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 31 毫秒
1.
In the present work, the catalytic effect of TiF3 on the dehydrogenation properties of LiAlH4 has been investigated. Decomposition of LiAlH4 occurs during ball milling in the presence of 4 mol% TiF3. Different ball milling times have been used, from 0.5 h to 18 h. With ball milling time increasing, the crystallite sizes of LiAlH4 get smaller (from 69 nm to 43 nm) and the dehydrogenation temperature becomes lower (from 80 °C to 60 °C). Half an hour ball milling makes the initial dehydrogenation temperature of doped LiAlH4 reduce to 80 °C, which is 70 °C lower than as-received LiAlH4. About 5.0 wt.% H2 can be released from TiF3-doped LiAlH4 after 18 h ball milling in the range of 60 °C–145 °C (heating rate 2 °C min−1). TiF3 probably reacts with LiAlH4 to form the catalyst, TiAl3. The mechanochemical and thermochemical reactions have been clarified. However, the rehydrogenation of LiAlH4/Li3AlH6 can not be realized under 95 bar H2 in the presence of TiF3 because of their thermodynamic properties.  相似文献   

2.
In this work, we report the synthesis, characterization and destabilization of lithium aluminum hydride by ad-mixing nanocrystalline magnesium hydride (e.g. LiAlH4 + nanoMgH2). A new nanoparticulate complex hydride mixture (Li–nMg–Al–H) was obtained by solid-state mechano-chemical milling of the parent compounds at ambient temperature. Nanosized MgH2 is shown to have greater and improved hydrogen performance in terms of storage capacity, kinetics, and initial temperature of decomposition, over the commercial MgH2. The pressure–composition isotherms (PCI) reveal that the destabilized LiAlH4 + nanoMgH2 possess ∼5.0 wt.% H2 reversible capacity at T ≤ 350 °C. Van't Hoff calculations demonstrate that the destabilized (LiAlH4 + nanoMgH2) complex materials have comparable enthalpy of hydrogen release (∼85 kJ/mole H2) to their pristine counterparts, LiAlH4 and MgH2. However, these new destabilized complex hydrides exhibit reversible hydrogen sorption behavior with fast kinetics.  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.
The present paper reports the catalytic effect of carbon nanomaterials, particularly carbon nanotubes (CNTs) and graphitic nanofibres (GNFs) with two different structure morphology, namely planar GNFs (PGNFs) and helical GNFs (HGNFs) as the catalyst for improving the dehydrogenation and rehydrogenation behavior of sodium aluminum hydride (NaAlH4). It has been observed that HGNFs posses superior catalytic activity than other carbon nanoforms in improving the desorption kinetics and decreasing the desorption temperature of NaAlH4. Temperature programmed desorption (TPD) reveals that HGNFs admixed NaAlH4 undergo hydrogen desorption at a much lower temperature than PGNFs and CNTs (SWCNTs and MWCNTs) admixed NaAlH4. Thus for the heating rate of 2 °C/min, the peak desorption temperature corresponds to initial step decomposition of NaAlH4 admixed with 2 wt.% HGNFs and 2 wt.% PGNFs has been lowered to 143.6 °C and 152.6 °C, respectively (for pristine NaAlH4, it is ∼170 °C). In addition to the enhancement in desorption kinetics, the HGNFs admixed NaAlH4 undergoes fast rehydrogenation at the moderate condition. Microstructural investigation reveals that the HGNFs were present on the surface of NaAlH4 grains, whereas CNTs were tunneled into the grains of NaAlH4 suggesting a distinct catalytic behavior of different carbon nanovariants.  相似文献   

5.
The effect of carbon nanofibres (CNFs) on the de/re-hydrogenation characteristics of 1:2 magnesium amide (Mg(NH2)2) and lithium hydride (LiH) mixture is investigated. It is found that the desorption as well as absorption characteristic of the 1:2 Mg(NH2)2/LiH mixture is improved with admixing of different shaped (planar and helical) CNFs separately. The different shaped CNFs were synthesized through catalytic decomposition of acetylene gas over LaNi5 alloy. The synthesized CNFs contain Ni-metal nano particles. Among two different types of nanofibres namely planar carbon nanofibres (PCNFs) and helical carbon nanofibres (HCNFs), the later was found to act as a better catalyst. The decomposition temperature of the pristine Mg(NH2)2/LiH mixture is ∼250 °C, reduced to 150 and 140 °C for the PCNF and HCNF admixed Mg(NH2)2/LiH mixture respectively. The activation energy for dehydrogenation reaction was found to ∼97.2 kJ/mol, which is further reduced to ∼67 and ∼65 kJ/mol for the PCNF and HCNF admixed Mg(NH2)2/LiH mixture respectively. The lowering of decomposition temperature and enhancement in desorption kinetics, with admixing of different shaped CNFs are described and discussed.  相似文献   

6.
The effects of TiO2 nanopowder addition on the dehydrogenation behaviour of LiAlH4 have been studied. The 5 wt.% TiO2-added LiAlH4 sample showed a significant improvement in dehydrogenation rate compared to that of undoped LiAlH4, with the dehydrogenation temperature reduced from 150 °C to 60 °C. Kinetic desorption results show that the added LiAlH4 released about 5.2 wt% hydrogen within 30 min at 100 °C, while the as-received LiAlH4 just released below 0.2 wt.% hydrogen within same time at 120 °C. From the Arrhenius plot of the hydrogen desorption kinetics, the apparent activation energy is 114 kJ/mol for pure LiAlH4 and 49 kJ/mol for the 5 wt.% TiO2 added LiAlH4, indicating that TiO2 nanopowder adding significantly decreased the activation energy for hydrogen desorption of LiAlH4. X-ray diffraction and Fourier transform infrared analysis show that there is no phase change in the cell volume or on the Al-H bonds of the LiAlH4 due to admixture of TiO2 after milling. X-ray photoelectron spectroscopy results show no changes in the Ti 2p spectra for TiO2 after milling and after dehydrogenation. The improved dehydrogenation behaviour of LiAlH4 in the presence of TiO2 is believed to be due to the high defect density introduced at the surfaces of the TiO2 particles during the milling process.  相似文献   

7.
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   

8.
Mg(AlH4)2 and CaAlH5 were synthesized by direct ball milling of AlH3 and MgH2 or AlH3 and CaH2 hydrides. The XRD profiles indicated crystalline compounds. Several ball-milling conditions were studied and the optimum parameters were found. Among these, the key parameter is the pause used to cool down the milling device, which allows reducing the temperature rise during milling. Thus, the maximum yield of complex hydrides was obtained by minimizing the desorbed alane amount. The decomposition properties were studied and were in agreement with those reported for different synthesis methods. Mg(AlH4)2 with a good hydrogen capacity and a decomposition reaction enthalpy close to 0 kJ/mol H2 can be a candidate for one-way storage systems. As for CaAlH5, it might be suitable for reversible hydrogen storage thanks to its dehydrogenation reaction enthalpy (26 kJ/mol H2). However, rather high activation energy values were evaluated for both compounds (119 and 161 kJ/mol, respectively).  相似文献   

9.
LiAlH4 containing 5 wt.% of nanometric Fe (n-Fe) shows a profound mechanical dehydrogenation by continuously desorbing hydrogen (H2) during high energy ball milling reaching ∼3.5 wt.% H2 after 5 h of milling. In contrast, no H2 desorption is observed during low energy milling of LiAlH4 containing n-Fe. Similarly, no H2 desorption occurs during high energy ball milling for LiAlH4 containing micrometric Fe (μ-Fe) and, for comparison, both the micrometric and nanometric Ni (μ-Ni and n-Ni) additive. X-ray diffraction studies show that ball milling results in a varying degree of the lattice expansion of LiAlH4 for both the Fe and Ni additives. A volumetric lattice expansion larger than 1% results in the profound destabilization of LiAlH4 accompanied by continuous H2 desorption during milling according to reaction: LiAlH4 (solid) → 1/3Li3AlH6 + 2/3Al + H2. It is hypothesized that the Fe ions are able to dissolve in the lattice of LiAlH4 by the action of mechanical energy, replacing the Al ions and forming a substitutional solid solution. The quantity of dissolved metal ions depends primarily on the total energy of milling per unit mass of powder generated within a prescribed milling time, the type of additive ion e.g. Fe vs. Ni and on the particle size (micrometric vs. nanometric) of metal additive. For thermal dehydrogenation the average apparent activation energy of Stage I (LiAlH4 (solid) → 1/3Li3AlH6 + 2/3Al + H2) is reduced from the range 76 to 96 kJ/mol for the μ-Fe additive to about 60 kJ/mol for the n-Fe additive. For Stage II dehydrogenation (1/3Li3AlH6 → LiH+1/3Al + 0.5H2) the average apparent activation energy is within the range 77–93 kJ/mol, regardless of the particle size of the Fe additive (μ-Fe vs. n-Fe). The n-Fe and n-Ni additives, the latter used for comparison, provide nearly identical enhancement of dehydrogenation rate during isothermal dehydrogenation at 100 °C. Ball milled (LiAlH4 + 5 wt.% n-Fe) slowly self-discharges up to ∼5 wt.% H2 during storage at room temperature (RT), 40 and 80 °C. Fully dehydrogenated (LiAlH4 + 5 wt.% n-Fe) has been partially rehydrogenated up to 0.5 wt.% H2 under 100 bar/160°C/24 h. However, the rehydrogenation parameters are not optimized yet.  相似文献   

10.
The effects of K2TiF6 on the dehydrogenation properties of LiAlH4 were investigated by solid-state ball milling. The onset decomposition temperature of 0.8 mol% K2TiF6 doped LiAlH4 is as low as 65 °C that 85 °C lower than that of pristine LiAlH4. Isothermal dehydrogenation properties of the doped LiAlH4 were studied by PCT (pressure–composition–temperature). The results show that, for the 0.8 mol% K2TiF6 doped LiAlH4 that dehydrogenated at 90 °C, 4.4 wt% and 6.0 wt% of hydrogen can be released in 60 min and 300 min, respectively. When temperature was increased to 120 °C, the doped LiAlH4 can finish its first two dehydrogenation steps in 170 min. DSC results show that the apparent activation energy (Ea) for the first two dehydrogenation steps of LiAlH4 are both reduced, and XRD results suggest that TiH2, Al3Ti, LiF and KH are in situ formed, which are responsible for the improved dehydrogenation properties of LiAlH4.  相似文献   

11.
A mixture of [3LiBH4 + MnCl2] was processed by high energy ball milling in ultra-high purity hydrogen gas for 0.5 and 1 h. The XRD patterns of milled powders show the sole diffraction peaks of LiCl. The reaction occurring during milling of [3LiBH4 + MnCl2] seems to have all characteristics of the metathesis-type reactions occurring between borohydrides (LiBH4 and NaBH4) and metal chlorides (MCln) induced in a solid state by a mechano-chemical activation synthesis (MCAS). Under pressure of 0.1 MPa H2 (atmospheric) the ball milled [3LiBH4 + MnCl2] mixture is able to desorb ∼4.0 wt.% H2 at 100 °C within 21,000 s and ∼4.5 wt.% H2 at 120 and 200 °C within 8000 s and 4000 s, respectively. The addition of n-Ni with SSA = 60.5 m2/g allows desorption of ∼3.7wt.%H2 within 8,700 s at 100 °C. This is one of the highest H2 desorption capacities obtained for a complex hydride at 100 °C under atmospheric pressure of H2 taking into account the fact that the microstructure contains some amount of a useless LiCl constituent. The activation energy of hydrogen desorption for a ball milled undoped [3LiBH4 + MnCl2] is ∼102 kJ/mol and ∼98 and 92 kJ/mol after doping with 5 wt.% of nanometric Ni having specific surface area (SSA) of 9.5 and 60.5 m2/g, respectively. After volumetric desorption from 100 to 450 °C the XRD patterns show only LiCl. The n-Ni additive slightly lowers the total quantity of desorbed H2. Re-absorption tests, under pressure of 10 MPa H2 at 200 °C, show that the system is, most likely, irreversible. Flammability studies show that the ball milled [3LiBH4 + MnCl2] mixture can be ignited by scraping the cylinder walls with a metal tool as well when it is thrown and dispersed in air in a powder form. It also reacts violently in contact with water and a nitric acid.  相似文献   

12.
A significant decrease in the dehydrogenation temperature of Mg(AlH4)2 was achieved by low-energy ball milling with TiF4. Approximately 8.0 wt% of hydrogen was released from the Mg(AlH4)2-0.025TiF4 sample with an on-set temperature of 40 °C, which represents a decrease of 75 °C relative to pristine Mg(AlH4)2. In contrast to the three-step reaction for pristine Mg(AlH4)2, hydrogen desorption from the TiF4-doped sample involves a two-step process because the Ti-based species participates in the dehydrogenation reaction. The presence of TiF4 alters the nucleation and growth of the dehydrogenation product, significantly decreasing the activation energy barrier of the first step in the dehydrogenation of Mg(AlH4)2. Further hydrogenation measurements revealed that the presence of the Ti-based species was also advantageous for hydrogen uptake, as the on-set hydrogenation temperature was only 100 °C for the dehydrogenated TiF4-doped sample, compared with 130 °C for the additive-free sample.  相似文献   

13.
Lithium aluminum hydride (LiAlH4) is considered as an attractive candidate for hydrogen storage owing to its favorable thermodynamics and high hydrogen storage capacity. However, its reaction kinetics and thermodynamics have to be improved for the practical application. In our present work, we have systematically investigated the effect of NiCo2O4 (NCO) additive on the dehydrogenation properties and microstructure refinement in LiAlH4. The dehydrogenation kinetics of LiAlH4 can be significantly increased with the increase of NiCo2O4 content and dehydrogenation temperature. The 2 mol% NiCo2O4-doped LiAlH4 (2% NCO–LiAlH4) exhibits the superior dehydrogenation performances, which releases 4.95 wt% H2 at 130 °C and 6.47 wt% H2 at 150 °C within 150 min. In contrast, the undoped LiAlH4 sample just releases <1 wt% H2 after 150 min. About 3.7 wt.% of hydrogen can be released from 2% NCO–LiAlH4 at 90 °C, where total 7.10 wt% of hydrogen is released at 150 °C. Moreover, 2% NCO–LiAlH4 displayed remarkably reduced activation energy for the dehydrogenation of LiAlH4.  相似文献   

14.
The main objective of this work was to investigate the different effects of transition metals (TiO2, VCl3, HfCl4) on the hydrogen desorption/absorption of NaAlH4. The HfCl4 doped NaAlH4 showed the lowest temperature of the first desorption at 85 °C, while the one doped with VCl3 or TiO2 desorbed at 135 °C and 155 °C, respectively. Interestingly, the temperature of desorption in subsequent cycles of the NaAlH4 doped with TiO2 reduced to 140 °C. On the contrary, in the case of NaAlH4 doped with HfCl4 or VCl3, the temperature of desorption increased to 150 °C and 175 °C, respectively. This may be because Ti can disperse in NaAlH4 better than Hf and V; therefore, this affected segregation of the sample after the desorption. The maximum hydrogen absorption capacity can be restored up to 3.5 wt% by doping with TiO2, while the amount of restored hydrogen was lower for HfCl4 and VCl3 doped samples. XRD analysis demonstrated that no Ti-compound was observed for the TiO2 doped samples. In contrast, there was evidence of Al–V alloy in the VCl3 doped sample and Al–Hf alloy in the HfCl4 doped sample after subsequent desorption/absorption. As a result, the V- or Hf-doped NaAlH4 showed the lower ability to reabsorb hydrogen and required higher temperature in the subsequent desorptions.  相似文献   

15.
The thermal transformations in the lithium alanate-amide system consisting of lithium aluminum hydride (LiAlH4) and lithium amide (LiNH2), mixed in a 1:1 M ratio, were investigated using the pressure-composition-temperature analysis, solid-state nuclear magnetic resonance, X-ray powder diffraction, and residual gas analysis. Below 250 °C, the alanate decomposes into Al, LiH and H2, through the formation of Li3AlH6, whereas the amide remains largely intact. The release of gaseous hydrogen corresponds to approximately 5 wt%. Above 250 °C, additional ∼4 wt% of hydrogen is produced through solid-state reactions among LiNH2, LiH and metallic Al, through the formation of intermetallic Li-Al binary alloy and an unidentified intermediate. The overall reaction of the thermochemical transformation of the LiAlH4-LiNH2 mixture results in the production of Li3AlN2, metallic Al, LiH and the release of 9 wt% of gaseous hydrogen. The reaction mechanism of the thermal decomposition is different from one identified earlier during mechanical treatment of the same system. Rehydrogenation of the thermally-decomposed products of LiAlH4-LiNH2 mixture using high hydrogen pressure (180 bar) and heating (275 °C) yields LiNH2 and amorphous aluminum nitride (AlN).  相似文献   

16.
Dehydrogenation behavior of LiAlH4 (lithium alanate) admixed with multi-walled carbon nanotubes (MWCNTs) was investigated by using high-pressure thermal gravimetric analysis (HPTGA) and in-situ synchrotron X-ray diffraction (XRD) technique. HPTGA results indicated that MWCNTs play a catalytic role in the dehydrogenation of LiAlH4, subsequently decreasing the dehydrogenation temperature and improving the desorption kinetics. With the proper amount of MWCNT, the initial dehydrogenation temperature of LiAlH4 decreased by approximately 60 °C. Furthermore, in-situ synchrotron XRD analysis confirmed the dehydrogenation reaction products and paths of LiAlH4, with and without MWCNT addition.  相似文献   

17.
18.
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   

19.
The structure stability of nanometric-Ni (n-Ni) produced by Vale Inco Ltd. Canada as a catalytic additive for MgH2 has been investigated. Each n-Ni filament is composed of nearly spherical interconnected particles having a mean diameter of 42 ± 16 nm. After ball milling of the MgH2 + 5 wt.%n-Ni mixture for 15 min the n-Ni particles are found to be uniformly embedded within the particles of MgH2 and at their surfaces. Neither during ball milling of the MgH2 + 5 wt.%n-Ni mixture nor its first decomposition at temperatures of 300, 325, 350 and 375 °C the elemental n-Ni reacts with the elemental Mg to form the Mg2Ni intermetallic phase (and eventually the Mg2NiH4 hydride). The n-Ni additive acts as a strong catalyst accelerating the kinetics of desorption. From the Arrhenius and Johnson–Mehl–Avrami–Kolmogorov theory the activation energy for the first desorption is determined to be ∼94 kJ/mol. After cycling at 300 °C the activation energy for desorption is determined to be ∼99 kJ/mol. This is much lower than ∼160 kJ/mol observed for the undoped and ball milled MgH2. During cycling at 275 and 300 °C the n-Ni additive is converted into Mg2Ni (Mg2NiH4). The newly formed Mg2NiH4 has a nanosized grain on the order of 20 nm. Its catalytic potency seems to be similar to its n-Ni precursor. The formation of Mg2Ni (Mg2NiH4) may be one of the factors responsible for the systematic decrease of hydrogen capacity observed upon cycling at 275 and 300 °C.  相似文献   

20.
The catalytic effects of rare earth fluoride REF3 (RE = Y, La, Ce) additives on the dehydrogenation properties of LiAlH4 were carefully investigated in the present work. The results showed that the dehydrogenation behaviors of LiAlH4 were significantly altered by the addition of 5 mol% REF3 through ball milling. The destabilization ability of these catalysts on LiAlH4 has the order: CeF3>LaF3>YF3. For instance, the temperature programmed desorption (TPD) analyses showed that the onset dehydrogenation temperature of CeF3 doped LiAlH4 was sharply reduced by 90 °C compared to that of pristine LiAlH4. Based on differential scanning calorimetry (DSC) analyses, the dehydriding activation energies of the CeF3 doped LiAlH4 sample were 40.9 kJ/mol H2 and 77.2 kJ/mol H2 for the first and second dehydrogenation stages, respectively, which decreased about 40.0 kJ/mol H2 and 60.3 kJ/mol H2 compared with those of pure LiAlH4. In addition, the sample doped with CeF3 showed the fastest dehydrogenation rate among the REF3 doped LiAlH4 samples at both 125 °C and 150 °C during the isothermal desorption. The phase changes in REF3 doped LiAlH4 samples during ball milling and dehydrogenation were examined using X-ray diffraction and the mechanisms related to the catalytic effects of REF3 were proposed.  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