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 reaction pathway and rate-limiting step of dehydrogenation of the LiNH2 + LiH mixture have been investigated. The study reveals that dehydrogenation of the LiNH2 + LiH mixture is diffusion-controlled and the rate-limiting step is NH3 diffusion through the Li2NH product layer outside the LiNH2 shrinking core. This phenomenon is explained based on a model describing the major steps of the dehydriding reaction of the mixture, and related to the evidence obtained from X-ray diffraction and specific surface area measurements of the mixture before and after isothermal hydrogen uptake/release cycles at high homologous temperatures.  相似文献   

2.
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   

3.
The reaction rate of MgH2 with NH3 is studied using a two-layered structure containing a top MgH2 layer and a bottom LiNH2 layer. Quantification of the effluent gas composition from the two-layered structure indicates substantial NH3 emission, while the X-ray diffraction analysis reveals little formation of the reaction products between MgH2 and NH3. In contrast, the study of the two-layered structure containing a top LiH layer and a bottom LiNH2 layer reveals that the reaction between LiH and NH3 is much faster than that between MgH2 and NH3.  相似文献   

4.
Stepwise reactions were observed in the ball milling and heating process of the LiBH4-NaNH2 system by means of X-ray diffraction (XRD) and Fourier transform infrared spectrometry (FT-IR). During the ball milling process, two concurrent reactions take place in the mixture: 3LiBH4 + 4NaNH2 → Li3Na(NH2)4 + 3NaBH4 and LiBH4 + NaNH2 → LiNH2 + NaBH4. The heating process from 50 °C to 400 °C is mainly the concurrent reactions of Li3Na(NH2)4 + 3LiBH4 → 2Li3BN2 + NaBH4 + 8H2 and 2LiNH2 + LiBH4 → Li3BN2H8 → Li3BN2 + 4H2, where the intermediate phases Li3Na(NH2)4 and LiNH2 serve as the reagents to decompose LiBH4. The merged equations for the mechanochemical and the heating reactions below 400 °C can be denoted as 3LiBH4 + 2NaNH2 → Li3BN2 + 2NaBH4 + 4H2. The maximum dehydrogenation capacity in closed system below 400 °C is 5.1 wt.% H2, which agrees well with the theoretical capacity (5.5 wt.% H2) of the merged equation and thus demonstrates the suggested pathway. The subsequent step consists of the decompositions of NaBH4 and Li3Na(NH2)4 within the temperature range of 400 °C-600 °C. The apparent activation energies of the two steps are 114.8 and 123.5 kJ/mol, respectively. They are all lower than that of our previously obtained bulk LiBH4.  相似文献   

5.
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   

6.
The effect of Ti0.4Cr0.15Mn0.15V0.3 (termed BCC due to the body centered cubic structure) alloy on the hydrogen storage properties of MgH2 was investigated. It was found that the hydrogenated BCC alloy showed superior catalysis properties compared to the quenched and ingot samples. As an example, the 1 h milled MgH2 + 20 wt.% hydrogenated BCC shows a peak temperature of dehydrogenation of about 294 °C. This is 16, 27 and 74 °C lower than those of MgH2 ball milled with quenched BCC, ingot BCC and an uncatalysed MgH2 sample, respectively. The hydrogenated BCC alloy is much easier to crush into small particles, and embed in MgH2 aggregates as revealed by X-ray diffraction and scanning electron microscope results. The BCC not only increases the hydrogen atomic diffusivity in the bulk Mg but also promotes the dissociation and recombination of hydrogen. The activation energy, Ea, for the dehydrogenation of the MgH2/hydrogenated BCC mixture was found to be 71.2 ± 5 kJ mol H2−1 using the Kissinger method. This represents a significant decrease compared to the pure MgH2 (179.7 ± 5 kJ mol H2−1), suggesting that the catalytic effect of the BCC alloy significantly decreases the activation energy of MgH2 for dehydrogenation by surface activation.  相似文献   

7.
An MgH2 + 1 mol% Nb2O5 system was modified by heptane and acetone through a high-energy ball milling process, and their rehydrogenation performances were investigated. XRD results indicated that except MgH2 and Nb2O5 phases Mg and MgO phases existed after ball milling. The rehydrogenation results showed that after modification by heptane the capacity increased from 3.0 wt% and 4.2 wt% to 5.0 wt% and 5.5 wt% within 110 s at 523 K and 573 K, respectively. The hydriding rate increased from 0.08 wt%/s after 20 s to 0.22 wt%/s after 10 s at 523 K. However, after modification by acetone it only absorbed 1.8 wt% and 2.0 wt% of hydrogen even within 8000 s at 523 K and 573 K, respectively. Rietveld refinement results indicated that after modification by the heptane the content of MgO was reduced from 6.8 wt% to 4.2 wt%, while after the modification by the acetone the content of MgO was significantly increased from 6.8 wt% to 23.8 wt%. The difference in the rehydrogenation performance was believed to be attributed to the different contents of the MgO phase, which led to the difference in the contents of the MgH2 phase. It implied that the heptane acted as a solvent without oxygen element in it to prevent the MgH2 + Nb2O5 system from aggregation, crystallization and oxidation. It suggested heptane was suitable for the improvement of the rehydrogenation performance of MgH2 system.  相似文献   

8.
Chemical effects of added CO2 on flame extinction characteristics are numerically studied in H2/CO syngas diffusion flames diluted with CO2. The two representative syngas flames of 80% H2 + 20% CO and 20% H2 + 80% CO are inspected according to the composition of fuel mixture diluted with CO2 and global strain rate. Particular concerns are focused on impact of chemical effects of added CO2 on flame extinction characteristics through the comparison of the flame characteristics between well-burning flames far from extinction limit and flames at extinction. It is seen that chemical effects of added CO2 reduce critical CO2 mole fraction at flame extinction and thus extinguish the flame at higher flame temperature irrespective of global strain rate. This is attributed by the suppression of the reaction rate of the principal chain branching reaction through the augmented consumption of H-atom from the reaction CO2 + H→CO + OH. As a result the overall reaction rate decreases. These chemical effects of added CO2 are similar in both well-burning flames far from extinction limit and flames at extinction. There is a mismatching in the behaviors between critical CO2 mole fraction and maximum flame temperature at extinction. This anomalous phenomenon is also discussed in detail.  相似文献   

9.
To improve the dehydrogenation properties of MgH2, a novel hydrogen storage system, MgH2–Li3AlH6, is prepared by mechanochemical milling. Three physical mixtures containing different mole ratios (1:4, 1:1 and 4:1) of MgH2 and Li3AlH6 are studied and there exists a mutual destabilization effect between the components. The last mixture shows a capacity of 6.5 wt% H2 with the lowest starting temperature of dehydrogenation (170 °C). First, Li3AlH6 decomposes into Al, LiH and H2, and then the as-formed Al can easily destabilize MgH2 to form the intermetallic compound Mg17Al12 at a temperature of 235 °C, which is about 180 °C lower than the decomposition temperature of pristine MgH2. Finally, the residual MgH2 undergoes a self-decomposition whose apparent activation energy has been reduced by about 22 kJ mol−1 compared with pristine MgH2. At a constant temperature of 250 °C, the mixture can dehydrogenate completely under an initial vacuum and rehydrogenate to form MgH2 under 2 MPa H2, showing good cycle stability after the first cycle with a capacity of 4.5 wt% H2. The comparison between 4 MgH2 + Li3AlH6 and 4 MgH2 + LiAlH4 mixtures is also investigated.  相似文献   

10.
A new type of Li1−xFe0.8Ni0.2O2–LixMnO2 (Mn/(Fe + Ni + Mn) = 0.8) material was synthesized at 350 °C in air atmosphere using a solid-state reaction. The material had an XRD pattern that closely resembled that of the original Li1−xFeO2–LixMnO2 (Mn/(Fe + Mn) = 0.8) with much reduced impurity peaks. The Li/Li1−xFe0.8Ni0.2O2–LixMnO2 cell showed a high initial discharge capacity above 192 mAh g−1, which was higher than that of the parent Li/Li1−xFeO2–LixMnO2 (186 mAh g−1). We expected that the increase of initial discharge capacity and the change of shape of discharge curve for the Li/Li1−xFe0.8Ni0.2O2–LixMnO2 cell is the result from the redox reaction from Ni2+ to Ni3+ during charge/discharge process. This cell exhibited not only a typical voltage plateau in the 2.8 V region, but also an excellent cycle retention rate (96%) up to 45 cycles.  相似文献   

11.
Hydride nanocomposites in the (LiNH2 + nMgH2) system have been synthesized by ball milling with varying input of milling energy injected into powder particles, QTR (kJ/g). The grain (crystallite) size of LiNH2 and MgH2 decreases rapidly with increasing QTR up to approximately 150–200 kJ/g and subsequently more or less saturates at the value of 10–20 nm. For the injected energy QTR ≈ 250–350 kJ/g the specific surface area (SSA) increases from the initial 2.4 m2/g for powder mixtures before milling to 30–37 m2/g for nanocomposites after milling. After injecting QTR ≈ 550 kJ/g there is a further increase of SSA to 52 m2/g which is over 20-fold increase of SSA from its initial value. That clearly indicates that a profound reduction of particle size has occurred. The hydride phases formed during ball milling with relatively low QTR are identified as a-Mg(NH2)2 (amorphous magnesium imide) and LiH. The ball milled (LiNH2 + nMgH2) nanocomposite system with n = 0.5–0.9 can effectively desorb about 4–5 wt.% H2 with a reasonable rate at the temperature range close to 200 °C. Within a low temperature range up to ∼250 °C, regardless of the molar ratio n and the injected energy QTR the thermal desorption of the (LiNH2 + nMgH2) nanocomposites occurs without any release of ammonia, NH3. For all molar ratios, n, the hydride nanocomposites are fully reversible at 175 °C under a relatively mild pressure of 50 bar H2. The quantity of H2 desorbed decreases with increasing molar ratio n, due to increasing fraction of inactive, retained MgH2. However, at 125 °C the dehydrogenation rate is very sluggish and the quantity of released H2 is minimal. At the temperature range lower than ∼250 °C dehydrogenation of ball milled nanocomposites occurs through formation of the Li2Mg(NH)2 hydride phase. The value of the measured dehydrogenation enthalpy change of 46.7 kJ/molH2 is relatively low and apparently, it is not responsible for sluggish dehydrogenation at 125 °C. The measurements of thermal conductivity for non-milled powders and ball milled nanocomposites show a dramatic reduction of thermal conductivity after ball milling. It seems that this could be a principal factor responsible for such a low dehydrogenation rate at low temperatures.  相似文献   

12.
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   

13.
The effect of CO2 reactivity on CH4 oxidation and H2 formation in fuel-rich O2/CO2 combustion where the concentrations of reactants were high was studied by a CH4 flat flame experiment, detailed chemical analysis, and a pulverized coal combustion experiment. In the CH4 flat flame experiment, the residual CH4 and formed H2 in fuel-rich O2/CO2 combustion were significantly lower than those formed in air combustion, whereas the amount of CO formed in fuel-rich O2/CO2 combustion was noticeably higher than that in air. In addition to this experiment, calculations were performed using CHEMKIN-PRO. They generally agreed with the experimental results and showed that CO2 reactivity, mainly expressed by the reaction CO2 + H → CO + OH (R1), caused the differences between air and O2/CO2 combustion under fuel-rich condition. R1 was able to advance without oxygen. And, OH radicals were more active than H radicals in the hydrocarbon oxidation in the specific temperature range. It was shown that the role of CO2 was to advance CH4 oxidation during fuel-rich O2/CO2 combustion. Under fuel-rich combustion, H2 was mainly produced when the hydrocarbon reacted with H radicals. However, the hydrocarbon also reacted with the OH radicals, leading to H2O production. In fact, these hydrocarbon reactions were competitive. With increasing H/OH ratio, H2 formed more easily; however, CO2 reactivity reduced the H/OH ratio by converting H to OH. Moreover, the OH radicals reacted with H2, whereas the H radicals did not reduce H2. It was shown that OH radicals formed by CO2 reactivity were not suitable for H2 formation. As for pulverized coal combustion, the tendencies of CH4, CO, and H2 formation in pulverized coal combustion were almost the same as those in the CH4 flat flame.  相似文献   

14.
The understanding of hydrogen bonding in magnesium and magnesium based alloys is an important step toward its prospective use. In the present study, a density functional theory (DFT) based, full-potential augmented plane waves method of calculation, extended with local orbitals (FP-APW+lo), was used to investigate the stability of MgH2 and MgH2:TM (TM = Ti and Co) 10 wt % alloys and the influence of this alloying on hydrogen storage properties of MgH2 compound. Effects of a possible spin polarisation induced in the system by transition metal (TM) ions were considered too. It has been found that TM-H bonding is stronger than the Mg–H bond, but at the same time it weakens other bonds in the second and third coordination around a TM atom, which leads to overall destabilization of the MgH2 compound. Due to a higher number of d-electrons, this effect is more pronounced for Co alloying, where in addition, the spin polarisation has a noticeable and stabilising influence on the compound structure.  相似文献   

15.
Submicron-sized LiNi1/3Co1/3Mn1/3O2 cathode materials were synthesized using a simple self-propagating solid-state metathesis method with the help of ball milling and the following calcination. A mixture of Li(ac)·2H2O, Ni(ac)2·4H2O, Co(ac)2·4H2O, Mn(ac)2·4H2O (ac = acetate) and excess H2C2O4·2H2O was used as starting material without any solvent. XRD analyses indicate that the LiNi1/3Co1/3Mn1/3O2 materials were formed with typical hexagonal structure. The FESEM images show that the primary particle size of the LiNi1/3Co1/3Mn1/3O2 materials gradually increases from about 100 nm at 700 °C to 200–500 nm at 950 °C with increasing calcination temperature. Among the synthesized materials, the LiNi1/3Co1/3Mn1/3O2 material calcined at 900 °C exhibits excellent electrochemical performance. The steady discharge capacities of the material cycled at 1 C (160 mA g−1) rate are at about 140 mAh g−1 after 100 cycles in the voltage range 3–4.5 V (versus Li+/Li) and the capacity retention is about 87% at the 350th cycle.  相似文献   

16.
Studies of the electrochemical behavior of K0.27MnO2·0.6H2O in K2SO4 show the reversible intercalation/deintercalation of K+-ions in the lattice. An asymmetric supercapacitor activated carbon (AC)/0.5 mol l−1 K2SO4/K0.27MnO2·0.6H2O was assembled and tested successfully. It shows an energy density of 25.3 Wh kg−1 at a power density of 140 W kg−1; at the same time it keeps a very good rate behavior with an energy density of 17.6 Wh kg−1 at a power density of 2 kW kg−1 based on the total mass of the active electrode materials, which is higher than that of AC/0.5 mol l−1 Li2SO4/LiMn2O4. In addition, this asymmetric supercapacitor shows excellent cycling behavior without the need to remove oxygen from the electrolyte solution. This can be ascribed in part to the stability of the lamellar structure of K0.27MnO2·0.6H2O. This asymmetric aqueous capacitor has great promise for practical applications due to high energy density at high power density.  相似文献   

17.
A novel Ba0.5Sr0.5Co0.8Fe0.2O3 − δ + LaCoO3 (BSCF + LC) composite oxide was investigated for the potential application as a cathode for intermediate-temperature solid-oxide fuel cells based on a Sm0.2Ce0.8O1.9 (SDC) electrolyte. The LC oxide was added to BSCF cathode in order to improve its electrical conductivity. X-ray diffraction examination demonstrated that the solid-state reaction between LC and BSCF phases occurred at temperatures above 950 °C and formed the final product with the composition: La0.316Ba0.342Sr0.342Co0.863Fe0.137O3 − δ at 1100 °C. The inter-diffusion between BSCF and LC was identified by the environmental scanning electron microscopy and energy dispersive X-ray examination. The electrical conductivity of the BSCF + LC composite oxide increased with increasing calcination temperature, and reached a maximum value of ∼300 S cm−1 at a calcination temperature of 1050 °C, while the electrical conductivity of the pure BSCF was only ∼40 S cm−1. The improved conductivity resulted in attractive cathode performance. An area-specific resistance as low as 0.21 Ω cm2 was achieved at 600 °C for the BSCF (70 vol.%) + LC (30 vol.%) composite cathode calcined at 950 °C for 5 h. Peak power densities as high as ∼700 mW cm−2 at 650 °C and ∼525 mW cm−2 at 600 °C were reached for the thin-film fuel cells with the optimized cathode composition and calcination temperatures.  相似文献   

18.
In order to increase the hydrogen storage capacity of Mg-based materials, a mixture with a composition of 2LiBH4 + MgF2 and LiBH4, which has a hydrogen storage capacity of 18.4 wt%, were added to MgH2. Ti isopropoxide was also added to MgH2 as a catalyst. A MgH2 composite with a composition of 40 wt%MgH2 + 25 wt%LiBH4 + 30 wt% (2LiBH4 + MgF2) + 5 wt%Ti isopropoxide (corresponding to 40 wt%MgH2 + 37 wt%LiBH4 + 18 wt%MgF2 + 5 wt%Ti isopropoxide) was prepared by reactive mechanical grinding. The hydrogen storage properties of the sample were then examined. Hydrogen content vs. desorption time curves for consecutive 1st desorptions of 40 wt%MgH2 + 37 wt%LiBH4 + 18 wt%MgF2 + 5 wt%Ti isopropoxide from room temperature to 823 K showed that the total desorbed hydrogen quantity for consecutive 1st desorptions was 8.30 wt%.  相似文献   

19.
The In-doped HLaNb2O7 oxide semiconductors synthesized by solid-state reaction followed by an ion-exchange reaction were found to be a novel composite photocatalyst system with enhanced activity for water splitting. Pt was incorporated in the interlayer of In-doped HLaNb2O7 by the stepwise intercalation reaction. The In-doped HLaNb2O7 powder samples were characterized with X-ray diffraction (XRD) and UV-vis diffuse reflectance spectrometry. The photocatalytic activities of Pt-loaded In-doped HLaNb2O7 and individual precursor materials were evaluated by H2 evolution from aqueous CH3OH solution under UV light irradiation. It was found that the composite In-doped HLaNb2O7 showed a higher H2 evolution rate in comparison with individual materials. The hydrogen production activity of In-doped HLaNb2O7 was greatly enhanced by Pt co-incorporation. The In content in the In-doped HLaNb2O7 system was discussed in relation to the photophysical and photocatalytic properties. As In content equal 5 mol%, the HLaNb2O7:In/Pt showed a photocatalytic activity of 354 cm3 g−1 hydrogen evolution in 10 vol% methanol solution under irradiation from a 100 W mercury lamp at 333 K for 3 h.  相似文献   

20.
The effect of mesoporous Co3O4, NiCo2O4 and NiO on the hydrogen sorption performance of MgH2 was investigated. These oxides were synthesized by multi-step nanocasting and introduced during the high-energy ball milling of MgH2 powder to act as catalysts. Hydrogen desorption on the as-milled powders was assessed upon heating the samples from room temperature to 400 °C. In all cases, the onset temperature for desorption was lowered by taking advantage of the introduced additives. The NiO-doped sample displayed the best response, the desorption rate being 7 times faster than in pure MgH2. Complementary kinetic studies on this particular sample revealed that the sorption activation energies were much lower (50 kJ/mol for absorption and 335 kJ/mol for desorption) than the corresponding ones for undoped MgH2 (57 kJ/mol for absorption and 345 kJ/mol for desorption), thus proving the catalytic activity of the mesoporous NiO oxide. Significantly, the X-ray powder diffraction (XRPD) patterns taken on the NiO-doped sample after discharging/charging cycles revealed that Mg could fully hydrogenate at the end of the charging process, while Mg metal was still detected in the undoped (pure) sample. Favored conditions for dissociative chemisorption of hydrogen could be ascribed to the formation of metallic Ni arising from complete or partial reduction of NiO, as observed in the XRPD patterns.  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