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.
Nanocrystalline titanium dioxide loaded carbon spheres (Ti-CSs) with 10wt% TiO2 were synthesized through an easy one-step method using phenolic resols, titania nanoparticles and Pluronic F127 as organic carbon sources, inorganic precursors and surfactant, respectively. The results show that the as-prepared Ti-CSs composite is spherical shape with a diameter ranging from 0.3 to 2 μm, and rutile TiO2 nanoparticles are distributed on the surface of the carbon spheres. Then the kinetics of NaAlH4 was improved through depositing it on the surface of as-prepared Ti-CSs by melt infiltration. The results show that NaAlH4 with Ti-CSs exhibits better hydrogen desorption kinetics than TiF3 or nanocrystalline TiO2 catalysted-NaAlH4, and it starts to release hydrogen at about 40 °C and releases about 25% of the hydrogen content during heating to 60 °C. The results from SEM and XPS show that hydrogen storage properties of NaAlH4 were considerably improved due to the formation of special structure during melt infiltration and the nanocrystalline TiO2 and/or amorphous phase Ti–Al clusters near the subsurface sites, which succeed in combining catalyst addition (TiO2 nanoparticles) and nanoconfinement to improve the kinetics of NaAlH4.  相似文献   

2.
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   

3.
NaAlH4 has been homogeneously mixed with micron- and nano-sized TiO2 powders by high-energy ball milling and their sorption properties have been investigated during hydrogen absorption/desorption cycles. NaAlH4 with TiO2 nanopowder exhibits as good desorption kinetics as NaAlH4 with TiCl3, whereas poor desorption kinetics is observed with micron-sized TiO2 powder. NaAlH4 with TiO2 nanopowder also provides improved cyclic property compared to NaAlH4 with TiCl3 in terms of both desorption rate and hydrogen capacity. X-ray diffraction analysis shows that micron-sized TiO2 remains stable with NaAlH4 after milling, although thermodynamic calculation predicts that TiO2 reacts with NaAlH4.  相似文献   

4.
In a previous paper, it was demonstrated that a MgH2–NaAlH4 composite system had improved dehydrogenation performance compared with as-milled pure NaAlH4 and pure MgH2 alone. The purpose of the present study was to investigate the hydrogen storage properties of the MgH2–NaAlH4 composite in the presence of TiF3. 10 wt.% TiF3 was added to the MgH2–NaAlH4 mixture, and its catalytic effects were investigated. The reaction mechanism and the hydrogen storage properties were studied by X-ray diffraction, thermogravimetric analysis, differential scanning calorimetry (DSC), temperature-programmed-desorption and isothermal sorption measurements. The DSC results show that MgH2–NaAlH4 composite milled with 10 wt.% TiF3 had lower dehydrogenation temperatures, by 100, 73, 30, and 25 °C, respectively, for each step in the four-step dehydrogenation process compared to the neat MgH2–NaAlH4 composite. Kinetic desorption results show that the MgH2–NaAlH4–TiF3 composite released about 2.4 wt.% hydrogen within 10 min at 300 °C, while the neat MgH2–NaAlH4 sample only released less than 1.0 wt.% hydrogen under the same conditions. From the Kissinger plot, the apparent activation energy, EA, for the decomposition of MgH2, NaMgH3, and NaH in the MgH2–NaAlH4–TiF3 composite was reduced to 71, 104, and 124 kJ/mol, respectively, compared with 148, 142, and 138 kJ/mol in the neat MgH2–NaAlH4 composite. The high catalytic activity of TiF3 is associated with in situ formation of a microcrystalline intermetallic Ti–Al phase from TiF3 and NaAlH4 during ball milling or the dehydrogenation process. Once formed, the Ti–Al phase acts as a real catalyst in the MgH2–NaAlH4–TiF3 composite system.  相似文献   

5.
The present investigation describes the hydrogen storage properties of 2:1 molar ratio of MgH2–NaAlH4 composite. De/rehydrogenation study reveals that MgH2–NaAlH4 composite offers beneficial hydrogen storage characteristics as compared to pristine NaAlH4 and MgH2. To investigate the effect of carbon nanostructures (CNS) on the de/rehydrogenation behavior of MgH2–NaAlH4 composite, we have employed 2 wt.% CNS namely, single wall carbon nanotubes (SWCNT) and graphene nano sheets (GNS). It is found that the hydrogen storage behavior of composite gets improved by the addition of 2 wt.% CNS. In particular, catalytic effect of GNS + SWCNT improves the hydrogen storage behavior and cyclability of the composite. De/rehydrogenation experiments performed up to six cycles show loss of 1.50 wt.% and 0.84 wt.% hydrogen capacity in MgH2–NaAlH4 catalyzed with 2 wt.% SWCNT and 2 wt.% GNS respectively. On the other hand, the loss of hydrogen capacity after six rehydrogenation cycles in GNS + SWCNT (1.5 + 0.5) wt.% catalyzed MgH2–NaAlH4 is diminished to 0.45 wt.%.  相似文献   

6.
LiBH4 nano-particles are incorporated into mesoporous TiO2 scaffolds via a chemical impregnation method. And the enhanced desorption properties of the composite have been investigated. The LiBH4/TiO2 sample starts to release hydrogen at 220 °C and the maximal desorption peak occurs at about 330 °C, much lower compared to the bulk LiBH4. Furthermore, the composite exhibits excellent dehydrogenation kinetics, with 11 wt% of hydrogen liberated from LiBH4 at 300 °C within 3 h. X-ray diffraction and Fourier transform infrared spectroscopy are used to confirm the nanostructure of LiBH4 in the TiO2 scaffold. This work demonstrates that confinement within active porous scaffold host is a promising approach for enhancing hydrogen decomposition properties of light-metal complex hydrides.  相似文献   

7.
The co-effects of lanthanide oxide Tm2O3 and porous silica on the hydrogen storage properties of sodium alanate are investigated. NaAlH4-Tm2O3 (10 wt%) and NaAlH4-Tm2O3 (10 wt%)-porous SiO2 (10 wt%) are prepared by the ball milling method, and their hydrogen desorption/re-absorption capacities are compared. Dehydrogenation process was performed at 150 °C under vacuum and rehydrogenation was performed at 150 °C for 4 h under ∼9 MPa in highly pure hydrogen. The results show that Tm2O3 has a catalytic effect on the hydrogen desorption and re-absorption of NaAlH4. The hydrogen desorption capacity of Tm2O3 single-doped NaAlH4 is 4.6 wt%, higher than that of undoped NaAlH4 (4.3 wt%). During the dehydrogenation process, NaAlH4 is completely decomposed and no intermediate product Na3AlH6 is detected. The addition of porous silica improves the dehydrogenation performance of NaAlH4. Tm2O3 and porous silica co-doped NaAlH4 could release a maximum hydrogen amount of 4.7 wt%, higher than that of undoped NaAlH4 and Tm2O3 single-doped NaAlH4. Moreover, porous silica improves the reversibility of hydrogen storage in NaAlH4.  相似文献   

8.
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   

9.
By directly introducing LaCl3, La3Al11, SmCl3, SmAl3 into NaAlH4 system using one-step synthesis method, the effects of these additives on NaAlH4 were systematically investigated with regard to hydriding and dehydriding properties. Results showed that the materials doped with aluminide exhibit similar kinetics to the chloride-doped NaAlH4. The apparent activation energy Ea of doped NaAlH4 were calculated to be 86.4-93.0 kJ/mol and 96.1-99.3 kJ/mol for the first and second dehydrogenation step respectively by using Kissinger’s approach, much lower than those of pristine NaAlH4. A reversible hydrogen capacity of 4.8 wt% can be achieved for the La3Al11- and SmAl3-doped NaAlH4, which is 10-20% higher than chloride-doped NaAlH4. Investigations on the phase evolvement and microstructure in the cycling in LaCl3- and La3Al11-doped NaAlH4 clearly demonstrate that La species is presented as the form of La-Al nanoclusters in the materials. The combination of hydrogen storage properties and the microstructures unequivocally reveal that the in situ formed rare-earth-Al species play a crucial rule in catalyzing the chloride-doped NaAlH4.  相似文献   

10.
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   

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.
TiF3-doped NaH/Al mixture was hydrogenated into Na3AlH6 and NaAlH4 complex hydrides by reactive ball-milling at room temperature through the optimization of milling duration and hydrogen pressure. The analysis of the preparation of NaAlH4 samples during reactive ball-milling process has been performed by XRD and TG/DSC. It has been found that Na3AlH6 was formed under 0.5 MPa hydrogen pressure and 30 h milling duration, while NaAlH4 was formed under 0.8 MPa hydrogen pressure and 45 h milling duration. The process of preparing NaAlH4 by ball-milling was found accomplished via two reaction steps, namely: (1) NaH + Al + H2 → Na3AlH6 and (2) Na3AlH6 + Al + H2 → NaAlH4. As the hydrogen pressure and milling duration increase, the synthetic yield of NaAlH4 and its corresponding dehydriding capacity are both increased. With increased hydrogen pressure (0.8-3 MPa) and milling duration (45-60 h), the cell volume of Na3AlH6 decreases while that of NaAlH4 increases gradually. The abundance of Na3AlH6 phase decreases from 57.76 (x = 0.8, y = 45) to 8.69 wt.% (x = 3, y = 60), and the abundance of NaAlH4 phase increases from 20.63 (x = 0.8, y = 45) to 86.50 wt.% (x = 3, y = 60). All the samples prepared in this way have fairly good activation behavior and fast hydriding/dehydriding reaction kinetics, which are capable of absorbing 4.26 wt.% hydrogen at 120 °C and desorbing 4.12 wt.% hydrogen at 150 °C, respectively. The improvement of hydriding/dehydriding properties is ascribed to the favorable microstructure and ultrafine particle features of nanosized NaAlH4 formed during ball-milling at the optimum synthetic condition.  相似文献   

13.
Sodium aluminum hydride (NaAlH4) was directly synthesized by ball milling NaH/Al co-doped with CeCl3 + KH under a hydrogen pressure of 3 MPa at room temperature. Out of various samples corresponding to xNaH/Al + 0.02CeCl3 + yKH (x + y = 1; y = 0, 0.02, 0.04 mol%) composites, the composite with y = 0.02 exhibits the optimum de/hydrogenation properties. It shows that the addition of KH can effectively improve the dehydrogenation properties of second step reaction of NaAIH4 system. The composite with y = 0.02 starts to release hydrogen from 87 °C and completes dehydrogenation within 20 min at 170 °C, with good cycling de/hydrogenation kinetics at relatively lower temperature (100–140 °C). After ball milling, the CeCl3 precursor can be changed into CeH2 catalytic active component in the first several de/hydrogenation cycles. Apparent activation energy of the second decomposition step of NaAIH4 system can be effectively decreased by addition of KH, resulting in the decrease of desorption temperatures. Based on the microstructure analyses combined with hydrogen storage performances, the improved dehydrogenation properties of sodium aluminum hydride system are ascribed to the lattice volume expansion of Na3AlH6 during the dehydrogenation process resulted from the addition of KH. Moreover, by analyzing the reaction kinetics of CeCl3 + KH co-doped sample, both of the decomposition steps of composite with y = 0.02 were conformed to the two-dimension phase-boundary growth mechanism. The mechanistic investigations gained here could help to understand the de/rehydrogenation behaviors of catalyzed complex metal hydride systems.  相似文献   

14.
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   

15.
LiFePO4/C composite was synthesized at 600 °C in an Ar atmosphere by a soluble starch sol assisted rheological phase method using home-made amorphous nano-FePO4 as the iron source. XRD, SEM and TEM observations show that the LiFePO4/C composite has good crystallinity, ultrafine sphere-like particles of 100-200 nm size and in situ carbon. The synthesized LiFePO4 could inherit the morphology of FePO4 precursor. The electrochemical performance of the LiFePO4 by galvanostatic cycling studies demonstrates excellent high-rate cycle stability. The Li/LiFePO4 cell displays a high initial discharge capacity of more than 157 mAh g−1 at 0.2C and a little discharge capacity decreases from the first to the 80th cycle (>98.3%). Remarkably, even at a high current density of 30C, the cell still presents good cycle retention.  相似文献   

16.
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   

17.
In this study, NaAlH4?based hydrogen storage materials with dopants were prepared by a two-steps in-situ ball milling method. The dopants adopted included Ce, few layer graphene (FLG), Ce + FLG, and CeH2.51. The hydrogen storage materials were studied by non-isothermal and isothermal hydrogen desorption measurements, X-ray diffractions analysis, cycling sorption tests, and morphology analysis. The hydrogen storage performance of the as-prepared NaAlH4 with Ce addition is much better than that with CeH2.51 addition. This is due to that the impact of Ce occurs from the body to the surface of the materials. The addition of FLG further enhances the impact of Ce on the hydrogen storage performance of the materials. The hydrogen storage capacity, hydrogen sorption kinetics, and cycle performance of NaAlH4 with Ce + FLG additions are all better than NaAlH4 materials with the addition of either Ce or FLG alone. The NaAlH4 with Ce and FLG addition starts to release hydrogen at 85 °C and achieves a capacity of 5.06 wt% after heated to 200 °C. The capacity maintains at 4.91 wt% (94.7% of the theoretical value) for up to 8 cycles. At 110 °C, the material can release isothermally a hydrogen capacity of 2.8 wt% within 2 h. The activation energies for the two hydrogen desorption steps of NaAlH4 with Ce and FLG addition are estimated to be 106.99 and 125.91 kJ mol?1 H2, respectively. The related mechanisms were studied with first-principle and experimental methods.  相似文献   

18.
Carbon coated titanium dioxide supported on two-dimensional titanium carbide (C@TiO2/Ti3C2) is synthesized by simple annealing under a flowing acetylene (C2H2) atmosphere, and applied to improve the hydriding/dehydriding behavior of sodium alanate (NaAlH4). The results indicate that as-prepared C@TiO2/Ti3C2 composite exhibits excellent catalytic activity. The initial temperature for hydrogen desorption is reduced by 70 °C compared with the pristine sample. About 4.0 wt% hydrogen is released in 13 min at 140 °C. The apparent activation energies (Ea) of 10 wt% C@TiO2/Ti3C2 catalyzing NaAlH4 for the first two-steps dehydrogenation are 72.41 and 64.27 kJ mol−1 respectively. The structural analyses reveal that C@TiO2/Ti3C2 interacts with NaAlH4 by using ball milling and decomposes to form Ti-species which works in combination with carbon to improve the dehydrogenation performance of NaAlH4. This result provides an important progress in the hydrogen storage of NaAlH4 catalyzed by MXene.  相似文献   

19.
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   

20.
2LiBH4 + MgH2 system is considered as an attractive candidate for reversible hydrogen storage with high capacity and favorable thermodynamics. However, its reaction kinetics has to be further improved for the practical application. In this work, we investigated the effect of NbCl5 additive on the de/hydrogenation kinetics and microstructure refinement in 2LiH–MgB2 composite systematically. The hydrogenation and dehydrogenation kinetics of 2LiH–MgB2 composite can be significantly enhanced with the increase of NbCl5 content. The 3 mol% NbCl5 doped 2LiH–MgB2 composite exhibits the superior reversible hydrogen storage performance, which requires 50 min to uptake 9.0 wt% H2 at 350 °C and release 8.5 wt% H2 at 400 °C, respectively. In contrast, the undoped 2LiH–MgB2 sample uptakes 6.2 wt% H2 and releases 3.1 wt% H2 under identical measurement conditions. Moreover, the 3 mol% NbCl5 doped 2LiH–MgB2 composite can release more than 9.0 wt% H2 within 300 min at 400 °C without obvious degradation of capacity over the first 10 cycles. Microstructure analyses clearly indicate that NbCl5 additive first reacts with LiH to form Nb and LiCl during ball-milling process, and then NbH is formed after the first hydrogenation and stabilized upon further de/hydrogenation cycling. The well-distributed NbH active species play an important role in the improvement of de/hydrogenation kinetics for Li–Mg–B–H system through facilitating hydrogen diffusion rapidly as well as prevent the particles from further growth in the subsequent hydrogenation and dehydrogenation processes.  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