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.
The structural relationship between the hydride phases in Ti–Mo–H solid solution system (Mo content up to 15 at% in the alloy) during dehydrogenation process under annealing has been studied by conventional and in situ X-ray powder diffraction and transmission electron microscopy (TEM) analysis. During dehydrogenation, the saturated hydrides of the Ti–Mo alloys with fcc δ-phase structure transfer into bcc β-phase at higher temperatures. An associated hydrogen concentration reduction for the δ-phase hydride is observed in the process. However, as the hydrogen concentrations decrease to certain values (H/M  1.1–1.7), the unsaturated δ-phase formed at high temperature would become unstable at lower temperature, and transfer into a tetragonal phase (denoted the -phase here). Unlike that of the -phase in Ti–H system, the phase transition does not occur for the saturated δ-phase with hydrogen concentration close to the stoichiometric limit. The hydrogen concentration of this -phase hydride is in between that of the tetragonal γ and -phase in Ti–H system, but more close to the γ-phase. The occurrence region of this -phase expands along with the increase of the Mo content in the alloys. The phase has a lattice similar to that of the -phase in Ti–H system with corresponding fct unit-cell c/a < 1.  相似文献   

2.
Lattice parameters, hydrogen absorption properties and electrochemical cycling properties up to 240 cycles have been measured as a function of the Ce content for alloys of composition La0.82−xCexNd0.15Pr0.03Ni3.55Mn0.4Al0.3Co0.75 (0≤x≤0.82). The results show the strong increase of the plateau pressure correlated to the cell volume decrease as a function of x. On the other hand, the hydrogen capacity measured in solid–gas reaction as well as the electrochemical capacity decreases slightly. The results show that both La and Ce have to be present to achieve a good cycle life, the cycling degradation being almost independent of their relative quantities in a broad range of concentrations.  相似文献   

3.
Zr7Ni10 has three hydrogen occlusion phases, , β and γ, and the following unusual features are known for the phase transitions in the Zr7Ni10–H2 system: (1) The intermediate hydride phase (β) appears only during dehydrogenation but not during hydrogenation, and (2) The continuous hydrogen solid solution phase () exhibits a much higher hydrogen solubility during hydrogenation than during dehydrogenation. In order to clarify the mechanism about the difference in the hydrogen solubility of the phase, the relation between the pressure-composition isotherms and corresponding structural change has been examined by a conventional volumetric method and X-ray diffraction. Through the examination, we discovered that the crystal structure of the phase, which undergoes hydrogenation followed by dehydrogenation, is different from that of its pure metal phase, where the crystal structure of the dehydrogenated phase changes from an orthorhombic structure to a tetragonal structure. The conditions causing the structural change were then examined, and it has been found that the phase maintains its original orthorhombic structure as long as it is hydrogenated so as not to absorb enough hydrogen to change it to the hydride with a higher hydrogen content (γ). The phenomenon can be understood as one of the hydrogen-assisted phase transitions such as hydrogen-induced amorphization (HIA) in the sense that the phase transition requires hydrogenation under special conditions.  相似文献   

4.
The kinetics of hydrogen absorption and desorption by magnesium has been investigated by a volumetric technique. Experimental data have been analysed in order to find the rate determining step for both the absorption and desorption processes. It is shown that a nucleation and growth (NG) mechanism, with exponent values n=2 for desorption and n=0.5 to n=l for absorption provides suitable equations in order to fit the experimental data. The influence of hydrogen pressure and temperature on the process rate has been studied to obtain an expression for the driving force of the reaction and its activation energy. The driving force for desorption seems to follow a parabolic law though the experimental data are also compatible with a linear law. According to our data, the rate determining step in the desorption of hydrogen by magnesium, and probably also in the absorption process, seems to be the hydrogen diffusion through the β phase. An activation energy for such a diffusion process of 100±10 kJ mol−1H has been obtained from the desorption data.  相似文献   

5.
The single phase nature of the alloys LaNi4.9In0.1, LaNi4.8In0.2, NdNi4.9In0.1, NdNi4.8In0.2 of the systems LaNi5−xInx and NdNi5−xInx was confirmed by means of X-ray powder diffractometry. Nonstoichiometric alloys LaNi4.8 and NdNi4.8 were prepared and were also found to be good single phase materials. All these alloys crystallize with the same hexagonal structure of the CaCu5 type (space group P6/mmm) as do their prototypes LaNi5 and NdNi5. In order to determine the interaction with hydrogen the alloys were exposed to hydrogen gas and the pressure composition desorption isotherms were measured. It was found that all alloys react readily and reversibly absorb large amounts of up to 6.54 hydrogen atoms per alloy formula unit. Generally the equilibrium pressure and the hydrogen capacity decrease with the decreasing nickel content. Presence of indium in the alloy acts in favour of these trends. Furthermore, the increasing content of indium in the alloy system drastically alters the slope and the pressure of the plateau observed at higher pressure of the two isotherm plateaux of the NdNi5–hydrogen system. The final result is a merge of both plateaux into a single one for the hydrogen desorption isotherms of NdNi4.8In0.2. However, the isotherms of nonstoichiometric NdNi4.8 still exhibit two separated pressure plateau regions. The thermodynamic parameters of hydride formation, i.e., the entropy change, the enthalpy and the Gibbs free energy of formation have also been extracted for all alloy–hydrogen systems.  相似文献   

6.
Complex borohydrides of light metals are promising hydrogen storage materials due to their high hydrogen capacity. However, they exhibit two main drawbacks: their high thermodynamic stability and their slow kinetics. In the present work, the effect of various reactants on the formation kinetics of complex borohydrides is investigated. It is found that the kinetic barriers for the formation of LiBH4, NaBH4 and Ca(BH4)2 are drastically reduced when MgB2 is used instead of B as starting material. Since this kinetic enhancement is observed in all borohydride studied so far, the observed effect is attributed to the higher reactivity of B in MgB2 to form [BH4] complexes. In addition, by using MgB2 instead of elemental B, the corresponding reaction enthalpies are reduced by about 10 kJ/mol H, while the high gravimetric hydrogen capacities are largely preserved, i.e. LiBH4 + MgH2 with 11.4 wt%, Ca(BH4)2 + MgH2 with 8.3 wt%, and NaBH4 + MgH2 with 7.8 wt%.  相似文献   

7.
Ti–V–Mn based hydrides are one family of alloys with improved hydrogenation properties and they have a great potential to replace the AB5 alloys as the sorption materials in hydrogen compression systems, although there still are many problems associated with their use, including unstable reversible hydrogen capacity and unfavorable thermodynamic properties. To gain a better understanding on the effect of the substitution elements and to optimize the alloy composition for high storage capacity, the influence of the alloy stoichiometry was investigated. Ti–Zr–V–Mn alloys were prepared by arc melting technique and were annealed in vacuum at temperature above 900 °C to obtain great sorption properties. Hydrogen absorption and desorption kinetics and PCT characteristics of these alloys at ambient temperature were measured and compared. These hydrogen storage features were also discussed in relation to the effect of alloy element compositions. Ti–Zr–V–Mn alloy cycling behavior was also examined.  相似文献   

8.
Hydrogen energy may provide the means to an environmentally friendly future. One of the problems related to its application for transportation is “on board” storage. Hydrogen storage in solids has long been recognized as one of the most practical approaches for this. Recently the hydrogen storage system, (Li3N + 2H2  LiNH2 + 2LiH), was introduced by Chen et al. [P. Chen, Z. Xiong, J. Luo, J. Lin, K.L. Tan, Nature 420 (2002) 302–304. [1]]. This type of material has attracted a great attention of the researchers from the metal hydride research community due to its high reversible storage capacity, up to 11.5 wt%. Currently the Li–Mg–N–H system has been shown to be able to deliver 5.2 wt% reversibly at a H2 pressure of 30 bar and temperature of 200 °C. The accessibility of the capacity beyond 5.2 wt% is being actively explored. One of the issues related to the application of the metal–N–H storage systems is NH3 formation that takes place simultaneously with H2 release. NH3 formation will not only damage the catalyst in a fuel cell, but also accelerate the cyclic instability of the H-storage material since the metal–N–H system turns into a metal–H system after loosing nitrogen and, therefore, it would not function at the temperature and pressure range designed for the metal–N–H system. The accurate determination of the amounts of NH3 in the H2 is, therefore, very important and has not been previously reported. Here a novel method to quantify NH3 in the desorbed H2, the Draeger Tube, is reported as being suitable for this purpose. The results indicate that the concentration of NH3 in desorbed H2 increases with the desorption temperature. For the (2LiNH2 + MgH2) system the NH3 concentration was found to be 180 ppm at 180 °C and 720 ppm at 240 °C.  相似文献   

9.
The in-situ XRD measurements on dehydrogenation/rehydrogenation of the Li–Mg–N–H system were performed in this work. The ballmilled mixture of 8LiH and 3Mg(NH2)2 as a hydrogenated phase gradually changed into Li2NH as a dehydrogenated phase during heat-treatment at 200 °C in vacuum for 50 h. Neither Mg-related phases nor other intermediate phases were recognized in the dehydrogenated phase. With respect to the hydrogenation process, the dehydrogenated state gradually returned to the mixed phase of the LiH and Mg(NH2)2 without appearance of any intermediate phases during heat treatment at 200 °C under 5 MPa H2 for 37 h and during slow cooling down to room temperature through 24 h. In the hydrogenation process at 200 °C under 1 MPa H2, however, the growing up of the LiNH2 and LiH phase was observed in the XRD profiles before the 3Mg(NH2)2 and 8LiH phases were formed as the final hydrogenated state. This indicates that the LiNH2 and LiH phase essentially appears as an intermediate state in the Li–Mg–N–H system composed of 3Mg(NH2)2 and 8LiH.  相似文献   

10.
Isothermal hydrogen absorption properties of the ball milled mixture of 3Mg(NH2)2 and 8LiH after dehydrogenation at 200 °C under high vacuum were investigated at two different temperatures of 150 and 200 °C. The pressure–composition isotherm (PCT) curve at 200 °C revealed a two-plateaus-like behavior, while the PCT curve at 150 °C showed a single-plateau-like behavior. The hydrogenated phases were composed of LiH and Mg(NH2)2 under 9 MPa at 200 °C, while those were observed as mixed phases of LiH and LiNH2 at 150 °C without any trace of Mg(NH2)2 in XRD measurements. These results indicate that there are two-step hydrogenation processes corresponding to high and low pressures at 200 °C, but the kinetics at 150 °C is too slow to proceed with the second hydrogenating step at high pressure region.  相似文献   

11.
Mg6Ir2H11 has been synthesised by both hydrogenation of the intermetallic compound Mg3Ir at 20 bar and 300 °C, and sintering of the elements at 500 °C under 50 bar hydrogen pressure. Neutron powder diffraction on the deuteride indicates a monoclinic structure (space group P21/c, Mg6Ir2D11: a=10.226(1), b=19.234(2), c=8.3345(9) Å, β=91.00(1)°, T=20 °C) that is closely related to orthorhombic Mg6Co2H11. It contains a square-pyramidal [IrH5]4− complex and three saddle-like [IrH4]5− complexes of which one is ordered and two are disordered. Five hydride anions H are exclusively bonded to magnesium. The compound has a red colour, is presumably non-metallic and decomposes under 3 bar argon at 500 °C into Mg3Ir, iridium and a previously unreported intermetallic compound of composition Mg5Ir2.  相似文献   

12.
An effective and durable hydrogen storage module was required to fuel micro-power systems. Two primary specifications for the hydrogen fuel module in this application were a high volumic storage capacity and rapid hydrogen storage and release under atmospheric pressure or lower at room temperature. In addition, the hydrogen module should be operable for thousands of cycles with fast hydriding and dehydriding rates and be resistant to deactivation on exposure to air for many months and longer. In our prior work, mechanical grinding a small amount of palladium with the hydrogen storage alloys was shown to greatly improve the hydrogen storage performance. The palladium treatment of three intermetallic alloys, AB5 type LaNi4.7Al0.3 and CaNi5, and A2B type Mg2Ni, lowered the activation pressure to sub-atmospheric pressure at room temperature and also significantly increased the hydrogen absorption and desorption rates. This work focused on the durability of hydrogen absorption and desorption performances after exposure of the storage materials to air. The palladium treated hydrogen storage alloys retained both low activation pressures and fast absorption and desorption rates even after more than 2 years air exposure.  相似文献   

13.
The hydrogen absorption–desorption performance of the body-centered-cubic (bcc) Ti–V–Cr–Pd alloys have been investigated. Ti33V33Cr34 ingots with 0, 0.05, 0.5 at.% Pd were prepared by arc melting. X-ray diffraction (XRD) revealed that all of these alloys were homogeneous bcc solid solutions. Pd-containing (0.05, 0.5 at.% Pd) Ti–V–Cr alloys have better initial activation properties than those without Pd, and the desorption plateau pressure of the (Ti33V33Cr34)99.5Pd0.5 alloy was substantially higher than that of the alloy without Pd. It is also found that the hysteresis difference is smaller in these alloys and degradation of hydrogen absorption capacity becomes steady after the 25th cycling test. (Ti33V33Cr34)99.5Pd0.5 alloy exhibits large hydrogen absorption and desorption capacity of up to 3.42 and 2.07 mass% at 353 K, respectively.  相似文献   

14.
A kinetic and electron microscopy study of first hydriding of magnesium preliminary mechanically alloyed with the addition of NaF or NaCl salt has been performed. The salts have been found to modify the magnesium particle surfaces in a different way in the course of both mechanical alloying and hydriding. The action of NaCl consists in the local destruction of the oxide layer on the magnesium surface, facilitating the hydride nucleation process, whereas formation of NaMgF3 has been observed at the very initial stages of hydriding of the Mg–NaF mechanical alloy. This ternary fluoride has been shown to play an active role in the process of first hydriding affecting the reaction kinetics and altering the overall course of the reaction.  相似文献   

15.
We tried to improve the hydrogen sorption properties of Mg by mechanical grinding under H2 (reactive mechanical grinding) with oxides Cr2O3, Al2O3 and CeO2. The hydriding rates of Mg are reportedly controlled by the diffusion of hydrogen through a growing Mg hydride layer. The added oxides can help pulverization of Mg during mechanical grinding. A part of Mg is transformed into MgH2 during reactive mechanical grinding. The Mg+10wt.%Cr2O3 powder has the largest transformed fraction 0.215, followed in order by Mg+10wt.%CeO2 and Mg+10wt.%Al2O3. The Mg+10wt.%Cr2O3 powder has the largest hydriding rates at the first and fifth hydriding cycle, followed in order by Mg+10wt.%Al2O3 and Mg+10wt.%CeO2. Mg+10wt.%Cr2O3 absorbs 5.87wt.% H at 573 K, 11 bar H2 during 60 min at the first cycle. The Mg+10wt.%Cr2O3 powder has the largest dehydriding rates at the first and fifth dehydriding cycle, followed by Mg+10wt.%CeO2 and Mg+10wt.%Al2O3. It desorbs 4.44 wt.% H at 573 K, 0.5 bar H2 during 60 min at the first cycle. All the samples absorb and desorb less hydrogen at the fifth cycle than at the first cycle. It is considered that this results from the agglomeration of the particles during hydriding–dehydriding cycling. The average particle sizes of the as-milled and cycled powders increase in the order of Mg+10wt.%Cr2O3, Mg+10wt.%Al2O3 and Mg+10wt.%CeO2. The quantities of hydrogen absorbed or desorbed for 1 h for the first and fifth cycles decrease in the order of Mg+10wt.%Cr2O3, Mg+10wt.%Al2O3 and Mg+10wt.%CeO2. The quantities of absorbed or desorbed hydrogen increase as the average particle sizes decrease. As the particle size decreases, the diffusion distance shortens. This leads to the larger hydriding and dehydriding rates. The Cr2O3 in the Mg+10wt.%Cr2O3 powder is reduced after hydriding–dehydriding cycling. The much larger chemical affinity of Mg than Cr for oxygen leads to a reduction of Cr2O3 after cycling.  相似文献   

16.
Hydrogen strorage alloys with formula La1.5Mg0.5Ni7 were prepared by induction melting followed by different annealing treatments (1073, 1123 and 1173 K) for 24 h. The alloy composition, alloy microstructure and electrochemical properties were investigated, respectively. The results showed that the multi-phase structure of as-cast alloy was converted into a double-phase structure (Gd2Co7-type phase and Ce2Ni7-type phase) through annealing treatments. Mg atoms were mainly located in Laves unit of Gd2Co7-type unit cell and Ce2Ni7-type unit cell. The electrochemical capacity of alloy electrodes after annealing treatment could be up to 390 mAh/g. The cyclic stability of alloy electrodes was significantly improved by annealing treatments; After 150 charge/discharge cycles, the capacity retention ratio of alloy annealed at 1173 K was the highest (81.9%). The high rate dischargeability of alloy electrodes was also improved due to annealing treatment.  相似文献   

17.
The structural and electrochemical properties of the as-cast and rapidly solidified Ti0.8Zr0.2V2.4Mn0.48Cr0.72Ni0.9 alloys were studied. Both the as-cast and the rapidly solidified alloys were mainly composed of a C14 Laves phase matrix with hexagonal structure and a V-based solid solution phase with body centered cubic (BCC) structure. The V-based solid solution phase showed very fine dendrites in the rapidly solidified alloy instead of the large dendrites as observed in the as-cast alloy. In addition, the content of the C14 Laves phase in the alloy decreased greatly after rapid solidification. Electrochemical measurements indicated that the rapidly solidified alloy had a lower discharge capacity, a slower activation rate, a worse high rate dischargeability, a smaller exchange current density and limiting current density, but an improved cycle life compared with that of the as-cast alloy.  相似文献   

18.
The phase relations and hydrogenation behavior of Sr(Al1−xMgx)2 alloys were studied. The pseudobinary C36-type Laves phase Sr(Al,Mg)2 was found as a structural intermediate between the Zintl phase and the C14 Laves phase. The single-phase regions for the Zintl phase, C36 phase and C14 phase, were determined to be x=0–0.10, 0.45–0.68 and 0.80–1, respectively. The Mg-substituted Zintl phase Sr(Al0.95Mg0.05)2 can be hydrogenated to Sr(Al,Mg)2H2 at about 473 K. However, the Sr(Al,Mg)2H2 directly decomposes into SrH2 and Sr(Al,Mg)4 starting at 513 K. When the temperature is 573 K, the C36 Laves phase Sr(Al0.5Mg0.5)2 can be hydrogenated into SrMgH4 and Al, while the C14 Laves phase Sr(Al0.1Mg0.9)2 is hydrogenated into SrMgH4, Mg17Al12 and Mg.  相似文献   

19.
The hydrogenation disproportionation desorption recombination (HDDR) process was performed on the generally used alloy composition of Nd12.6Fe63.1Co17.4Zr0.1Ga0.3B6.5 and a low rare earth content alloy composition of Nd11.0Fe65.0Co17.8Zr0.1Ga0.3B5.8. A detailed evaluation was made of the relationship between the microstructure and magnetic properties of these HDDR-processed magnetic powders with respect to their different rare earth element concentrations. The HDDR-processed powders of both alloy compositions were transformed to the Nd2Fe14B phase consisting of fine recombined crystal grains of around 400–500 nm in size and maintained the anisotropic magnetic characteristic that was present before HDDR processing. However, reduction of the rare earth content drastically reduced coercivity, and the alloy composition of Nd11.0Fe65.0Co17.8Zr0.1Ga0.3B5.8 did not manifest magnetic properties. From the results of an examination of their microstructures, it was inferred that the coercivity decreased due to a decline in the concentration of the rare earth element at the grain boundaries of the fine Nd2Fe14B grains with the reduction of the rare earth content of the alloys. Accordingly, in magnetic powders obtained by the HDDR process, the nucleation type of coercivity mechanism predominates, in which rare earth-rich regions present at the grain boundaries of fine Nd2Fe14B grains play a large role in the manifestation of coercivity.  相似文献   

20.
The hydrogen storage properties of thin films of FeTi evaporated on Si substrates and covered with 20 nm Pd were studied. The films serve as a model system for powdered FeTi, with grains that are (partly) covered with Pd. This material could serve as a practical hydrogen storage material. The 20 nm Pd layer prevents the oxidation of the FeTi layer during air exposure up to temperatures of 200 °C and during H charging and discharging in impure hydrogen. The FeTi is a mixture of amorphous and nano-crystalline material. Two FeTi compositions (43 at.% Fe, i.e. Ti-rich, and 56 at.% Fe, i.e. Fe-rich) were studied. The H charging and discharging characteristics as a function of temperature and pressure are determined from a differential pressure measurement for Fe and Ti-rich material before and after annealing. After discharging in vacuum at a temperature of 150 °C a H residue of H/M  0.12 is observed. The recoverable charging capacity of FeTi (also after many cycles) is 0.9 ΔH/M (H atoms per metal atom) for RT charging at 2700 mbar and vacuum discharging at 150 °C.  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