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 15 毫秒
1.
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   

2.
3.
The hydrogen desorption properties of MgH2–LiAlH4 composites obtained by mechanical milling for different milling times have been investigated by Thermal Desorption Spectroscopy (TDS) and correlated to the sample microstructure and morphology analysed by X-ray Diffraction (XRD) and Scanning Electron Microscopy (SEM). The MgH2–LiAlH4 composites show improved hydrogen desorption properties in comparison with both as-received and ball-milled MgH2. Mixing of MgH2 with small amount of LiAlH4 (5 wt.%) using short mechanical milling (15 min) shifts, in fact, the hydrogen desorption peak to lower temperature than those observed with both as-received and milled MgH2 samples. Longer mixing times of the MgH2–LiAlH4 composites (30 and 60 min) reduce the catalytic activity of the LiAlH4 additive as revealed by the shift of the hydrogen desorption peak to higher temperatures.  相似文献   

4.
The mutual destabilization of LiAlH4 and MgH2 in the reactive hydride composite LiAlH4-MgH2 is attributed to the formation of intermediate compounds, including Li-Mg and Mg-Al alloys, upon dehydrogenation. TiF3 was doped into the composite for promoting this interaction and thus enhancing the hydrogen sorption properties. Experimental analysis on the LiAlH4-MgH2-TiF3 composite was performed via temperature-programmed desorption (TPD), differential scanning calorimetry (DSC), isothermal sorption, pressure-composition isotherms (PCI), and powder X-ray diffraction (XRD). For LiAlH4-MgH2-TiF3 composite (mole ratio 1:1:0.05), the dehydrogenation temperature range starts from about 60 °C, which is 100 °C lower than for LiAlH4-MgH2. At 300 °C, the LiAlH4-MgH2-TiF3 composite can desorb 2.48 wt% hydrogen in 10 min during its second stage dehydrogenation, corresponding to the decomposition of MgH2. In contrast, 20 min was required for the LiAlH4-MgH2 sample to release so much hydrogen capacity under the same conditions. The hydrogen absorption properties of the LiAlH4-MgH2-TiF3 composite were also improved significantly as compared to the LiAlH4-MgH2 composite. A hydrogen absorption capacity of 2.68 wt% under 300 °C and 20 atm H2 pressure was reached after 5 min in the LiAlH4-MgH2-TiF3 composite, which is larger than that of LiAlH4-MgH2 (1.75 wt%). XRD results show that the MgH2 and LiH were reformed after rehydrogenation.  相似文献   

5.
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   

6.
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   

7.
In this work, we have successfully prepared urchin-like NiCo2O4 spheres via a facile hydrothermal method without any templates or surfactants. The as-prepared urchin-like NiCo2O4 spheres have uniform diameters of 6 μm with numerous small nanorods radially-grown from the center. Typical nanorods have diameters of 20–50 nm and lengths of 2–3 μm. Remarkably, urchin-like NiCo2O4 nanostructure as an alternative non-precious metal electrocatalyst for ORR exhibits catalytic performance and excellent cycling stability in alkaline environment.  相似文献   

8.
This study investigated the effect of Nd2O3 and Gd2O3 as catalyst on hydrogen desorption behavior of NaAlH4. Pressure-content-temperature (PCT) equipment measurement proved that both two oxides enhanced the dehydrogenation kinetics distinctly and increasing Nd2O3 and Gd2O3 from 0.5 mol% to 5 mol% caused a similar effect trend that the dehydrogenation amount and average dehydrogenation rate increased firstly and then decreased under the same conditions. 1 mol% Gd2O3–NaAlH4 presented the largest hydrogen desorption amount of 5.94 wt% while 1 mol% Nd2O3–NaAlH4 exerted the fastest dehydrogenation rate. Scanning Electron microscopy (SEM) analysis revealed that Gd2O3–NaAlH4 samples displayed uniform surface morphology that was bulky, uneven and flocculent. The difference of Nd2O3–NaAlH4 was that with the increasing of Nd2O3 content, the particles turned more and more big. Compared to dehydrogenation behavior, this phenomenon demonstrated that small particles structure were beneficial to hydrogen desorption. Besides, the further study found that different catalysts and addition amounts had different effects on the microstructure of NaAlH4.  相似文献   

9.
Spinel NiCo2O4 in different morphologies is of current interest in the design and development of electrochemical supercapacitors. In this work, we synthesized two different morphologies of NiCo2O4 by facile hydrothermal method employing CTAB as a soft template and urea as hydrolysis controlling agent. This study has been undertaken to determine the effect of synthesis temperature on the morphology and pseudocapacitance behavior of the NiCo2O4. We find that the temperature variation in the synthesis procedure has a strong effect on the morphology of NiCo2O4, producing urchin-like morphology at 120 °C (NiCoO-120-cal) and sheaf-like morphology at 200 °C (NiCoO-200-cal) with hierarchical porous textures. The effect of morphology on the electrochemical pseudocapacitance behavior was studied by CV, CP and EIS techniques. Both NiCo2O4 samples show higher electrochemical performance than the parent NiO and Co3O4 synthesized under similar conditions. The maximum specific capacitance values obtained for NiCoO-120-cal and NiCoO-200-cal are 636 and 504 F g−1 respectively, at a current density of 0.5 A g−1. The capacitance retention of NiCoO-120-cal and NiCoO-200-cal samples, respectively, are 76% and 69% after 1000 charge–discharge cycles at a current density of 1 A g−1.  相似文献   

10.
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   

11.
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   

12.
CuCr2O4/TiO2 heterojunction has been successfully synthesized via a facile citric acid (CA)-assisted sol-gel method. Techniques of X-ray diffraction (XRD), transmission electron microscopy (TEM), and UV-vis diffuse reflectance spectrum (UV-vis DRS) have been employed to characterize the as-synthesized nanocomposites. Furthermore, photocatalytic activities of the as-obtained nanocomposites have been evaluated based on the H2 evolution from oxalic acid solution under simulated sunlight irradiation. Factors such as CuCr2O4 to TiO2 molar ratio in the composites, calcination temperature, photocatalyst mass concentration, and initial oxalic acid concentration affecting the photocatalytic hydrogen producing have been studied in detail. The results showed that the nanocomposite of CuCr2O4/TiO2 is more efficient than their single part of CuCr2O4 or TiO2 in producing hydrogen. The optimized composition of the nanocomposites has been found to be CuCr2O4·0.7TiO2. And the optimized calcination temperature and photocatalyst mass concentration are 500 °C and 0.8 g l−1, respectively. The influence of initial oxalic acid concentration is consistent with the Langmuir model.  相似文献   

13.
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   

14.
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   

15.
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   

16.
To improve hydrogen desorption properties of MgH2, mechanical milling of MgH2 with low concentration (2 and 5%) of NaNH2 has been performed. Pre-milling of MgH2 for 10 h has been done and then six samples have been synthesised with different milling times from 15 to 60 min. Microstructural characterisation has been performed using X-ray diffraction (XRD), scanning electron microscopy (SEM) and laser scattering measurements (PSD), and correlated to desorption properties examined using Differential Scanning Calorimetry (DSC) and Hydrogen Sorption Analyser (HSA). Thermal analysis shows that desorption temperatures are shifted towards lower values. It also highlights the significance of milling time and additive concentration on desorption behaviour.  相似文献   

17.
One-dimensional alpha manganese dioxide (α-MnO2) nanorods synthesized by a hydrothermal route were explored as the starting material for preparing lithium manganese spinel LiMn2O4. Pure and highly crystalline spinel LiMn2O4 was easily obtained from α-MnO2 nanorods through a low-temperature solid-state reaction route, while Mn2O3 impurity was present along with the spinel phase when commercial MnO2 was used as starting material. The particle size of LiMn2O4 prepared from α-MnO2 nanorods was about 100 nm with a homogenous distribution. Electrochemical tests demonstrated that the LiMn2O4 thus prepared exhibited a higher capacity than that prepared from commercial MnO2. Therefore, α-MnO2 nanorods are proved to be a promising starting material for the preparation of high quality LiMn2O4.  相似文献   

18.
MgH2 is a perspective hydrogen storage material whose main advantage is a relatively high hydrogen storage capacity (theoretically, 7.6 wt.% H2). This compound, however, shows poor hydrogen desorption kinetics. Much effort was devoted in the past to finding possible ways of enhancing hydrogen desorption rate from MgH2, which would bring this material closer to technical applications. One possible way is catalysis of hydrogen desorption. This paper investigates separate catalytic effects of Ni, Mg2Ni and Mg2NiH4 on the hydrogen desorption characteristics of MgH2. It was observed that the catalytic efficiency of Mg2NiH4 was considerably higher than that of pure Ni and non-hydrated intermetallic Mg2Ni. The Mg2NiH4 phase has two low-temperature modifications below 508 K: un-twinned phase LT1 and micro-twinned phase LT2. LT1 was observed to have significantly higher catalytic efficiency than LT2.  相似文献   

19.
The physical properties and photoelectrochemical characterization of the spinel ZnFe2O4, elaborated by chemical route, have been investigated for the hydrogen production under visible light. The forbidden band is found to be 1.92 eV and the transition is indirectly allowed. The electrical conduction occurs by small polaron hopping with activation energy of 0.20 eV. p-type conductivity is evidenced from positive thermopower and cathodic photocurrent. The flat band potential (0.18 VSCE) determined from the capacitance measurements is suitably positioned with respect to H2O/H2 level (−0.85 VSCE). Hence, ZnFe2O4 is found to be an efficient photocatalyst for hydrogen generation under visible light. The photoactivity increases significantly when the spinel is combined with a wide band gap semiconductor. The best performance with a hydrogen rate evolution of 9.2 cm3 h−1 (mg catalyst)−1 occurs over the new hetero-system ZnFe2O4/SrTiO3 in Na2S2O3 (0.025 M) solution.  相似文献   

20.
In this study, various nanoscale metal oxide catalysts, such as CeO2, TiO2, Fe2O3, Co3O4, and SiO2, were added to the LiBH4/2LiNH2/MgH2 system by using high-energy ball milling. Temperature programmed desorption and MS results showed that the Li–Mg–B–N–H/oxide mixtures were able to dehydrogenate at much lower temperatures. The order of the catalytic effect of the studied oxides was Fe2O3 > Co3O4 > CeO2 > TiO2 > SiO2. The onset dehydrogenation temperature was below 70 °C for the samples doped with Fe2O3 and Co3O4 with 10 wt.%. More than 5.4 wt.% hydrogen was released at 140 °C. X-ray diffraction indicated that the addition of metal oxides inhibited the formation of Mg(NH2)2 during ball milling processes. It is thought that the changing of the ball milling products results from the interaction of oxide ions in metal oxide catalysts with hydrogen atoms in MgH2. The catalytic effect depends on the activation capability of oxygen species in metal oxides on hydrogen atoms in hydrides.  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