首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 250 毫秒
1.
A porous MgH2/C composite can be synthesized through decomposition of an organo-magnesium precursor under hydrogen pressure. XRD patterns of the porous MgH2/C composite exhibit a pure MgH2 phase with a tetragonal structure. The morphology of the resulted samples is significantly dependent on the synthesis temperature and hydrogen pressure. The samples exhibit a rod-like structure and composed of nano-crystallites of MgH2 with a size of less than 5 nm. TPD spectra of a sample synthesized at 220 °C for 4 h show a remarkable decrease of the onset hydrogen release temperature. Further, this sample also exhibits fast hydrogen adsorption kinetics adsorbing 6 wt % of hydrogen in 3 min at 250 °C. The same amount of hydrogen is adsorbed in 11 min at 200 °C and 40 min at 150 °C, respectively. N2 physisorption measurements of this sample indicate meso-porosity with a BET surface area of 40.9 m2 g−1 and an average pore diameter of 20 nm.  相似文献   

2.
The effect of Ti0.4Cr0.15Mn0.15V0.3 (termed BCC due to the body centered cubic structure) alloy on the hydrogen storage properties of MgH2 was investigated. It was found that the hydrogenated BCC alloy showed superior catalysis properties compared to the quenched and ingot samples. As an example, the 1 h milled MgH2 + 20 wt.% hydrogenated BCC shows a peak temperature of dehydrogenation of about 294 °C. This is 16, 27 and 74 °C lower than those of MgH2 ball milled with quenched BCC, ingot BCC and an uncatalysed MgH2 sample, respectively. The hydrogenated BCC alloy is much easier to crush into small particles, and embed in MgH2 aggregates as revealed by X-ray diffraction and scanning electron microscope results. The BCC not only increases the hydrogen atomic diffusivity in the bulk Mg but also promotes the dissociation and recombination of hydrogen. The activation energy, Ea, for the dehydrogenation of the MgH2/hydrogenated BCC mixture was found to be 71.2 ± 5 kJ mol H2−1 using the Kissinger method. This represents a significant decrease compared to the pure MgH2 (179.7 ± 5 kJ mol H2−1), suggesting that the catalytic effect of the BCC alloy significantly decreases the activation energy of MgH2 for dehydrogenation by surface activation.  相似文献   

3.
Both kinetics and thermodynamics properties of MgH2 are significantly improved by the addition of Mg(AlH4)2. The as-prepared MgH2–Mg(AlH4)2 composite displays superior hydrogen desorption performances, which starts to desorb hydrogen at 90 °C, and a total amount of 7.76 wt% hydrogen is released during its decomposition. The enthalpy of MgH2-relevant desorption is 32.3 kJ mol−1 H2 in the MgH2–Mg(AlH4)2 composite, obviously decreased than that of pure MgH2. The dehydriding mechanism of MgH2–Mg(AlH4)2 composite is systematically investigated by using x-ray diffraction and differential scanning calorimetry. Firstly, Mg(AlH4)2 decomposes and produces active Al. Subsequently, the in-situ formed Al reacts with MgH2 and forms Mg–Al alloys. For its reversibility, the products are fully re-hydrogenated into MgH2 and Al, under 3 MPa H2 pressure at 300 °C for 5 h.  相似文献   

4.
This paper reports the catalytic effects of mischmetal (Mm) and mischmetal oxide (Mm-oxide) on improving the dehydrogenation and rehydrogenation behaviour of magnesium hydride (MgH2). It has been found that 5 wt.% is the optimum catalyst (Mm/Mm-oxide) concentration for MgH2. The Mm and Mm-oxide catalyzed MgH2 exhibits hydrogen desorption at significantly lower temperature and also fast rehydrogenation kinetics compared to ball-milled MgH2 under identical conditions of temperature and pressure. The onset desorption temperature for MgH2 catalyzed with Mm and Mm-oxide are 323 °C and 305 °C, respectively. Whereas the onset desorption temperature for the ball-milled MgH2 is 381 °C. Thus, there is a lowering of onset desorption temperature by 58 °C for Mm and by 76 °C for Mm-oxide. The dehydrogenation activation energy of Mm-oxide catalyzed MgH2 is 66 kJ/mol. It is 35 kJ/mol lower than ball-milled MgH2. Additionally, the Mm-oxide catalyzed dehydrogenated Mg exhibits faster rehydrogenation kinetics. It has been noticed that in the first 10 min, the Mm-oxide catalyzed Mg (dehydrogenated MgH2) has absorbed up to 4.75 wt.% H2 at 315 °C under 15 atmosphere hydrogen pressure. The activation energy determined for the rehydrogenation of Mm-oxide catalyzed Mg is ∼62 kJ/mol, whereas that for the ball-milled Mg alone is ∼91 kJ/mol. Thus, there is a decrease in absorption activation energy by ∼29 kJ/mol for the Mm-oxide catalyzed Mg. In addition, Mm-oxide is the native mixture of CeO2 and La2O3 which makes the duo a better catalyst than CeO2, which is known to be an effective catalyst for MgH2. This takes place due to the synergistic effect of CeO2 and La2O3. It can thus be said that Mm-oxide is an effective catalyst for improving the hydrogen sorption behaviour of MgH2.  相似文献   

5.
MgH2 is one of the most promising hydrogen storage materials due to its high capacity and low cost. In an effort to develop MgH2 with a low dehydriding temperature and fast sorption kinetics, doping MgH2 with NiCl2 and CoCl2 has been investigated in this paper. Both the dehydrogenation temperature and the absorption/desorption kinetics have been improved by adding either NiCl2 or CoCl2, and a significant enhancement was obtained in the case of the NiCl2 doped sample. For example, a hydrogen absorption capacity of 5.17 wt% was reached at 300 °C in 60 s for the MgH2/NiCl2 sample. In contrast, the ball-milled MgH2 just absorbed 3.51 wt% hydrogen at 300 °C in 400 s. An activation energy of 102.6 kJ/mol for the MgH2/NiCl2 sample has been obtained from the desorption data, 18.7 kJ/mol and 55.9 kJ/mol smaller than those of the MgH2/CoCl2, which also exhibits an enhanced kinetics, and of the pure MgH2 sample, respectively. In addition, the enhanced kinetics was observed to persist even after 9 cycles in the case of the NiCl2 doped MgH2 sample. Further kinetic investigation indicated that the hydrogen desorption from the milled MgH2 is controlled by a slow, random nucleation and growth process, which is transformed into two-dimensional growth after NiCl2 or CoCl2 doping, suggesting that the additives reduced the barrier and lowered the driving forces for nucleation.  相似文献   

6.
To improve hydrogen desorption properties of magnesium hydride, a composite material with composition of MgH2-5 at% Ni3FeMn has been prepared by co-milling MgH2 powder with Ni3FeMn alloy either in the form of as-cast (sample A) or melt-spun ribbon (sample B). The study has shown that the addition of Ni3FeMn alloy to magnesium hydride can yield a finer particle size after mechanical alloying (MA). As a consequence, the desorption temperature of mechanically activated MgH2 for 30 h has decreased from 319 °C to 307 °C for sample A and to 290 °C for sample B. Furthermore, some favorable effects of Ni3FeMn alloy on hydrogen desorption kinetics have been observed. Further improvement in the hydrogen desorption of melt-spun containing composite can be related to higher hardness value of the melt-spun powder compared to the as-cast alloy, and probably a more homogeneous distribution of the alloyed elements.  相似文献   

7.
In this paper, amorphous NiB nanoparticles were fabricated by chemical reduction method and the effect of NiB nanoparticles on hydrogen desorption properties of MgH2 was investigated. Measurements using temperature-programmed desorption system (TPD) and volumetric pressure–composition isotherm (PCI) revealed that both the desorption temperature and desorption kinetics have been improved by adding 10 wt% amorphous NiB. For example, the MgH2–10 wt%NiB mixture started to release hydrogen at 180 °C, whereas it had to heat up to 300 °C to release hydrogen for the pure MgH2. In addition, a hydrogen desorption capacity of 6.0wt% was reached within 10 min at 300 °C for the MgH2–10 wt%NiB mixture, in contrast, even after 120 min only 2.0 wt% hydrogen was desorbed for pure MgH2 under the same conditions. An activation energy of 59.7 kJ/mol for the MgH2/NiB composite has been obtained from the desorption data, which exhibits an enhanced kinetics possibly due to the additives reduced the barrier and lowered the driving forces for nucleation. Further cyclic kinetics investigation using high-pressure differential scanning calorimetry technique (HP-DSC) indicated that the composite had good cycle stability.  相似文献   

8.
A systematic investigation was performed on the hydrogen storage behaviors of ball-milled MgH2-activated carbon (AC) composites. Differential Scanning Calorimetry (DSC) measurement on the desorption temperature was carried out and indicated that the onset and peak temperatures both decreased with increasing AC adding amount, for example, the desorption peak temperature shifted from 349 °C for 1 wt% AC to 316 °C for 20 wt% AC. Furthermore, it is noted that the hydrogen absorption capacity and hydriding kinetics of the composites were also dependent on the adding amount of AC, and the optimum condition could be achieved by mechanical milling of MgH2 with 5 wt% AC. The Mg-5wt%AC composite can absorb about 6.5 wt% hydrogen within 7 min at 300 °C and 6.7 wt% within 2 h at 200 °C, respectively. It is also demonstrated that MgH2-5wt% AC exhibited good hydrogen desorption property that could release 6.5 wt% at 330 °C within 30 min. X-ray diffraction patterns (XRD) and transmission electron microscopy (TEM) observations revealed that the grain size of the synthesized composites decreased with increasing AC amount. This may contribute to the improvement of hydrogen storage in MgH2-AC composites.  相似文献   

9.
In the present work, the hydrogen storage properties of MgH2-X wt.% FeCl3 (X = 5, 10, 15 and 20) are investigated experimentally. It is found that the MgH2 + 10 wt.% FeCl3 sample exhibits the best comprehensive hydrogen storage properties, in terms of the onset dehydrogenation temperature, the hydrogen amounts de/reabsorbed as well as the relative de/rehydrogenation rates. The onset dehydrogenation temperature of the 10 wt.% FeCl3-doped MgH2 sample is reduced by about 90 °C compared to the as-milled MgH2, and the sorption kinetics measurements indicate that the FeCl3-doped sample displays an average dehydrogenation rate 5–6 times faster than that of the undoped MgH2 sample. Higher levels of doping introduce negative effects, such as lower capacity and slower absorption/desorption rates compared to samples with lower FeCl3 doping levels. The apparent activation energy for hydrogen desorption is decreased from 166 kJ•mol−1 for as-milled MgH2 to 130 kJ•mol−1 by the addition of 10 wt.% FeCl3. It is believed that the improvement of the MgH2 sorption properties in the MgH2/FeCl3 composite is due to the catalytic effects of the in-situ generated Fe species and MgCl2 that are formed during the heating process.  相似文献   

10.
Scandium(II)hydride, ScH2, and scandium(III)chloride, ScCl3, are explored as additives to facilitate hydrogen release and uptake for magnesium hydride. These additives are expected to form more homogeneous composites with Mg/MgH2 as compared to metallic scandium. However, scandium(III)chloride, reacts with MgH2 during mechano-chemical treatment and form ScH2 and MgCl2 (that later crystallise during heat treatment). The composite MgH2−ScH2 was investigated using in-situ synchrotron radiation powder X-ray diffraction during up to five cycles of continuous release and uptake of hydrogen at isothermal conditions at 320, 400 and 450 °C and p(H2) = 100–150 or 10−2 bar. The data were analysed by Rietveld refinement and no reaction is observed between either MgH2/ScH2 or Mg/ScH2 during cycling. The extracted sigmoidal shaped curves for formation or decomposition of Mg/MgH2 suggest that a nucleation process is preceding the crystal growth. The reaction rate increases with increasing number of cycles of hydrogen release and uptake at isothermal conditions possibly due to activation effects. This kinetic enhancement is strongest between the first cycles and may be denoted an activation effect.  相似文献   

11.
The intermetallic compound Mg0.65Sc0.35 was found to form a nano-structured metal hydride composite system after a (de)hydrogenation cycle at temperatures up to 350 °C. Upon dehydrogenation phase separation occurred forming Mg-rich and Sc-rich hydride phases that were clearly observed by SEM and TEM with the Sc-rich hydride phase distributed within Mg/MgH2-rich phase as nano-clusters ranging in size from 40 to 100 nm. The intermetallic compound Mg0.65Sc0.35 showed good hydrogen uptake, ca. 6.4 wt.%, in the first charging cycle at 150 °C and in the following (de)hydrogenation cycles, a reversible hydrogen capacity (up to 4.3 wt.%) was achieved. Compared to the as-received MgH2, the composite had faster cycling kinetics with a significant reduction in activation energy Ea from 159 ± 1 kJ mol−1 to 82 ± 1 kJ mol−1 (as determined from a Kissinger plot). Two-dehydrogenation events were observed by DSC and pressure–composition-isotherm (PCI) measurements, with the main dehydrogenation event being attributed to the Mg-rich hydride phase. Furthermore, after the initial two cycles the hydrogen storage capacity remained unchanged over the next 55 (de)hydrogenation cycles.  相似文献   

12.
The MgH2 + 0.02Ti-additive system (additives = 35 nm Ti, 50 nm TiB2, 40 nm TiC, <5 nm TiN, 10 × 40 nm TiO2) has been studied by high-resolution synchrotron X-ray diffraction, after planetary milling and hydrogen (H) cycling. TiB2 and TiN nanoparticles were synthesised mechanochemically whilst other additives were commercially available. The absorption kinetics and temperature programmed desorption (TPD) profiles have been determined, and compared to the benchmark system MgH2 + 0.01Nb2O5 (20 nm). TiC and TiN retain their structures after milling and H cycling. The TiB2 reflections appear compressed in d-spacing, suggesting Mg/Ti exchange has occurred in the TiB2 structure. TiO2 is reduced, commensurate with the formation of MgO, however, the Ti is not evident anywhere in the diffraction pattern. The 35 nm Ti initially forms an fcc Mg47.5Ti52.5 phase during milling, which then phase separates and hydrides to TiH2 and MgH2. At 300 °C, the MgH2 + 0.02 (Ti, TiB2, TiC, TiN, TiO2) samples display equivalent absorption kinetics, which are slightly faster than the MgH2 + 0.01Nb2O5 (20 nm) benchmark. All samples are contaminated with MgO from the use of a ZrO2 vial, and display rapid absorption to ca. 90% of capacity within 20 s at 300 °C. TPD profiles of all samples show peak decreases compared to the pure MgH2 milled sample, with many peak profiles displaying bi-modal splitting. TPD measurements on two separate instruments demonstrate that on a 30 min milling time scale, all samples are highly inhomogenous, and samplings from the exact same batch of milled MgH2 + 0.02Ti-additive can display differences in TPD profiles of up to 30 °C in peak maxima. The most efficient Ti based additive cannot be discerned on this basis, and milling times ? 30 min are necessary to obtain homogenous samples, which may lead to artefactual benefits, such as reduction in diffusion distances by powder grinding or formation of dense microstructure. For the hydrogen cycled MgH2 + 0.01Nb2O5 system, we observe a face centred cubic Mg/Nb exchanged Mg0.165Nb0.835O phase, which accounts for ca. 60% of the originally added Nb atoms.  相似文献   

13.
Nanostructured MgH2/0.1TiH2 composite was synthesized directly from Mg and Ti metal by ball milling under an initial hydrogen pressure of 30 MPa. The synthesized composite shows interesting hydrogen storage properties. The desorption temperature is more than 100 °C lower compared to commercial MgH2 from TG-DSC measurements. After desorption, the composite sample absorbs hydrogen at 100 °C to a capacity of 4 mass% in 4 h and may even absorb hydrogen at 40 °C. The improved properties are due to the catalyst and nanostructure introduced during high pressure ball milling. From the PCI results at 269, 280, 289 and 301 °C, the enthalpy change and entropy change during the desorption can be determined according to the van’t Hoff equation. The values for the MgH2/0.1TiH2 nano-composite system are 77.4 kJ mol−1 H2 and 137.5 J K−1 mol−1 H2, respectively. These values are in agreement with those obtained for a commercial MgH2 system measured under the same conditions. Nanostructure and catalyst may greatly improve the kinetics, but do not change the thermodynamics of the materials.  相似文献   

14.
Nanocrystalline Mg films with thicknesses between 45 and 900 nm were prepared by e-beam on fused-SiO2 substrates and hydrogenated at 280 °C to investigate the H-absorption/desorption process. Films were characterized by XRD, RBS, Raman, FEG, “in situ” optical measurements and TPD-MS. Whereas practically full conversion into MgH2 is observed in thinner films (d < 150–200 nm), higher amount of hydrogen is not absorbed by thicker films (d > 200–250 nm) that is attributed to the formation of Mg2Si–MgO phases (observed by RBS and Raman) as well as the slow kinetics of MgH2 formation. H-desorption process is controlled by a nucleation and growth process and hydrogen is released at lower desorption temperatures (Td = 425 °C) than bulk MgH2. Td are slightly lower (ΔT ∼ 25 °C) in thickest hydrogenated films (d > 200–250 nm) suggesting an influence of Mg2Si and MgO phases, formed during hydrogenation.  相似文献   

15.
The various Mg–B–Al–H systems composed of Mg(BH4)2 and different Al-sources (metallic Al, LiAlH4 and its decomposition products) have been investigated as potential hydrogen storage materials. The role of Al was studied on the dehydrogenation and the rehydrogenation of the systems. The results indicate that the different Al-sources exhibit a similar improving effect on the dehydrogenation properties of Mg(BH4)2. Taking the Mg(BH4)2 + LiAlH4 system as an example, at first LiAlH4 rapidly decomposes into LiH and Al, then Mg(BH4)2 decomposes into MgH2 and B, finally the MgH2 reacts with Al, LiH and B, which forms Mg3Al2 and MgAlB4. The system starts to desorb H2 at 140 °C and desorbs 3.6 wt.% H2 below 300 °C, while individual Mg(BH4)2 starts to desorb H2 at 250 °C and desorbs only 1.3 wt.% H2 below 300 °C. The isothermal desorption kinetics of the Mg–B–Al–H systems is about 40% faster than that of Mg(BH4)2 at the hydrogen desorption ratio of 90%. In addition, the Mg–B–Al–H systems show partial reversibility at moderate temperature and pressure. For Al-added system, the product of rehydrogenation is MgH2, while for LiAlH4-added system the product is composed of LiBH4 and MgH2.  相似文献   

16.
Pure magnesium powders were ball milled under a hydrogen pressure of 50 bar at room temperature, using reactive ball milling (RBM) approach. The results have shown that a single stable phase of β-MgH2 is obtained upon RBM for 25 h. Increasing the RBM time leads to a significant decreasing on the grain size and an increase in the iron contamination that were introduced to the powders upon using hard steel milling tools. Remarkable changes in the transformed mass fractions of β-MgH2 phase to a metastable γ-MgH2 phase with increasing the RBM time could be detected. Cyclic β-γ-β phase transformations were observed several times upon changing the RBM time. After 200 h of RBM time, the decomposition temperature and activation energy were recorded to be 399 °C and 131 kJ/mol, respectively. Moreover, the times required for complete absorption and desorption of 7 wt.% of hydrogen at 250 °C were recorded to be 3140 s and 35,207 s under 10 and 0 bar, respectively. At 300 °C, 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. Improvements of hydrogenation/dehydrogenation kinetics are attributed to the presence of γ-phase, the existence of Fe contamination and the nanocrystallinity of the ball milled powders.  相似文献   

17.
In the present work, high-energy mechanical alloying (MA) was employed to synthesize a nanostructured magnesium-based composite for hydrogen storage. The preparation of the composite material with composition of MgH2-5 at% (TiCr1.2Fe0.6) was performed by co-milling of commercial available MgH2 powder with the body-centered cubic (bcc) alloy either in the form of Ti-Cr-Fe powder mixture with the proper mass fraction (sample A) or prealloyed TiCr1.2Fe0.6 powder (sample B). The prealloyed powder with an average crystallite size of 14 nm and particle size of 384 nm was prepared by the mechanical alloying process. It is shown that the addition of the Ti-based bcc alloy to magnesium hydride yields a finer particle size and grain structure after mechanical alloying. As a result, the desorption temperature of mechanically activated MgH2 for 4 h decreased from 327 °C to 262 °C for sample A and 241 °C for sample B. A high dehydrogenation capacity (∼5 wt%) at 300 °C is also obtained. The effect of the Ti-based alloy on improvement of the dehydrogenation is discussed.  相似文献   

18.
In this paper, the hydrogen storage properties and reaction mechanism of the 4MgH2 + LiAlH4 composite system with the addition of Fe2O3 nanopowder were investigated. Temperature-programmed-desorption results show that the addition of 5 wt.% Fe2O3 to the 4MgH2 + LiAlH4 composite system improves the onset desorption temperature to 95 °C and 270 °C for the first two dehydrogenation stage, which is lower 40 °C and 10 °C than the undoped composite. The dehydrogenation and rehydrogenation kinetics of 5 wt.% Fe2O3-doped 4MgH2 + LiAlH4 composite were also improved significantly as compared to the undoped composite. Differential scanning calorimetry measurements indicate that the enthalpy change in the 4MgH2–LiAlH4 composite system was unaffected by the addition of Fe2O3 nanopowder. The Kissinger analysis demonstrated that the apparent activation energy of the 4MgH2 + LiAlH4 composite (125.6 kJ/mol) was reduced to 117.1 kJ/mol after doping with 5 wt.% Fe2O3. Meanwhile, the X-ray diffraction analysis shows the formation of a new phase of Li2Fe3O4 in the doped composite after the dehydrogenation and rehydrogenation process. It is believed that Li2Fe3O4 acts as an actual catalyst in the 4MgH2 + LiAlH4 + 5 wt.% Fe2O3 composite which may promote the interaction of MgH2 and LiAlH4 and thus accelerate the hydrogen sorption performance of the MgH2 + LiAlH4 composite system.  相似文献   

19.
The desorption mechanism of as-milled 2NaBH4 + MgH2 was investigated by volumetric analysis, X-ray diffraction and electron microscopy. Hydrogen desorption was carried out in 0.1 bar hydrogen pressure from room temperature up to 450 °C at a heating rate of 3 °C min−1. Complete dehydrogenation was achieved in two steps releasing 7.84 wt.% hydrogen. Desorption reaction in this system is kinetically restricted and limited by the growth of MgB2 at the Mg/Na2B12H12 interface where the intermediate product phases form a barrier to diffusion. During desorption, MgB2 particles are observed to grow as plates around NaH particles.  相似文献   

20.
The effect of lithium borohydride (LiBH4) on the hydriding/dehydriding kinetics and thermodynamics of magnesium hydride (MgH2) was investigated. It was found that LiBH4 played both positive and negative effects on the hydrogen sorption of MgH2. With 10 mol.% LiBH4 content, MgH2–10 mol.% LiBH4 had superior hydrogen absorption/desorption properties, which could absorb 6.8 wt.% H within 1300 s at 200 °C under 3 MPa H2 and completed desorption within 740 s at 350 °C. However, with the increasing amount of LiBH4, the hydrogenation/dehydrogenation kinetics deteriorated, and the starting desorption temperature increased and the hysteresis of the pressure-composition isotherm (PCI) became larger. Our results showed that the positive effect of LiBH4 was mainly attributed to the more uniform powder mixture with smaller particle size, while the negative effect of LiBH4 might be caused by the H–H exchange between LiBH4 and MgH2.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号