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 0 毫秒
1.
The effect of the vanadium content on the cyclic stability of V–Ti binary alloys was investigated. V1−xTix, x = 0.2 and 0.5 samples were hydrogenated and dehydrogenated at 410 K and 553 K respectively, for more than 100 times. During hydrogen cycling, reduction in the reversible hydrogen storage capacity was clearly observed from both samples. No prominent V-effect was found. In fact, the reduction rates of two samples were similar; both samples showed a ∼25% reduction in the reversible hydrogen storage capacity after 100 cycles. In addition, the shape of the pressure–composition-isotherm (PCT) curves was significantly altered over the testing cycle period; the absorption and desorption plateaus got markedly inclined and the hysteresis became evidently smaller. We found that even after the hydrogen storage capacity of V1−xTix was significantly reduced, at low enough temperature V1−xTix was able to absorb hydrogen as much as it did at the first cycle. Furthermore, the reversible hydrogen storage capacity of V0.8Ti0.2 at 410 K was recovered to a certain degree after hydrogenating the sample at low temperatures.  相似文献   

2.
Pd capped MgxTi1−x films have been prepared by magnetron sputtering, and their electrochemical hydrogen storage properties have been investigated. Results show that the Mg0.85Ti0.15 and Mg0.72Ti0.28 samples exhibit the most promising electrochemical properties, including short activation period, large discharge capacities, excellent cyclic stabilities and superior anti-corrosion behaviors. The maximum capacity of Mg0.85Ti0.15 film is achieved to ∼1100 mAh g−1 after 30 cycles, 80% of which (∼810 mAh g−1) can be maintained even after 150 cycles. Such excellent properties make Pd capped MgxTi1−x films competitive candidates as anode materials of alkaline secondary batteries.  相似文献   

3.
The effect of Mg content on the structural characteristics and hydrogen storage properties of the Ca3.0−xMgxNi9 (x = 0.5, 1.0, 1.5 and 2.0) alloys was investigated. The lattice parameters and unit cell volume of the PuNi3-type (Ca, Mg)Ni3 main phase decreased with increasing Mg content. The 6c site of PuNi3-type structure was occupied by both Ca and Mg atoms. Moreover, the occupation factor of Ca on the 6c site decreased with the increase of Mg content. The hydrogen absorption capacity of the alloys decreased due to higher Mg content. However, the thermodynamic properties of hydrogen absorption and desorption were improved and the plateau pressures were increased. When x = 1.5–2.0, the Ca3.0−xMgxNi9 alloys had favorable enthalpy (ΔH) and entropy (ΔS) of hydride formation.  相似文献   

4.
ZrCo1−xNix (x = 0, 0.1, 0.2 and 0.3) alloys were prepared and their hydrogen storage behavior were studied. ZrCo1−xNix alloys of compositions with x = 0, 0.1, 0.2 and 0.3 prepared by arc-melting method and characterized by X-ray diffraction analysis. XRD analysis showed that the alloys of composition with x = 0, 0.1, 0.2 and 0.3 forms cubic phase similar to ZrCo with traces of ZrCo2 phase. A trace amount of an additional phase similar to ZrNi was found for the alloy with composition x = 0.3. Hydrogen desorption pressure–composition–temperature (PCT) measurements were carried out using Sievert's type volumetric apparatus and the hydrogen desorption pressure–composition isotherms (PCIs) were generated for all the alloys in the temperature range of 523–603 K. A single sloping plateau was observed for each isotherm and the plateau pressure was found to increase with increasing Ni content in ZrCo1−xNix alloys at the same experimental temperature. A van't Hoff plot was constructed using plateau pressure data of each pressure–composition isotherm and the thermodynamic parameters were calculated for desorption of hydrogen in the ZrCo1−xNix–H2 systems. The enthalpy and entropy change for desorption of hydrogen were calculated. In addition, the hydrogen absorption–desorption cyclic life studies were performed on ZrCo1−xNix alloys at 583 K up to 50 cycles. It was observed that with increasing Ni content the durability against disproportionation of alloys increases.  相似文献   

5.
New photocatalysts of Sb2TixSy (x = 0, 0.5, 1.0, 1.5 mol and y = 3, 4, 5, 6 mol) fan blade-like core-shell nanorods have been designed ultimately to enhance hydrogen production. The nanorods of 500 nm long and 60–100 nm wide are Sb2S3 nanorod surrounded by an amorphous TiS2 membrane, showing absorption band edges of above 600 nm. The evolution of H2 from methanol/water (1:1) photo-splitting over Sb2TixSy nanorods in the liquid system is doubled, compared to that over pure Sb2S3. Particularly, 52 μmol of H2 gas is produced after 10 h when 0.5 g of Sb2Ti1.0S5 nanorods is used at pH = 7, and the performance is increased by more than 50% at higher pH. Based on cyclic voltammetry (CV) and UV-Visible absorption spectra, the high photocatalytic activity can be attributed to the existence of an appropriate band-gap state, which includes the scope of the redox potential of water in Sb2Ti1.0S5 nanorods, resulting in the promotion of the redox reaction of water.  相似文献   

6.
Photocatalytic hydrogen production was investigated over ZnS1−x−0.5yOx(OH)y-ZnO using sulfide ion (Na2S-Na2SO3) as an electron donor from NaCl saltwater. NaCl can affect markedly the activity for photocatalytic hydrogen production, depending on NaCl concentration. When NaCl concentration is lower, the activity is lower than that in pure water, whereas when NaCl concentration is higher, the activity is higher than that in pure water. NaCl decreases not only the surface charge of ZnS1−x−0.5yOx(OH)y-ZnO but also the surface hydration. When ZnS1−x−0.5yOx(OH)y-ZnO was impregnated with the electron donor (Na2S-Na2SO3), ZnO was transformed partly into ZnS. The impregnated ZnS1−x−0.5yOx(OH)y-ZnO exhibits higher activity than the non-impregnated one. The possible mechanisms were discussed.  相似文献   

7.
A two-dimensional sample array synthesis has been used to screen carbon-coated Li(1−x)Mgx/2FePO4 and LiFe(1−y)MgyPO4 powders as potential positive electrode materials in lithium ion batteries with respect to x, y and carbon content. The synthesis route, using sucrose as a carbon source as well as a viscosity-enhancing additive, allowed introduction of the Mg dopant from solution into the sol–gel pyrolysis precursor. High-throughput XRD and cyclic voltammetry confirmed the formation of the olivine phase and percolation of the electronic conduction path at sucrose to phosphate ratios between 0.15 and 0.20. Measurements of the charge passed per discharge cycle showed that the capacity deteriorated on increasing magnesium in Li(1−x)Mgx/2FePO4, but improved with increasing magnesium in LiFe(1−y) MgyPO4, especially at high scan rates. Rietveld-refined XRD results on samples of LiFe(1−y)MgyPO4 prepared by a solid-state route showed a single phase up to y = 0.1 according to progressive increases in unit cell volume with increases in y. Carbon-free samples of the same materials showed conductivity increases from 10−10 to 10−8 S cm−1 and a decrease of activation energy from 0.62 to 0.51 eV. Galvanostatic cycling showed near theoretical capacity for y = 0.1 compared with only 80% capacity for undoped material under the same conditions.  相似文献   

8.
NiS was loaded on ZnS1−x−0.5yOx(OH)y–ZnO photocatalyst by two methods (depositing NiS on the photocatalyst via impregnation with a NiS sol and via a precipitation reaction). The activity of the NiS-modified photocatalyst for hydrogen evolution from pure water and seawater was investigated using mixed electron donors (sulfide and sulfite). The activity (in pure water-electron donor solution) is dependant of the modification method and loading content of NiS. When Na2S·9H2O and Na2SO3 are added into seawater, a mixture precipitate (Mg(OH)2 and CaSO3) produces. The precipitate is detrimental to the photocatalytic hydrogen evolution over the unmodified ZnS1−x−0.5yOx(OH)y–ZnO. However, for the NiS-modified photocatalyst, the detrimental effect decreases notably. Interestingly, the activity of the modified photocatalyst in seawater-electron donor solution (with the precipitate) is higher than that in the pure water-electron-donor solution (without the precipitate). The possible mechanism was discussed.  相似文献   

9.
Three series of Ti–Cr–Mn–Fe based alloys with high hydrogen desorption plateau pressures for hybrid hydrogen storage vessel application were prepared by induction levitation melting, as well as their crystallographic characteristics and hydrogen storage properties were investigated. The results show that all of the alloys were determined as a single phase of C14-type Laves structure. As the Fe content in the TiCr1.9−xMn0.1Fex (x = 0.4–0.6) alloys increases, the hydrogen absorption and desorption plateau pressures increase, and the hydrogen storage capacity and plateau slope factor decrease respectively. The same trends are observed when increasing the Mn content in the TiCr1.4−yMnyFe0.6 (y = 0.1–0.3) alloys, except for the plateau slope factor. Compared with the stoichiometric TiCr1.1Mn0.3Fe0.6 alloy, the titanium super-stoichiometric Ti1+zCr1.1Mn0.3Fe0.6 (z = 0.02, 0.04) alloys have larger hydrogen storage capacities and lower hydrogen desorption plateau pressures. Among the studied alloys, Ti1.02Cr1.1Mn0.3Fe0.6 has the best overall properties for hybrid hydrogen storage application. Its hydrogen desorption pressure at 318 K is 41.28 MPa, its hydrogen storage capacity is 1.78 wt.% and its dissociation enthalpy (ΔHd) is 16.24 kJ/mol H2.  相似文献   

10.
Cd1−xZnxS solid solutions (x = 0.05–0.3) supported on mesoporous silica SBA-16 substrate with 3D cubic structure were investigated for hydrogen production from water splitting under visible light. The influence of Zn concentration (x) in the Cd1−xZnxS solid solution and support morphology were investigated. The bare SBA-16 substrate was synthetized by the hydrothermal method whereas the Cd1−xZnxS photocatalysts were prepared by coprecipitation of metal sulfides from aqueous solutions of Cd2+ and Zn2+ using Na2S as precipitating agent. An attempt has been made to determine the photocatalyst structures using several techniques including elemental analysis, N2 adsorption–desorption, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), high resolution transmission electron microscopy (HRTEM), UV–Vis diffuse reflectance spectroscopy (UV–Vis DRS) and Raman spectroscopy. Surface characterization of the samples by XPS indicates that Cd1−xZnxS nanoparticles are unevenly distributed on both external surface and within the pore network. An increase of the band gap energy with increasing Zn loading up to x = 0.2 in the Cd1−xZnxS solid solution was observed. As a consequence, H2 evolution increases gradually with an increase of the Zn loading in the photocatalysts from 0.05 to 0.2 wt% being the Cd0.8Zn0.2S/SBA-16 system the most active among the catalysts studied. The highest activity of this photocatalyst was explained in terms not only of its large band gap energy but also by the enhancement of the interaction between the particles of solid solution and the SBA-16 substrate.  相似文献   

11.
The hydrogen permeability of cast Nb30Ti35Ni35−xCox (x = 0…35) alloys is found to increase with the Co content, induced by the tailored microstructures containing high fractions of eutectic and less primary phase. Among the cast alloys, Nb30Ti35Co35 consisting entirely of eutectic exhibits the highest permeability, particularly 2.65 × 10−8 mol H2 m−1 s−1 Pa−0.5 at 673 K. This permeability is further increased to 3.58 × 10−8 mol H2 m−1 s−1 Pa−0.5 by aligning eutectic grains perpendicularly to the membrane surface using directional solidification. The high permeability is attributed to the high hydrogen solubility and diffusivity in alloys. Our work demonstrates that hydrogen permeable alloys containing high fractions of eutectic may exhibit high permeability by adequately optimizing morphology, volume fraction and alignment of the bcc-Nb phase in the eutectic as the pathways for hydrogen permeation.  相似文献   

12.
Ternary Mg1−xCaxNi2−y solid solutions were synthesized by powder sintering. The phase structures and hydrogen storage properties of the sintered samples were investigated. In a certain range of x and y values, the samples are a single C15 Laves phase with various types of defects. The reduction of Ni content leads to the formation of omission solid solution with vacancies on the sites of Ni. These vacancies increase the hydrogen storage capacity, but decrease the reversibility of hydrogen absorption and desorption.  相似文献   

13.
In situ X-ray diffraction was used to identify the crystal structures of as-deposited and hydrogenated MgyTi100-y thin film alloys containing 70, 80 and 90 at.% Mg. The preferred crystallographic orientation of the films in both the as-prepared and hydrogenated state made it difficult to unambiguously identify the crystal structure up to now. In this work, identification of the unit cells was achieved by in situ recording diffraction patterns at various tilt angles. The results reveal a hexagonal closed packed structure for all alloys in the as-deposited state. Hydrogenating the layers under 105 Pa H2 transforms the unit cell into face-centered cubic for the Mg70Ti30 and Mg80Ti20 compounds, whereas the unit cell of hydrogenated Mg90Ti10 has a body-centered tetragonal symmetry. The (de)hydrogenation kinetics changes along with the crystal structure of the metal hydrides from rapid for fcc-structured hydrides to sluggish for hydrides with a bct symmetry and emphasizes the influence of the crystal structure on the hydrogen transport kinetics.  相似文献   

14.
Polycrystalline TiO2−x pellets were equilibrated at 1220–1420 K in a flow of Ar + 7 vol.% H2 gas mixture. The resulting deviation x from stoichiometric composition of titanium dioxide was determined gravimetrically. X-ray diffractometry revealed coexistence of rutile with TinO2n−1 Magnéli phases. Optical reflectance and photocurrent spectroscopies served as experimental methods of the band gap determination. Polycrystalline TiO2−x were used as photoanodes in a photoelectrochemical cell. The photocurrent response to the applied voltage was studied. It was found that TiO2−x with x ca. 0.006 exhibited the best photoelectrochemical performance.  相似文献   

15.
Hydrogenation properties of LaNi5  xInx alloys (x = 0.1, 0.2 and 0.5) were examined by their direct reaction with gaseous hydrogen and by cathodic charging in 6 M KOH solution. The gas phase measurements were carried out using Sievert's type apparatus in 300–400 K temperature range and at hydrogen pressures up to 40 bars. Indium substitution for Ni in LaNi5 significantly modifies the hydrogenation behavior, decreasing the equilibrium pressure of hydrogen and limiting the hydrogen capacity as compared to LaNi5. The LaNi4.9In0.1 revealed a distinct presence of two pressure plateaus on the high temperature isotherms. Apart from the α-phase (hydrogen solid solution) and β-phase (LaNi5H6 hydride), formation of a new σ*-hydride phase was postulated at the hydrogen content extended over the region of H/f.u. = 1.3–1.8. Thermodynamic functions: enthalpy and entropy of the hydrogen absorption process were calculated from the H2-pressure/composition (p–c) isotherms at several temperatures, applying the Van't Hoff's (lnp − 1/T) dependence. Electrochemical galvanostatic hydrogenation experiments at 185 mA/g charge/discharge rate revealed the greatest discharge current capacity of 319 mAh/g for LaNi4.9In0.1 alloy after 4–5 cycles. The hydrogen discharge capacities decrease with further increase of indium content in the alloy.  相似文献   

16.
Current paper comprises the electrodeposition of nanostructured porous Co1−xNix layered double hydroxide (Co1−xNix LDHs) thin films on to stainless steel substrate by a potentiodynamic mode. The compositional impacts on the various properties of Co1−xNix LDHs are examined via structural, morphological, surface wettability and electrochemical studies. The nanocrystalline Co1−xNix LDHs thin films possess varying porous, nanoflake like morphology and superhydrophilic behavior by the composition influence. Electrochemical studies demonstrate the supercapacitive performance of Co1−xNix LDHs thin film electrodes. The maximal specific capacitance for Co1−xNix LDHs electrode is found to be ∼1213 F g−1 for composition Co0.66Ni0.34 LDH in 2 M KOH electrolyte at 5 mV s−1 scan rate owing specific energy of 104 Whkg−1, specific power of 1.44 kW kg−1 with ∼94% of coulomb efficiency and stability of electrode retained to 77% after 10,000th cycle. The high capacitance retention proposes the deposited Co1−xNix LDHs thin film as promising contender for supercapacitor applications.  相似文献   

17.
A single phase mixed oxide ion-electron conducting electrochemical catalyst of Ce1−xNixO2−y is employed as an anode functional reformation layer for a coking-resistant solid oxide fuel cell (SOFC) based on oxide ion conducting electrolyte operated in methane and ethanol. The high catalytic activity of Ce1−xNixO2−y oxide for fuel reformation is demonstrated by the excellent cell performances in various fuels at relatively low temperatures (550–650 °C). The fast oxygen ions flux and formed steam at anode side are also found to be favorable for hydrocarbon reformation to promote the cell performance and long term stability. At 650 °C, maximum power densities of 415 and 271 mW cm−2 are achieved in methane and ethanol respectively. The resistance against carbon deposition is significantly improved with stable voltage output in a long-term durability operation.  相似文献   

18.
A network of CoxNiyAlz layered triple hydroxides (LTHs) nanosheets was prepared by the potentiostatic deposition process at −1.0 V (vs. Ag/AgCl) onto stainless steel electrodes. X-ray diffraction patterns show that the CoxNiyAlzLTHs belong to the hexagonal system with layered structure. Cyclic voltammetry and charge discharge measurements in the potential range of −0.1 to 0.5 V and 0.0–0.4 V, respectively, vs. Ag/AgCl in 1 M KOH electrolyte indicate that CoxNiyAlzLTHs have excellent supercapacitive characteristics. The maximum specific capacitance of ∼1263 F g−1 was obtained for Co0.59Ni0.21Al0.20LTH. The impedance studies indicated highly conducting nature of the CoxNiyAlzLTHs.  相似文献   

19.
The influence of titanium doping level in Ba0.6Sr0.4Co1−yTiyO3−δ (BSCT) oxides on their phase structure, electrical conductivity, thermal expansion coefficient (TEC), and single-cell performance with BSCT cathodes has been investigated. The incorporation of Ti can lead to the phase transition of Ba0.6Sr0.4CoO3−δ from hexagonal to cubic structure. The solid solution limitation of Ti in Ba0.6Sr0.4Co1−yTiyO3−δ is 0.15–0.3 under 1100 °C. BSCT shows a small polaron conduction behavior and the electrical conductivity increases steadily in the testing temperature range (300–900 °C), leading to a relatively high conductivity at high temperatures. The electrical conductivity decreases with increasing Ti content. The addition of Ti deteriorates the cathode performance of BSCT slightly but decreases the TEC significantly. The TEC of BSCT is about 14 × 10−6 K−1, which results in a good physical compatibility of BSCT with Gd0.2Ce0.8O2−δ (GDC) electrolyte. BSCT also shows excellent thermal cyclic stability of electrical conductivity and good chemical stability with GDC. These properties make BSCT a promising cathode candidate for intermediate temperature solid oxide fuel cells (IT-SOFCs).  相似文献   

20.
Hydrogen absorption/desorption has been investigated in the three series of solid solution bcc alloys Ti35VxCr65−x (x = 18,22), Ti40VxMn50−xCr10 (x = 32,36) and TixCr97.5−xMo2.5 (x = 43,46). It has been found that the H absorption at pressures smaller than 1 bar can only occur after elimination of the oxide films by heating the alloys to temperatures higher than 600 K. Hydrogen desorption from pre-loaded materials (nH = H/Me ≤ 0.27) takes place on heating at much lower temperatures in the Ti40VxMn50−xCr10 and Ti35VxCr65−x than in the TixCr97.5−xMo2.5 alloys. The H diffusion parameters W and Do deduced from high temperature (>450 K) absorption experiments are as follows: W = 0.318 ± 0.005 eV, Do = (4 ± 1)×10−7 m2/s for Ti40VxMn50−xCr10; W = 0.32 ± 0.02 eV, Do = (3 ± 2)×10−7 m2/s for Ti35VxCr65−x; W = 0.79 ± 0.06 eV, Do = (4 ± 2)×10−8 m2/s for TixCr97.5−xMo2.5. The higher value of the activation energy for H diffusion in Mo containing alloys is most likely due to remarkable attractive interactions between H and Mo atoms.  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