic peak appearing at an onset temperature of 692 K (419 oC). This endothermic event vanished at an end-
set temperature of about 750 K (477 oC), as shown in Fig. 4(c). Moreover, the onset temperature of the main
endothermic peak did not show significant changes when compared with the powders obtained after 12.5 h of the
RBM time.

After 37.5 h of the RBM time, the low- and high-temperature events clearly, are seen into two separated peaks
(Fig. 4(d)) appearing at onset temperatures of 679 K (409 oC) and 708 K (435 oC), respectively. In retrospect, Fig.
4(e) shows the two separated peaks tending to be overlapping with each other for the sample obtained after 50 h of
the RBM time. The onset temperatures of these two endothermic events were continuously shifted to the low
temperature side and recorded to be 670 K (397oC) and 709 K (336oC), respectively. Obviously, the area under the
first endothermic peak showed to be relatively larger than that for the second endothermic reaction, indicating that
the volume fraction of the existing phase in the milled powder was larger than -MgH2 phase. When the powders
were successively subjected to further RBM time (75 h), the ratio between the formed two phases (-MgH2 and-
MgH2) was changed and the volume mole fraction of the high-temperature phase becomes larger, as indicated by the
area under the second endothermic peak (Fig. 4(f). Moreover, the onset temperatures of the two endothermic events
shifted to the low temperature side recorded as 650 K (377oC) and 669 K (396oC), respectively, suggesting a
continuous destabilization of the formed hydride phases.

Fig. 4. DSC curves of MgH2 powders obtained after RBM for (a) 6, (b) 12.5, (c) 25, (d) 37.5, (e) 50,
(f) 75, (g) 100, (h) 150, (i) 200 and (k) 300 h.
Again, and after 100 h of the RBM time (Fig. 4(g)), the two peaks were seen to split into individual events
appearing at 652 K (379oC) and 672 K (399oC), respectively. This indicated a cyclic- phase transformations.
The -MgH2 phase, which was remarkably grown after 100 h of the RBM time, tended to be more pronounced for
the sample obtained after 150 h and 200 h of the RBM time, as displayed in Figs. 4(h) and 4(i), respectively. Further
RBM (300 h) was seen to lead to a dramatic  phase transformations, suggested by the obvious decrease in the
area under the first endothermic peak, related to the decomposition of -MgH2 phase, as presented in Fig. 4(k).
El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734 13731

The effect of the RBM time on the enthalpy change of hydrogen decomposition (Hdec) and the grain sizes of the
formed hydride phases are shown in Fig. 5(a). The total Hdec displayed in Fig. 5(a) refers to the total area under the
endothermic peaks of -MgH2 and -MgH2 phases at specific RBM time. TheHdec of the first endothermic peak
presented in Fig. 5(b) refers to the area under the first endothermic peak of -MgH2 at specific RBM time. The total
Hdec tended to increase drastically during the first few hours of the RBM time (6-37.5 h) to get a value of about 2
kJ/g, as shown in Fig. 5(a). This drastic increase could be associated to a dramatic decrease in the grain sizes that
showed a continuous decreasing from about 200 nm to almost 10 nm in diameter (Fig. 5(a)). In parallel to the
successive grain size refinement taking place upon further RBM time (50 h -300 h), where the grain sizes reached to
the level between 8 nm – 4 nm in sizes, a continuous increasing rate in the value of total Hdec can be seen to reach
to a value of about 5 kJ/g. This implies the dependence of hydrogen releasing on the grain size. As the grain sizes
are saturated at the values of 3-4 nm in diameter during the last stage of RBM (300-700 h), the changes in the value
of Hdec could hardly be seen and saturated at the values of 5 -5.8 kJ/g, as shown in Fig. 5(a).

The area under the first endothermic peaks (low-temperature peaks) presented in Fig. 4 refers to the enthalpy
changes in the decomposition (Hdec) for -MgH2. TheHdec values are very sensitive to the structural changes
occurring in the powders during the different stages of RBM. In order to investigate the cyclic growing and
degradation of -phase during the different stages RBM process, the calculated Hdec values for -MgH2 are plotted
against the RBM time in Fig. 5(b). During the first stage of RBM time (6-12.5 h), the values Hdec for the -phase
are nil (Fig. 5(b)), indicating the absence of this metastable phase in the ball-milled powders. At this stage of RBM,
the powders consisted of a single -phase, as previously shown in Fig. 1. A dramatic increase in the Hdec values (0-
1.6 kJ/g) can be seen upon increasing the RBM time (12.5-50 h). Subsequently, Hdec decreased to 0.95 kJ/g upon
milling for 75 h of the RBM time, suggesting a significant decreasing of -phase in the milled powders. However,
increasing the RBM time to 100 h led to destabilize the -phase to gradually transforming into a less stable -phase,
indicated by the monotonical increase of Hdec from 1.2 kJ/g (100 h) to a higher value at 2.1 kJ/g after 200 h of
RBM time, as shown in Fig. 5(b).

Fig. 5. Effect of RBM on the (a) total enthalpy change of decomposition and grain size of MgH2 powders, and (b) enthalpy change of
decomposition of the first endothermic DSC peaks for -MgH2 phase displayed in Fig. 4.
During the next stage of the RBM (200-400 h) the metastable-MgH2 phase complied with gradual phase
transformations into -MgH2, indicated by the dramatic decrease in the values of Hdec that reached to 0.5 kJ/g after
400 h of RBM time. Again, the formed stable -phase was destabilized and fell into -phase, suggested by that
increase observed in the Hdec value (2.2 kJ/g), as shown in Fig. 5(b). This metastable -phase was no seen; longer to
withstand the imperfections and the mechanical deformation generated by the milling media and partially
crystallized to -phase after 500 h of RBM time, implied by the drop of the Hdec value that reached to 1.4 kJ/g (Fig.
5(b)).
13732 El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734

Figure 6 shows the absorption (a, c) and desorption (b, d) kinetics of the samples obtained after 25 h and 200 h of
the RBM time, respectively. At 250oC, the 25 h sample absorbed about 5 wt.% H2 within 5000 s and saturated at 6.2
wt.% H2 after about 32000 s. When these values are compared with those obtained at 200oC (5 wt.% H2/8120 s) and
225oC (5 wt.% H2/6074 s), one can say that increasing the applied temperature leads to a moderate improve on the
kinetics of hydrogen absorption improved with the temperature increase, as shown in Fig. 6(a). The effect of
temperature on the kinetics of desorption for the same sample is hardly seen, as displayed in Fig. 6(b). The hydrogen
released after 40000s at temperature of 200oC and 225 oC were 0.08 and 0.18 wt.%, respectively. In contrast to the
25 h samples, the powders obtained after 200 h of RBM showed significant improvement in the hydrogen
uptake/release, as presented in Figs. 6 (c) and 6(d), respectively. At 250oC, the 200 h sample absorbed and released
about 5 wt.% H2 within 348 s and 1657 s, respectively. The maximum hydrogen capacities for absorption and
consequent desorption attained saturation values of about 6.8 wt.% within 3140 s and 7400 s, respectively, as
presented in Fig. 6(d). The kinetics of absorption decreased remarkably upon decreasing the temperatures to 225oC
and 200oC, as shown in Fig. 6(c). At 225oC, the kinetics of absorption reached to its saturation value of 6.9 wt.%
within 3240 s. However, 35207 s was required for a complete desorption, as shown in Fig. 6(d). At 200oC, further
duration of time (5529 s) was necessary for complete absorption (~6.61 wt.%). However, the sample failed to
release its absorbed hydrogen completely even after 12394 s, as presented in Fig. 6(d).

Fig. 6. Effects of RBM time and applied temperature on the hydrogenation (a, c) and dehydrogenation (b, d) kinetics of MgH2 powders obtained
after RBM time of 25 h (a, b) and 200 h (c, d).

The cycle-life-time for the 200 h sample presented in Fig. 7(a) shows an excellent hydrogenation properties,
indexed by high cyclic stability even after about 600 h (~ 600 cycles) with no obvious degradation on the hydrogen
storage capacity, which show nearly constant values of absorption and desorption of +7.50 wt.% and -7.50 wt.%,
respectively, as displayed in Fig. 7(a). The XRD analysis (Fig. 7(b)) of the sample taken after the completion of the
600 cycles of sorption/desorption showed sharp Bragg’s peaks corresponding to hcp-Mg, suggesting a complete
decomposition of MgH2 phase.
El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734 13733

4. Conclusion

Based on the results of the present study the conclusions derived are the following:
(1) The the mole fraction of rutile-structure -MgH2 phase increased with increases the RBM time. A single -
MgH2 phase was obtained after 25 h of the RBM time,
(2) The obtained -MgH2 phase is no longer withstand against the lattice imperfections introduced by the
milling media and hence transformed into metastable -MgH2 phase upon increasing the RBM time. Cyclic 
phase transformations were observed several times upon changing the milling time. We would propose that the
formation enthalpy of the metastable -MgH2 phase is comparable to the -MgH2 phase and the energy barrier
between these two phases is rather low to allow such cyclic phase transformations,
(3) Long RBM time was investigated to decrease the grain sizes of MgH2 and improve the kinetics of
hydrogen absorption/desorption,
(4) After 200 h of RBM time, the decomposition temperature was 399oC. Moreover, the times required for
complete absorption and desorption of 7 wt.% of hydrogen at 250oC were recorded to be 3140 s and 35207 s under
10 bar and 200mbar, respectively,
(5) At 300 oC, the powders that were obtained upon RBM time for 200 h possess excellent hydrogenation
properties for any pure MgH2 system, indexed by high hydrogen storage capacity (7.54 wt.%) with complete 600
absorption/desorption cycles.

Acknowledgements

Appreciation is extended to the Kuwait Foundation for the Advancement of Sciences (KFAS) for the partial
financial support of this study related to the Project EA061C under a contract number: P315-35EC-01. The financial
support received by the Kuwait Government through the Kuwait Institute for Scientific Research for purchasing the
equipment used in the present work, using the budget dedicated for the project led by the first author (P-KISR-06-
04) of Establishing Nanotechnology Center in KISR is highly appreciated

References

[1] A. Calka, Appl. Phys. Lett. 59 (1991) 1568-1570.


[2] M Sherif El-Eskandarany, K. Sumiyama, K. Aoki, K. Suzuki, Mater. Sci. Forum., 88 (1992) 801-808.
[3] T. Sadhasivama, M. Hudson, Leo Sterlin, K. Sunita Pandey, Ashish Bhatnagar, K. Milind, K. Gurunathan, O. Srivastav: Int. J. of Hydrogen
Energy 38 (2013) 7353-7362.
[4] A. Varin Robert, Tomasz Czujko, S. Wronski Zbigniew, Nanomaterials for Solid State Hydrogen Storage. first ed., New York, Springer,
2009, pp. 12.
[5] M. Sherif El-Eskandarany, H.S. AlMatrouk, Ehab Shaban: Materials Today: Proceedings 3 (2016) 2735–2743.
[6] H. Shao, G. Xin, J. Zheng, X. Li, E. Akiba, Nano Energy 1 (2012) 590–601.
[7] M. Sherif El-Eskandarany, Sci. Rep. 6, 26936; doi: 10.1038/srep26936 (2016).
[8] M. Sherif El-Eskandarany, Ehab Shaban, Badryiah Al-Halaili, Int. J Hydrogen Energy 39 (2014) 12727-12740.
13734 El-Eskandarany et al. / Materials Today: Proceedings 5 (2018) 13726–13734

[6] J. Huot, D. B. Ravnsbæk, J. Zhang, F. Cuevas, M. Latroche, T. R. Jensen, Prog. Mater. Sci. 58 (2013) 30-75.
[5] A. Varin Robert, Tomasz Czujko, S. Wronski Zbigniew, Nanomaterials for Solid State Hydrogen Storage. First ed., New York, Springer,
2009, pp. 22.
[6] C. X .Shang, M. Bououdina, Y. Song, Z.X. Guo, Int. J. Hydrogen Energy 29 (2004) 73–80.
[7] M. Porcu, A. K. Petford-Long, J M. Sykes, J. Alloys. Compd. 29 (2005) 341–346.
[8] G. Barkhordarian, T. Klassen, R. Bormann, J. Alloys Compd. 407 (2006) 249–55.
[9] N. Hanada, T. Ichikawa, S. I. Orimo, H. Fujii, J Alloys Compd 366 (2004) 269–273.
[10] R. Floriano, D.R. Leiva, J.A. Carvalho, T.T. Ishikawa, W.J. Botta, J. Hydrogen Energy 39 (2014) 4959-65.
[11] Hui Wu, et al. Nanotechnology 2009; 20: doi:10.1088/0957-4484/20/20/204002.

Das könnte Ihnen auch gefallen