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.
2.
The effect of Ti, Co, Ni and LaCl3 on hydrogen release of NaAlH4 was studied by pressure-content-temperature (PCT) equipment. The result showed that the sample doped with 3 mol% LaCl3 presented the largest amount of hydrogen release. Increasing the amount of LaCl3 from 1 to 6 mol% caused such marked changes in behavior that the amount and rate of hydrogen release increased first and then decreased. In addition, the study on the rehydrogenation temperature of NaAlH4 doped with 3 mol% LaCl3 showed that the doped sample during the first rehydrogenation cycle carried out in PCT at 110 °C under 8 MPa after being discharged of hydrogen at 270 °C presented the largest amount of hydrogen release.  相似文献   

3.
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   

4.
Nanocrystalline titanium dioxide/carbon composite (TiO2/C) was synthesized through a direct solution-phase carburization using tetrabutyl titanate (Ti(OBu)4) and resol as precursors. The prepared TiO2/C composite was mainly in the anatase structure with an average particle size under 20 nm, which was then introduced in NaAlH4 as a catalyst through ball milling. The desorption curves show that both nanocrystalline TiO2/C and TiO2 can obviously improve the kinetics of NaAlH4, while NaAlH4 with 3 mol% TiO2/C exhibits better cycling stability than NaAlH4 with 3 mol%TiO2. The hydrogen storage capacity of NaAlH4 with TiO2/C remains stable after 5th cycle, and about 94% of initial hydrogen is released, while the capacity of NaAlH4 with TiO2 decreases continuously during cycling, and only 88% of initial hydrogen is released after 10th cycle. Furthermore, NaAlH4 with 3 mol%TiO2/C exhibits good reversibility at relatively low hydrogen pressures, and it can reload 4.16 and 1.63wt% hydrogen at 50 and 30 bar hydrogen pressures, respectively.  相似文献   

5.
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   

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.
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   

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.
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   

10.
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   

11.
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   

12.
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   

13.
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   

14.
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   

15.
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   

16.
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   

17.
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   

18.
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   

19.
It is well known that the dehydrogenation pathway of the LiBH4–MgH2 composite system is highly reliant on whether decomposition is performed under vacuum or a hydrogen back-pressure. In this work, the effects of hydrogen back-pressure and NbF5 addition on the dehydrogenation kinetics of the LiBH4–MgH2 system are studied under either vacuum or hydrogen back-pressure, as well as the subsequent rehydrogenation and cycling. For the pristine sample, faster desorption kinetics was obtained under vacuum, but the performance is compromised by slow absorption kinetics. In contrast, hydrogen back-pressure remarkably promotes the absorption kinetics and increases the reversible hydrogen storage capacity, but with the penalty of much slower desorption kinetics. These drawbacks were overcome after doping with NbF5, with which the dehydrogenation and rehydrogenation kinetics was significantly improved. In particular, the enhanced kinetics was observed to persist well, even after 9 cycles, in the case of the NbF5 doped sample under hydrogen back-pressure, as well as the suppression of forming Li2B12H12. Furthermore, the mechanism that is behind these effects of NbF5 additive on the reversible dehydrogenation reaction of the LiBH4–MgH2 system is discussed.  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