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 843 毫秒
1.
Roy D. Raharjo  Donald R. Paul 《Polymer》2007,48(25):7329-7344
Pure and mixed gas n-C4H10 and CH4 permeability coefficients in poly(1-trimethylsilyl-1-propyne) (PTMSP) are reported at temperatures from −20 to 35 °C. CH4 partial pressures range from 1.1 to 14.6 atm, and n-C4H10 partial pressures range from 0.02 to 1.8 atm. CH4 permeability decreases with increasing n-C4H10 upstream activity (f/fsat) in the feed. For example, at −20 °C, CH4 permeability decreases by more than an order of magnitude, from 52,000 to 1700 Barrer, as n-C4H10 activity increases from 0 to 0.73. In contrast, n-C4H10 mixed gas permeability is essentially unaffected by the presence of CH4. The depression of CH4 permeability in mixtures is a result of competitive sorption and blocking effects, which reduce both CH4 mixture solubility and diffusivity, respectively. Diffusion coefficients of n-C4H10 and CH4 in mixtures were calculated from mixture permeability and mixture solubility data. The CH4 concentration-averaged diffusion coefficient generally decreases as n-C4H10 activity increases. On the other hand, the n-C4H10 diffusion coefficient is essentially unaffected by the presence of CH4. Pure and mixed gas activation energies of permeation and diffusion of CH4 and n-C4H10 are reported. The mixed gas n-C4H10/CH4 permeability selectivity increases with increasing n-C4H10 activity and decreasing temperature, and it is higher than pure gas estimates would suggest. Mixture diffusivity selectivity also increases with increasing n-C4H10 activity. The difference between pure and mixed gas permeability selectivity arises from both solubility and diffusivity effects. The dual mode mixed gas permeability model describes the mixture permeability data reasonably well for n-C4H10. However, the model must be modified to accurately describe the methane data by accounting for the decrease in methane diffusivity due to the presence of n-C4H10 (i.e., blocking). Even though the penetrant concentrations are rather significant at some of the conditions considered, no evidence is observed for phenomena such as multicomponent coupling that would require a model more complex than the binary form of Fick's law. That is, Fick's law in its simplest form adequately describes the experimental data.  相似文献   

2.
Poly(1-trimethylsilyl-1-propyne) (PTMSP) has been crosslinked using 3,3′-diazidodiphenylsulfone to improve its solvent resistance. This study reports the influence of crosslinker content on the solubility properties of PTMSP, its density, and its gas sorption and transport properties. Crosslinking PTMSP renders it insoluble even in excellent solvents for the uncrosslinked polymer. Gas permeability and fractional free volume (FFV) decreased as crosslinker content increased, while gas sorption was unaffected by crosslinking. Therefore, the reduction in permeability upon crosslinking PTMSP was due to decreases in diffusion coefficients. Permeability reductions due to crosslinking could be offset by adding nanoparticles to the films. The addition of 30 wt.% fumed silica nanoparticles increased the permeability of crosslinked PTMSP by approximately 80%. In mixed gas permeation experiments, when the composition of the feed gas was 98 mol% CH4 and 2 mol% n-C4H10, uncrosslinked PTMSP had an n-C4H10/CH4 selectivity of 31 and an n-C4H10 permeability of 114,000 barrers at 35 °C and 14 atm feed fugacity. At the same conditions, crosslinked PTMSP containing 5 wt.% crosslinker had an n-C4H10/CH4 selectivity of 28 and an n-C4H10 permeability of 73,000 barrers, and crosslinked PTMSP containing 5 wt.% crosslinker and 30 wt.% fumed silica nanoparticles had an n-C4H10/CH4 selectivity of 21 and an n-C4H10 permeability of 110,000 barrers.  相似文献   

3.
Crosslinking poly[1-(trimethylsilyl)-1-propyne] (PTMSP) films with 3,3′-diazidodiphenylsulfone, a bis(azide) crosslinker, rendered the films insoluble in common solvents for PTMSP such as toluene. At all temperatures, mixed gas CH4 and n-C4H10 permeabilities of crosslinked PTMSP were less than those of uncrosslinked PTMSP, which correlates with lower free volume in the crosslinked material. The presence of fumed silica (FS) nanoparticles in both uncrosslinked PTMSP and crosslinked PTMSP increased mixed gas CH4 and n-C4H10 permeabilities, consistent with the disruption of polymer chain packing by such nanoparticles. Mixed gas CH4 permeabilities of all films were significantly less than their corresponding pure gas CH4 permeabilities. For example, at 35 °C, the mixed gas CH4 permeabilities were approximately 60–80% less than their pure gas values. The greatest decrease was observed for uncrosslinked PTMSP, while nanocomposite PTMSP films showed the least decrease. The mixed gas n-C4H10/CH4 selectivities of crosslinked PTMSP and nanocomposite PTMSP films were less than those of uncrosslinked PTMSP at all temperatures. For example, at 35 °C, the mixed gas n-C4H10/CH4 selectivities of uncrosslinked PTMSP, crosslinked PTMSP containing 10 wt% crosslinker, and uncrosslinked PTMSP containing 30 wt% FS were 33, 27, and 17, respectively, when the feed gas contained 2 mol% n-C4H10 and the total upstream mixture fugacity was 11 atm. For all films, as temperature decreased, mixed gas n-C4H10 permeabilities increased, and mixed gas CH4 permeabilities decreased. Consequently, the mixed gas n-C4H10/CH4 selectivities increased substantially as temperature decreased and the mixed gas selectivity of uncrosslinked PTMSP increased from 33 to 170 as temperature decreased from 35 °C to ?20 °C when the feed gas contained 2 mol% n-C4H10 and the total upstream mixture fugacity was 11 atm.  相似文献   

4.
Using a manometric experimental setup, high-pressure sorption measurements with CH4 and CO2 were performed on three Chinese coal samples of different rank (VRr = 0.53%, 1.20%, and 3.86%). The experiments were conducted at 35, 45, and 55 °C with pressures up to 25 MPa on the 0.354-1 mm particle fraction in the dry state. The objective of this study was to explore the accuracy and reproducibility of the manometric method in the pressure and temperature range relevant for potential coalbed methane (CBM) and CO2-enhanced CBM (CO2-ECBM) activities (P > 8 MPa, T > 35 °C). Maximum experimental errors were estimated using the Gauss error propagation theorem, and reproducibility tests of the high-pressure sorption measurements for CH4 and CO2 were performed. Further, the experimental data presented here was used to explicitly study the CO2 sorption behaviour of Chinese coal samples in the elevated pressure range (up to 25 MPa) and the effects of temperature on supercritical CO2 sorption isotherms.The experiments provided characteristic excess sorption isotherms which, in the case of CO2 exhibit a maximum around the critical pressure and then decline and level out towards a constant value. The results of these manometric tests are consistent with those of previous gravimetric sorption studies and corroborate a crossover of the 35, 45, and 55 °C CO2 excess sorption isotherms in the high-pressure range. The measurement range could be extended, however, to significantly higher pressures. The excess sorption isotherms tend to converge, indicating that the temperature dependence of CO2 excess sorption on coals at high-pressures (>20 MPa) becomes marginal. Further, all CO2 high-pressure isotherms measured in this study were approximated by a three-parameter excess sorption function with special consideration of the density ratio of the “free” phase and the sorbed phase. This function provided a good representation of the experimental data.The maximum excess sorption capacity of the three coal samples for methane ranged from 0.8 to 1.6 mmol/g (dry, ash-free) and increased from medium volatile bituminous to subbituminous to anthracite. The medium volatile bituminous coal also exhibited the lowest overall excess sorption capacity for CO2. However, the subbituminous coal was found to have the highest CO2 sorption capacity of the three samples. The mass fraction of adsorbed substance as a function of time recorded during the first pressure step was used to analyze the kinetics of CH4 and CO2 sorption on the coal samples. CO2 sorption proceeds more rapidly than CH4 sorption on the anthracite and the medium volatile bituminous coal. For the subbituminous coal, methane sorption is initially faster, but during the final stage of the measurement CO2 sorption approaches the equilibrium value more rapidly than methane.  相似文献   

5.
Partial oxidative gasification of n-hexadecane (n-C16) and organosolv-lignin (lignin) was studied by use of a batch type reactor in supercritical water: 673 K, 0.52 cm−3 of water density (40 MPa of water pressure at 673 K), and 0.3 of O/C ratio for the n-C16 experiments; 673 K, 0.35 cm−3 of water density (30 MPa of water pressure at 673 K), and 1.0 of O/C ratio for the lignin experiments. The experiments without O2 were also conducted for lignin (lignin decomposition). For all the cases (n-C16 partial oxidation, lignin decomposition, lignin partial oxidation), NaOH or zirconia (ZrO2) was added in the system as catalysts. Through n-C16 studies, the catalytic effect of NaOH and ZrO2 on partial oxidation in supercritical water were examined. In the case of lignin partial oxidation, we studied the possibility of partial oxidation in supercritical water for gasification technique of wastes. The yield of H2 from n-C16 and lignin with zirconia was twice as same as that without catalyst at the same condition. The H2 yield with NaOH was 4 times higher than that without catalyst. Thus, a base catalyst has a positive effect on partial oxidation of n-C16 and lignin to produce H2. The catalytic effect of NaOH and ZrO2 was found to be enhancement of decomposition of intermediate (aldehyde and ketone) into CO, through n-C16 studies. In the case of lignin studies, the enhancement of decomposition of the carbonyl compounds by catalytic effect of NaOH and ZrO2 inhibit char formation and promotes CO and thus H2 formation.  相似文献   

6.
The alkane n-C198H398 has been crystallised in both extended chain and once-folded forms and annealed to produce materials with low concentrations of gauche bonds. The concentrations of the specific conformers detected by FTIR spectroscopy at −173 °C are calculated, using measurements on liquid n-hexadecane for calibration: values are all generally less than 2.0 per 100 carbon atoms, with extended chain samples showing values less than 1.0 per 100 carbon atoms. A subtraction spectrum (Once-folded chain sample minus Extended chain sample) shows positive bands at 1298, 1340, 1347 and 1369 cm−1, which are predicted in earlier calculations for a (110) fold, while additional positive bands at 1353 and 1363 cm−1 are assigned, respectively, to gg conformers and (tentatively) to strained gtg or gtg′ conformations.  相似文献   

7.
Lihui Cao  Weimin Dong  Xuequan Zhang 《Polymer》2007,48(9):2475-2480
The oxovanadium phosphonates (VO(P204)2 and VO(P507)2) activated by various alkylaluminums (AlR3, R = Et, i-Bu, n-Oct; HAlR2, R = Et, i-Bu) were examined in butadiene (Bd) polymerization. Both VO(P204)2 and VO(P507)2 showed higher activity than those of classical vanadium-based catalysts (e.g. VOCl3, V(acac)3). Among the examined catalysts, the VO(P204)2/Al(Oct)3 system (I) revealed the highest catalytic activity, giving the poly(Bd) bearing Mn of 3.76 × 104 g/mol, and Mw/Mn ratio of 2.9, when the [Al]/[V] molar ratio was 4.0 at 40 °C. The polymerization rate for I is of the first order with respect to the concentration of monomer. High thermal stability of I was found, since a fairly good catalytic activity was achieved even at 70 °C (polymer yield > 33%); the Mn value and Mw/Mn ratio were independent of polymerization temperature in the range of 40-70 °C. By IR and DSC, the poly(Bd)s obtained had high 1,2-unit content (>65%) with atactic configuration. The 1,2-unit content of the polymers obtained by I was nearly unchanged, regardless of variation of reaction conditions, i.e. [Al]/[V], ageing time, and reaction temperature, indicating the high stability of stereospecificity of the active sites.  相似文献   

8.
Members of the solid-solution series Ce1−xSrxPO4−δ (x = 0, 0.01, 0.02) with mixed protonic and electronic transport have been synthesized by a nitrate-decomposition method followed by sintering at 1450 °C. Impedance spectroscopy is employed to estimate the bulk electrical conductivity in wet (∼0.03 atm) and dry atmospheres of O2 and 10%H2:90%N2. Conductivity increases with dopant concentration (x), oxygen partial pressure (pO2) and water vapour partial pressure (pH2O) reaching ∼3.5 × 10−3 S cm−1 at 600 °C for x = 0.02 in wet O2. Activation energies (Ea) for the bulk conductivity of Ce0.98Sr0.02PO4−δ below 650 °C are 0.44 and 0.78 eV for wet oxidising and wet reducing conditions, respectively. A moderate but positive pO2+n power-law dependence (n < 1/10) of conductivity is exhibited in the pO2 range 10−2.5 to 10−1 atm, consistent with mixed ionic and p-type electronic transport. Thermogravimetric analysis indicates that the Sr-doped materials are stable in a CO2 atmosphere in the temperature range 25–1200 °C.  相似文献   

9.
This paper describes the synthesis and characterization of three-dimensional hybrid inorganic-organic networks prepared by a polycondensation reaction between Zr(O(CH2)3CH3)4 and polyethylene glycol 400 (PEG400). Eleven hybrid networks doped with varying concentrations of LiClO4 salt were prepared. On the basis of analytical data and FT-Raman studies it was concluded that these polymer electrolytes consist of inorganic-organic networks with zirconium atoms bonded together by PEG400 bridges. These polymers are transparent with a solid rubber consistency and are very stable under inert atmosphere. Scanning electron microscopy revealed a smooth glassy surface. X-ray fluorescence microanalysis with energy dispersive spectroscopy demonstrated that all the constituent elements are homogeneously distributed in the materials. Thermogravimetric measurements revealed that these materials are thermally stable up to 262 °C. Differential Scanning Calorimetry measurements indicated that the glass transition temperature Tg of these inorganic-organic hybrids varies from −43 to −15 °C with increasing LiClO4 concentration. FT-Raman investigations revealed the TGT (T=trans, G=gauche) conformation of polyether chains and allowed characterization of the types of ion-ion and ion-polymer host interactions in the bulk materials. The conductivity of the materials at different temperatures was determined by impedance spectroscopy over the 20 Hz-1 MHz frequency range. Results indicated that the materials conduct ionically and that their ionic conductivity is strongly influenced by the segmental motion of the polymer network and the type of ionic species distributed in the bulk material. Finally, it is to be highlighted that the hybrid network with a nLi/nO molar ratio of 0.0223 shows a conductivity of ca. 1×10−5 S cm−1 at 40 °C.  相似文献   

10.
The modification of activated carbon fibres prepared from a commercial textile acrylic fibre into materials with monolithic shape using phenolic resin as binder was studied. The molecular sieving properties for the gas separations CO2/CH4 and O2/N2 were evaluated from the gas uptake volume and selectivity at 100 s contact time taken from the kinetic adsorption curves of the individual gases. The pseudo-first order rate constant was also determined by the application of the LDF model. The samples produced show high CO2 and O2 rates of adsorption, in the range 3-35 × 10−3 s−1, and in most cases null or very low adsorption of CH4 and N2 which make them very promising samples to use in PSA systems, or similar. Although the selectivity was very high, the adsorption capacity was low in certain cases. However, the gas uptake in two samples reached 23 cm3 g−1 for CO2 and 5 cm3 g−1 for O2, which can be considered very good. The materials were heat-treated using a microwave furnace, which is a novel and more economic method, when compared with conventional furnaces, to improve the molecular sieves properties.  相似文献   

11.
W.B. Utomo 《Electrochimica acta》2006,51(16):3338-3345
The corrosion of titanium in H2SO4 electrolytes (0.001-1.0 M) at temperatures from ambient to 98 °C has been investigated using steady-state polarization measurements. Four distinct regions of behaviour were identified, namely active corrosion, the active-passive transition, passive region and the dielectric breakdown region. The active corrosion and active-passive transition were characterized by anodic peak current (im) and voltage (Em), which in turn were found to vary with the experimental conditions, i.e., d(log?(im))/dpH=−0.8±0.1 and dEm/dpH which was −71 mV at 98 °C, −58 mV at 80 °C and −28 mV at 60 °C. The activation energy for titanium corrosion, determined from temperature studies, was found to be 67.7 kJ mol−1 in 0.1 M H2SO4 and 56.7 kJ mol−1 in 1.0 M H2SO4. The dielectric breakdown voltage (Ed) of the passive TiO2 film was found to vary depending on how much TiO2 was present. The inclusion of Mn2+ into the H2SO4 electrolyte, as is done during the commercial electrodeposition of manganese dioxide, resulted in a decrease in titanium corrosion current, possibly due to Mn2+ adsorption limiting electrolyte access to the substrate.  相似文献   

12.
Dieter Heymann 《Carbon》2005,43(11):2235-2242
The mean lifetimes of polyyne C8H2 in hexane were determined at 50, 60, 80, and 100 °C and in methanol at 60 °C. The reactions are second order at all temperatures: ln k2 = 20.5 ± 1.5-10303 ± 520T−1 and the corresponding activation energy is 85.7 ± 6.3 kJ mol−1 (7164 cm−1). Extrapolation suggests that solutions at 1 mM concentration are significantly unstable at room temperature. Quantum chemical calculations show that polyynes CmH2 + CnH2 (m + n = 16) could be products, but these were not detected. Alternatively, C16H2 isomers could form. IR spectra of the solid residues from hexane and methanol solutions were obtained.  相似文献   

13.
Cobalt aluminate (CoAl2O4) thin films were grown in a low-pressure hot wall metal organic chemical vapour deposition (MOCVD) reactor on Si(1 0 0) and quartz substrates with a total pressure of 2 Torr using bis(η5-cyclopentadienyl)Co(II) [Co(η5-C5H5)2] and aluminium dimethylisopropoxide [AlMe2(OiPr)] as precursors at 500 and 900 °C. Films showed a dark-brown and dark-green colouration, respectively, and after an overnight heat treatment in air at 1200 °C, they turned blue. Film microstructure, composition and morphology were investigated in detail by X-ray diffraction (XRD), Rutherford backscattering spectroscopy (RBS), scanning electron microscopy (SEM) and secondary ion mass spectrometry (SIMS) analyses. Films were polycrystalline and the UV-vis spectra showed three electronic transitions allowed by the spin (540-630 nm range) characteristic of Co(II) ions with 3d7 configuration in tetrahedral coordination. SEM micrographs of the heat-treated CoAl2O4 samples revealed the presence of agglomerated crystallites with a highly porous structure.  相似文献   

14.
Xue Jiang  Wenshuai Zhu  Huoming Shu 《Fuel》2009,88(3):431-436
Oxidation of dibenzothiphene (DBT) in model oil with H2O2 using surfactant-type decatungstates Q4W10O32 (Q = (CH3)3NC16H33, (CH3)3NC14H29, (CH3)3NC12H25 and (CH3)3NC10H21) as catalysts was studied. The surfactant-type decatungstates were synthesized and characterized. The suitable reaction condition of deep desulfurization was suggested: n(DBT):n(catalyst):n(H2O2) = 1:0.01:3, 60 °C for 0.5 h, under which the DBT conversion can reach 99.6% with [(CH3)3NC16H33]4W10O32 as catalyst. The length of carbon chains of quaternary ammonium cations played a vital role in the catalytic activity of surfactant-type decatungstates, that is, the longer the carbon chain of quaternary ammonium cation of a catalyst was, the better the activity of this catalyst showed. [(CH3)3NC16H33]4W10O32 exhibited the best catalytic performance and can be recycled for six times without significant decrease in catalytic activity. Using benzothiphene (BT) and 4,6-dimethyldibenzothiphene (4,6-DMDBT) as substrates in model oil, surfactant-type decatungstates also showed high catalytic activity. During desulfurization process, BT conversion can reach 99.6% at 3.25 h, while 99.4% of 4,6-DMDBT conversion reached at 1.25 h, with the temperature of 60 °C under atmospheric pressure. The sulfone can be separated from the oil using N,N-dimethylformamide (DMF) as an extractant, and the sulfur content can be lowered from 1000 to 4 ppm. For real diesel, the sulfur removal can reach 93.5% after five times extraction.  相似文献   

15.
We report on the use of the polyoxometalate acids of the series [PMo(12 − n)VnO40](3 + n)− (n = 0-3) as electrocatalysts in both the anode and the cathode of polymer-electrolyte membrane (PEM) fuel cells. The heteropolyacids were incorporated as catalysts in a commercial gas diffusion electrode based on Vulcan XC-72 carbon which strongly adsorbed a low loading of the catalyst, ca. 0.1 mg/cm2. The moderate activity observed was independent of the number of vanadium atoms in the polyoxometalate. In the anode the electrochemistry is dominated by the V3+/4+ couple. With a platinum reference wire in contact with the anode, polarization curves are obtained withVOC of 650 mV and current densities of 10 mA cm−2 at 100 mV at 80 °C. These catalysts showed an order of magnitude more activity on the cathode after moderate heat treatment than on the anode,VOC = 750 mV, current densities of 140 mA cm−2 at 100 mV. The temperature dependence of the catalysts was also investigated and showed increasing current densities could be achieved on the anode up to 139 °C and the cathode to 100 °C showing the potential for these materials to work at elevated temperatures.  相似文献   

16.
In this study, fresh water methane-producing bacterium (MPB), strain FJ10, which used H2 as an electron donor and CO2 as an electron acceptor, was isolated and chosen as the primary methanogen for the conversion of CO2 into CH4. Improvements to culture medium to increase methane production were investigated using a fractional factorial design in 32 experiments with six variables under consumption of H2/CO2 at ratios of 4 and 1. The tested nutrient compositions were NaCl, NH4Cl, FeSO4, MgCl2, H2PO4 and yeast extract. Experimental results indicate that yeast extract was essential for the growth of strain FJ10 to impact the conversion of CO2 into CH4. Strain FJ10 generated maximum CH4 production with 5.0 g/l of yeast extract. Moreover, optimal culture conditions for methane production by strain FJ10 were 40 °C and pH 8. Approximately 22-25% of CO2 conversion into CH4 was achieved at an H2/CO2 ratio of 4 and roughly 2.5-6% of CO2 conversion into CH4 was obtained at an H2/CO2 ratio of 1 under different pressurized conditions of 1 atm, 50 atm and 100 atm. Under 100 atm, about 6780 μM CH4 was produced with an H2/CO2 ratio of 4 and 4240 μM CH4 was produced with an H2/CO2 ratio of 1 under the steady state condition. The kinetic model for H2/CO2 utilization and CH4 production under different pressures was verified by experimental data. Model predictions are in good agreement with experimental results. The experimental and modeling approaches in this study can be applied to evaluate the conversion of CO2 into CH4 as an energy source by geo-microorganisms in geological sequestration.  相似文献   

17.
Electrochemical behaviour of titanium tetrachloride solutions in 1-butyl-2,3-dimethyl imidazolium azide (BMMImN3) at 65 °C has been examined. Ti(IV) reduction was studied with chronopotentiometry and cyclic voltammetry methods in melts with different concentrations of TiCl4. According to IR spectra, Ti(IV) exists in form of a hexaazidotitanate complex. The electrochemical reduction of this complex was found to proceed irreversibly to Ti(III) species only. Diffusion coefficients of Ti(IV) in this ionic liquid at temperature 65 °C were calculated based on the chronopotentiometry measurements at different TiCl4 concentrations (DTi(IV) = 1.3 ± 0.6 × 10−6 cm2 s−1).  相似文献   

18.
Summary The kinetics of the dibutyltin diacetate (DBTA) – catalyzed polymerization reactions of (η5-C5H4CH2CH2OH)2Mo2(CO)6 with Hypol 2000 (an isocyanate-terminated polyether prepolymer) and with 1,4-butanediol were studied, as were the kinetics of a copolymerization involving (η5-C5H4CH2CH2OH)2Mo2(CO)6 and PEG-1000 (a poly(ethylene glycol)) with Hypol 2000. The purpose was to determine if (η5-C5H4CH2CH2OH)2Mo2(CO)6 appreciably affected the overall rate of the polymerization reaction and if it changed the mechanism of the reaction. The kinetics were analyzed with a fitting program, which allowed extraction of the rate constants for the individual elementary steps in the mechanism. The results showed that (η5-C5H4CH2CH2OH)2Mo2(CO)6 does not significantly alter the timescale of the reaction and that the same reaction mechanism is likely used as with the 1,4-butanediol and PEG-1000. There are some differences in the rate constants of the elementary steps, but these differences can be attributed to the increased steric crowding caused by the bulkier (η5-C5H4CH2CH2OH)2Mo2(CO)6 diol. The effect of the (η5-C5H4CH2CH2OH)2Mo2(CO)6 on the polymers’ physical properties was also investigated. As is the case with other segmented polyurethanes, the hydrogen bonding index (HBI) and the relative amount of soft segments of the (η5-C5H4CH2CH2OH)2Mo2(CO)6-containing polyurethane correlate in a general way with the physical properties of the polymer.  相似文献   

19.
LiNi0.5Co0.5VO4 nano-crystals were solvothermally prepared using a mixture of LiOH·H2O, Ni(NO3)2·6H2O, Co(NO3)2·6H2O and NH4VO3 in isopropanol at 150–200 °C followed by 300–600 °C calcination to form powders. TGA curves of the solvothermal products show weight losses due to evaporation and decomposition processes. The purified products seem to form at 500 °C and above. The products analyzed by XRD, selected area electron diffraction (SAED), energy dispersive X-ray (EDX) and atomic absorption spectrophotometer (AAS) correspond to LiNi0.5Co0.5VO4. V–O stretching vibrations of VO4 tetrahedrons analyzed using FTIR and Raman spectrometer are in the range of 620–900 cm−1. A solvothermal reaction at 150 °C for 10 h followed by calcination at 600 °C for 6 h yields crystals with lattice parameter of 0.8252 ± 0.0008 nm. Transmission electron microscope (TEM) images clearly show that the solvothermal temperatures play a more important role in the size formation than the reaction times.  相似文献   

20.
A perfluorosulfonic acid (PFSA) polymer with pendant side chain -O(CF2)4SO3H was doped with the heteropoly acids (HPAs), H3PW12O40 and H4SiW12O40. Infrared spectroscopy revealed a strong interaction between the HPA and the PFSA ionomer. Modes associated with the peripheral bonds of the HPA were shifted to lower wave numbers when doped into PFSA membranes. Small-angle X-ray scattering (SAXS) measurements showed the presence of large crystallites of HPA in the membrane with d spacings of ca. 10 Å, close to the lattice spacing observed in bulk HPA crystals. Under wet conditions the HPA was more dispersed and constrained the size of the sulfonic acid clusters to 20 Å at a 5 wt% HPA doping level, the same as in the vacuum treated ionomer samples. Under conditions of minimum hydration the HPA decreased the Ea for the self-diffusion of water from 27 to 15 kJ mol−1. The reverse trend was seen under 100% RH conditions. Proton conductivity measurements showed improved proton conductivity of the HPA doped PFSAs at a constant dew point of 80 °C for all temperatures up to 120 °C and at all relative hummidities up to 80%. The activation energy for proton conduction generally was lower than for the undoped materials at RH ≤80%. Significantly the Ea was 1/2 that of the undoped material at RHs of 40 and 60%. A practical proton conductivity of 113 mS cm−1 was observed at 100 °C and 80% RH.  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