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 11 毫秒
1.
In this work, the hydriding–dehydriding properties of the LiBH4–NbF5 mixtures were investigated. It was found that the dehydrogenation and reversibility properties of LiBH4 were significantly improved by NbF5. Temperature-programed dehydrogenation (TPD) showed that 5LiBH4–NbF5 sample started releasing hydrogen from as low as 60 °C, and 4 wt.% hydrogen could be obtained below 255 °C. Meanwhile, ∼7 wt.% H2 could be reached at 400 °C in 20LiBH4–NbF5 sample, whereas pristine LiBH4 only released ∼0.7 wt.% H2. In addition, reversibility measurement demonstrated that over 4.4 wt.% H2 could still be released even during the fifth dehydrogenation in 20LiBH4–NbF5 sample. The experimental results suggested that a new borohydride possibly formed during ball milling the LiBH4–NbF5 mixtures might be the source of the active effect of NbF5 on LiBH4.  相似文献   

2.
The effect of NbF5 on the hydrogen sorption performance of NaAlH4 has been investigated. It was found that the dehydrogenation/hydrogenation properties of NaAlH4 were significantly enhanced by mechanically milling with 3 mol% NbF5. Differential scanning calorimetry results indicate that the ball-milled NaAlH4-0.03NbF5 sample lowered the completion temperature for the first two steps dehydrogenation by 71 °C compared to the pristine NaAlH4 sample. Isothermal hydrogen sorption measurements also revealed a significant enhancement in terms of the sorption rate and capacity, in particular, at reduced operation temperatures. The apparent activation energy for the first-step and the second-step dehydrogenation of the NaAlH4-0.03NbF5 sample is estimated to be 88.2 kJ/mol and 102.9 kJ/mol, respectively, by using Kissinger’s approach, which is much lower than for pristine NaAlH4, indicating the reduced kinetic barrier. The rehydrogenation kinetics of NaAlH4 was also improved with 3 mol% NbF5 doping, absorbing ∼1.7 wt% hydrogen at 150 °C for 2 h under ∼5.5 MPa hydrogen pressure. In contrast, no hydrogen was absorbed by the pristine NaAlH4 sample under the same conditions. The formation of Na3AlH6 was detected by X-ray diffraction on the rehydrogenated NaAlH4-0.03NbF5 sample. Furthermore, the structural changes in the NbF5-doped NaAlH4 sample after ball milling and the hydrogen sorption were carefully examined, and the active species and mechanism of catalysis in NbF5-doped NaAlH4 are discussed.  相似文献   

3.
We investigated the effects of NbF5 addition by ball milling on the hydrogen storage properties of LiAlH4. Pressure-composition-temperature (PCT) experiments showed that addition of 0.5 and 1 mol% NbF5 in LiAlH4 improves the onset desorption temperature and results in little decrease in hydrogen capacity, with approximately 7.0 wt% released by 188 °C. Isothermal dehydriding kinetics measurements indicated that the NbF5-doped sample shows an average dehydrogenation rate 5–6 times faster than that of the as-received LiAlH4 sample. In the x-ray diffraction results, there are distinct peaks of Al and LiH that appear after desorption. There is no peak of NbF5 before or after desorption. Desorption kinetics measurements indicated that the activation energy, EA, for LiAlH4 + 1 mol% NbF5 is about 67 kJ/mol for first reaction stage and about 77 kJ/mol for second reaction stage. The desorption process was further characterised by differential scanning calorimetry, and the possible mechanism of the effects of NbF5 addition is discussed.  相似文献   

4.
This paper presents a comparative study of H2 absorption and desorption in MgH2 milled with NbF5 or NbH0.9. The addition of NbF5 or NbH0.9 greatly improves hydriding and dehydriding kinetics. After 80 h of milling the mixture of MgH2 with 7 mol.% of NbF5 absorbs 60% of its hydrogen capacity at 250 °C in 30 s, whereas the mixture with 7 mol.% of NbH0.9 takes up 48%, and MgH2 milled without additive only absorbs 2%. At the same temperature, hydrogen desorption in the mixture with NbF5 finishes in 10 min, whereas the mixture with NbH0.9 only desorbs 50% of its hydrogen content, and MgH2 without additive practically does not releases hydrogen. The kinetic improvement is attributed to NbH0.9, a phase observed in the hydrogen cycled MgH2 + NbF5 and MgH2 + NbH0.9 materials, either hydrided or dehydrided. The better kinetic performance of the NbF5-added material is attributed to the combination of smaller size and enhanced distribution of NbH0.9 with more favorable microstructural characteristics. The addition of NbF5 also produces the formation of Mg(HxF1-x)2 solid solutions that limit the practically achievable hydrogen storage capacity of the material. These undesired effects are discussed.  相似文献   

5.
Mg (200 nm) and LaNi5 (25 nm) nanoparticles were produced by the hydrogen plasma-metal reaction (HPMR) method, respectively. Mg–5 wt.% LaNi5 nanocomposite was prepared by mixing these nanoparticles ultrasonically. During the hydrogenation/dehydrogenation cycle, Mg–LaNi5 transformed into Mg–Mg2Ni–LaH3 nanocomposite. Mg particles broke into smaller particles of about 80 nm due to the formation of Mg2Ni. The nanocomposite showed superior hydrogen sorption kinetics. It could absorb 3.5 wt.% H2 in less than 5 min at 473 K, and the storage capacity was as high as 6.7 wt.% at 673 K. The nanocomposite could release 5.8 wt.% H2 in less than 10 min at 623 K and 3.0 wt.% H2 in 16 min at 573 K. The apparent activation energy for hydrogenation was calculated to be 26.3 kJ mol−1. The high sorption kinetics was explained by the nanostructure, catalysis of Mg2Ni and LaH3 nanoparticles, and the size reduction effect of Mg2Ni formation.  相似文献   

6.
Here we report the first investigation of the dehydriding and re-hydriding properties of 2LiBH4 + MgH2 mixtures in the solid state. Such a study is made possible by high-energy ball milling of 2LiBH4 + MgH2 mixtures at liquid nitrogen temperature with the addition of graphite. The 2LiBH4 + MgH2 mixture ball milled under this condition exhibits a 5-fold increase in the released hydrogen at 265 °C when compared with ineffectively ball milled counterparts. Furthermore, both LiBH4 and MgH2 contribute to hydrogen release in the solid state. The isothermal dehydriding/re-hydriding cycles at 265 °C reveal that re-hydriding is dominated by re-hydriding of Mg. These unusual phenomena are explained based on the formation of nanocrystalline and amorphous phases, the increased defect concentration in crystalline compounds, and possible catalytic effects of Mg, MgH2 and LiBH4 on their dehydriding and re-hydriding properties.  相似文献   

7.
Mg-5wt%Ni-2.5wt%Fe-2.5wt%V (named Mg-5Ni-2.5Fe-2.5V) powder was prepared by reactive mechanical grinding using a planetary ball mill. The activation process, the changes in phase and microstructure with hydriding-dehydriding cycling, and the variations in the hydriding and dehydriding rates with temperature were investigated. The rate-controlling step for the dehydriding reaction of Mg-5Ni-2.5Fe-2.5V was analyzed by using a spherical moving boundary model. As the temperature increased from 473 K through 623 K, the initial hydrogen absorption rate under 12 bar H2 decreased, while the hydrogen desorption rate under 1.0 bar H2 increased.  相似文献   

8.
The catalytic mechanism of Nb2O5 and NbF5 on the dehydriding property of Mg95Ni5 prepared by hydriding combustion synthesis and mechanical milling (HCS + MM) was studied. It was shown that NbF5 was more efficient than Nb2O5 in improving the dehydriding property. In particular, the dehydriding temperature onset decreases from 460 K for Mg95Ni5 to 450 K for Mg95Ni5with 2.0 at.% Nb2O5, whereas it decreases to 410 K for that with 2.0 at.% NbF5. By means of X-ray diffraction and X-ray photoelectron spectroscopy, it was confirmed that the interaction between the Nb ions and the H atoms and that between the anions (O2− or F) and Mg2+ existed in Mg95Ni5 doped with Nb2O5 or NbF5. Further, the pressure–concentration-isotherms analysis clarified that these interactions destabilized the Mg–H bonding, and that NbF5 had a better effect on the destabilization of the Mg–H bonding than Nb2O5 contributing to the better dehydriding property of (Mg95Ni5)2.0−NbF5.  相似文献   

9.
Nano-structured Cr2O3 powder could be produced by spray conversion (spray drying of an aqueous of Cr nitrate, oxidation and milling).The samples Mg—10 wt% Cr2O3 using nano-structured Cr2O3 synthesized by spray conversion were prepared by mechanical grinding under H2 (reactive mechanical grinding) under the optimum conditions, previously studied, for the preparation of the sample Mg—10 wt% Fe2O3 using purchased Fe2O3. The sample Mg—10 wt% Cr2O3 as milled absorbed 4.48 wt% H2 at 593 K under 12 bar H2 for 60 min. Its activation was accomplished after two hydriding–dehydriding cycles. The activated sample absorbed 5.48 wt% H2 for 10 min and 5.93 wt% H2 for 60 min at 593 K, 12 bar H2, and desorbed 3.65 wt% H2 at 603 K, 1.0 bar H2 for 60 min. H2-storage capacity was 6.38 wt% under 12 bar H2 at 593 K (from P–C-T curve). Reactive mechanical grinding of Mg with Cr2O3 by spray conversion increased the hydriding rate effectively but increased a little the dehydriding rate, compared with reactive mechanical grinding of Mg with Fe2O3 purchased, Fe2O3 by spray conversion, MnO and SiO2 by spray conversion.  相似文献   

10.
The dehydrogenation properties of Mg(BH4)2 with various additives (SiO2, VCl3, CoCl2 and NbF5) were investigated. The addition of NbF5 significantly improved the extent of hydrogen release as well as the kinetics. While neat Mg(BH4)2 starts to release hydrogen >270 °C, Mg(BH4)2 with NbF5 begins hydrogen release ∼75 °C, as confirmed by mass spectrometry and thermogravimetry. The maximum hydrogen yield of Mg(BH4)2, obtained in the presence of 15 wt% NbF5, was 3.7, 7.4, 10.0, 11.4 wt% for 150, 250, 300 and 350 °C, respectively. Using pXRD, we confirmed that the final crystalline product at 300 °C from Mg(BH4)2 + 15 wt% NbF5 was Mg, while it was MgH2 for neat Mg(BH4)2. Solid state 11B NMR analysis of Mg(BH4)2 with 15 wt% NbF5 at 300 °C showed significant selectivity toward the formation of Mg(B12H12) as intermediate, while neat Mg(BH4)2 showed β-Mg(BH4)2, Mg(B2H6) as well as some Mg(B12H12). Our results demonstrate that NbF5 is a promising additive to provide high hydrogen yield values from Mg(BH4)2 at moderate temperatures <300 °C.  相似文献   

11.
The cyclic stability and high rate discharge performance of (La,Mg)5Ni19 multiphase alloy were investigated in this work. The results show that the alloy is composed of Pr5Co19-type (2H), Ce5Co19-type (3R), CaCu5-type and Ce2Ni7-type phase after annealing at 1123 K. The total phase abundance of Pr5Co19-type (2H) and Ce5Co19-type (3R) is 63.3%. That composition decides the good cyclic stability both in repeated hydriding/dehydriding and charging/discharging process for this alloy. Moreover, the alloy shows the higher high rate discharge capacity at room temperature and it remains 140 mA h/g at the current density of 3600 mA/g.  相似文献   

12.
In this research, the effect of NbF5 as an additive on the hydrogen desorption kinetics of MgH2 was investigated and compared to TiH2, Mg2Ni and Nb2O5 catalysts. The kinetics measurements were done using a method in which the ratio of the equilibrium plateau pressure to the opposing pressure was the same for all the reactions. The data showed NbF5 to be vastly superior to the other catalysts for improving the desorption kinetics of MgH2. The rates of desorption were found to be in the order NbF5 ? Nb2O5 > Mg2Ni > TiH2 > Pure MgH2. Kinetic modeling measurements showed that chemical reaction at the phase boundary to be the likely process controlling the reaction rates. TPD analyses showed the mixture with NbF5 has the lowest desorption temperature although it was accompanied with some weight penalty.  相似文献   

13.
Among samples of Mg-Ni, Mg-Ni-5Fe2O3, and Mg-Ni-5Fe, Mg-Ni-5Fe had the highest hydriding and dehydriding rates. For the as-milled Mg-Ni-5Fe alloy and the hydrided Mg-Ni-5Fe alloy after activation, the weight percentages of the constituent phases were calculated using the FullProf program. The creation of defects and the diminution of Mg particle size through reactive mechanical grinding and hydriding-dehydriding cycling, and the formation of the Mg2Ni phase are considered to increase the hydriding and dehydriding rates. Mg-14Ni-2Fe-2Ti-2Mo had higher hydriding and dehydriding rates than did any of the other samples (Mg-Ni, Mg-Ni-5Fe2O3, Mg-Ni-5Fe, and Mg-14Ni-6Fe2O3) prepared in this work.  相似文献   

14.
The effects of annealing at 1123, 1148, 1173 and 1198 K for 16 h on the structure and properties of the LaY2Ni10Mn0.5 hydrogen storage alloy as the active material of the negative electrode in nickel–metal hydride (Ni–MH) batteries were systematically investigated by X-ray diffraction (XRD), scanning electron microscopy linked with an energy dispersive X-ray spectrometer (SEM–EDS), pressure-composition isotherms (PCI) and electrochemical measurements. The quenched and annealed LaY2Ni10Mn0.5 alloys primarily consist of Ce2Ni7- (2H) and Gd2Co7-type (3R) phases. The homogeneity of the composition and plateau characteristics of the annealed alloys are significantly improved, and the lattice strain is effectively reduced. The alloys annealed at 1148 K and 1173 K have distinctly greater hydrogen storage amounts, 1.49 wt% (corresponding to 399 mAh g?1 in equivalent electrochemical units) and 1.48 wt%, respectively, than the quenched alloy (1.25 wt%, corresponding to 335 mAh g?1 in equivalent electrochemical units). The alloys annealed at 1148 K and 1173 K have relatively good activation capabilities. The annealing treatment slightly decreases the discharge potentials of the alloy electrodes but markedly increases their discharge capacity. The maximum discharge capacities of the annealed alloy electrodes (372–391 mAh g?1) are greater than the extreme capacity of the LaNi5-type alloy (370 mAh g?1). The cycling stability of the annealed alloy electrodes was improved.  相似文献   

15.
The Ti1.4V0.6Ni ribbon alloy and AB3-type (La0.65Nd0.12Mg0.23Ni2.9Al0.1) alloy ingot are prepared by melt-spinning technique and induction levitation melting technique, respectively. The Ti1.4V0.6Ni + 20 wt.% AB3 mixture powders are synthesized by ball-milling the above prepared alloy ingots, and their structures and the electrochemical hydrogen storage properties are investigated. It is found that the icosahedral quasicrystal, Ti2Ni, BCC structural solid solution and AB3-type phases are all presented in the composite material. The maximum electrochemical discharge capacity of the composite electrode is 294.7 mAh/g at the discharge current density of 30 mA/g and 303 K. In addition, the electrode made of Ti1.4V0.6Ni and AB3 composite holds better high-rate discharge ability than that of Ti1.4V0.6Ni.  相似文献   

16.
Though LiBH4-MgH2 system exhibits an excellent hydrogen storage property, it still presents high decomposition temperature over 350 °C and sluggish hydrogen absorption/desorption kinetics. In order to improve the hydrogen storage properties, the influence of MoCl3 as an additive on the hydrogenation and dehydrogenation properties of LiBH4-MgH2 system is investigated. The reversible hydrogen storage performance is significantly improved, which leads to a capacity of about 7 wt.% hydrogen at 300 °C. XRD analysis reveals that the metallic Mo is formed by the reaction between LiBH4 and MoCl3, which is highly dispersed in the sample and results in improved dehydrogenation and hydrogenation performance of LiBH4-MgH2 system. From Kissinger plot, the activation energy for hydrogen desorption of LiBH4-MgH2 system with additive MoCl3 is estimated to be ∼43 kJ mol−1 H2, 10 kJ mol−1 lower than that for the pure LiBH4-MgH2 system indicating that the kinetics of LiBH4-MgH2 composite is significantly improved by the introduction of Mo.  相似文献   

17.
18.
Here we present the development of an aluminium alloy based hydrogen storage tank, charged with Ti-doped sodium aluminium hexahydride Na3AlH6. This hydride has a theoretical hydrogen storage capacity of 3 mass-% and can be operated at lower pressure compared to sodium alanate NaAlH4. The tank was made of aluminium alloy EN AW 6082 T6. The heat transfer was realised through an oil flow in a bayonet heat exchanger, manufactured by extrusion moulding from aluminium alloy EN AW 6060 T6. Na3AlH6 is prepared from 4 mol-% TiCl3 doped sodium aluminium tetrahydride NaAlH4 by addition of two moles of sodium hydride NaH in ball milling process. The hydrogen storage tank was filled with 213 g of doped Na3AlH6 in dehydrogenated state. Maximum of 3.6 g (1.7 mass-% of the hydride mass) of hydrogen was released from the hydride at approximately 450 K and the same hydrogen mass was consumed at 2.5 MPa hydrogenation pressure. 45 cycle tests (rehydrogenation and dehydrogenation) were carried out without any failure of the tank or its components. Operation of the tank under real conditions indicated the possibility for applications with stationary HT-PEM fuel cell systems.  相似文献   

19.
Mg(AlH4)2 submicron rods with 96.1% purity have been successfully synthesized in a modified mechanochemical reaction process followed by Soxhlet extraction. ∼9.0 wt% of hydrogen is released from the as-prepared Mg(AlH4)2 at 125–440 °C through a stepwise reaction. Upon dehydriding, Mg(AlH4)2 decomposes first to generate MgH2 and Al. Subsequently, the newly produced MgH2 reacts with Al to form a Al0.9Mg0.1 solid solution. Finally, further reaction between the Al0.9Mg0.1 solid solution and the remaining MgH2 gives rise to the formation of Al3Mg2. The first step dehydrogenation is a diffusion-controlled reaction with an apparent activation energy of ∼123.0 kJ/mol. Therefore, increasing the mobility of the species involved in Mg(AlH4)2 will be very helpful for improving its dehydrogenation kinetics.  相似文献   

20.
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   

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