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.
In this study, La0.4Sr0.6CoO3‐δ (LSC) oxide was synthesized via an EDTA‐citrate complexing process and its application as a mixed‐conducting ceramic membrane for oxygen separation was systematically investigated. The phase structure of the powder and microstructure of the membrane were characterized by XRD and SEM, respectively. The optimum condition for membrane sintering was developed based on SEM and four‐probe DC electrical conductivity characterizations. The oxygen permeation fluxes at various temperatures and oxygen partial pressure gradients were measured by gas chromatography method. Fundamental equations of oxygen permeation and transport resistance through mixed conducting membrane were developed. The oxygen bulk diffusion coefficient (Dv) and surface exchange coefficient (Kex) for LSC membrane were derived by model regression. The importance of surface exchange kinetics at each side of the membrane on oxygen permeation flux under different oxygen partial pressure gradients and temperatures were quantitatively distinguished from the oxygen bulk diffusion. The maximum oxygen flux achieved based on 1.6‐mm‐thick La0.4Sr0.6CoO3‐δ membrane was ~4.0 × 10?7 mol cm?2 s?1at 950°C. However, calculation results show theoretical oxygen fluxes as high as 2.98 × 10?5 mol cm?2 s?1 through a 5‐μm‐thick LSC membrane with ideal surface modification when operating at 950°C for air separation. © 2009 American Institute of Chemical Engineers AIChE J, 2009  相似文献   

2.
Dense BaCo0.7Fe0.2Ta0.1O3?δ (BCFT) perovskite membranes were successfully synthesized by a simple solid state reaction. In situ high‐temperature X‐ray diffraction indicated the good structure stability and phase reversibility of BCFT at high temperatures. The thermal expansion coefficient (TEC) of BCFT was determined to amount 1.02 × 10?5 K?1, which is smaller than those of Ba0.5Sr0.5Co0.8Fe0.2O3?δ (BSCF) (1.15 × 10?5 K?1), SrCo0.8Fe0.2O3?δ (SCF) (1.79 × 10?5 K?1), and BaCo0.4Fe0.4Zr0.2O3?δ (BCFZ) (1.03 × 10?5 K?1). It can be seen that the introduction of Ta ions into the perovskite framework could effectively lower the TEC. Thickness dependence studies of oxygen permeation through the BCFT membrane indicated that the oxygen permeation process was controlled by bulk diffusion. A membrane reactor made from BCFT was successfully operated for the partial oxidation of methane to syngas at 900°C for 400 h without failure and with the relatively high, stable oxygen permeation flux of about 16.8 ml/min cm2. © 2009 American Institute of Chemical Engineers AIChE J, 2010  相似文献   

3.
《Ceramics International》2022,48(1):415-426
The oxygen transport membrane (OTM) has huge application prospects in gas separation and carbon neutralization based on oxygen enriched combustion. In this paper, the family 60 wt.%Ce0.9Pr0.1O2-δ-40 wt.%Pr0.6Sr0.4Fe1-xInxO3-δ (CPO-PSF1-xIxO, x = 0.01, 0.025, 0.05, 0.075, 0.1) cobalt-free dual-phase MIEC OTMs doped with indium have been successfully prepared by Pechini method. The phase structure, surface morphology, element distribution, oxygen permeability, and long-term operation stability of these OTMs are systematically explored. Among these OTMs, the champion oxygen permeable flux of CPO-PSF0.99I0.01O reaches 1.07 mL min?1·cm?2 and 0.80 mL min?1·cm?2 at 1000 °C under air/He gradient and air/CO2 gradient. Meanwhile, CPO-PSF0.99I0.01O maintains the value of 0.80 mL min?1·cm?2 steadily at 1000 °C for 100 h when pure CO2 as the sweep gas. The surface element distribution and phase structure of the OTMs after long-term oxygen permeability reaction are investigated by XRD, SEM combining with EDS, where the spent membranes retain the same structure and component as the fresh membranes, demonstrating that the In-doped OTMs have an excellent CO2 tolerance. Suitable indium substitution for iron of these OTMs not only improves the oxygen permeability, but also maintains the long-term reaction stability of the material.  相似文献   

4.
We have prepared a novel oxygen carrier composed of Starburst dendrimer 3 hematoporphyrin iron(III) di‐1‐(3‐aminopropyl)imidazole di‐succinic acid (SDHAPI). This compound has a molecular weight of ~12.7 kD. O2 and CO binding to SDHAPI was demonstrated by UV‐vis spectroscopy. Oxygenated SDHAPI had an estimated auto‐oxidation half‐life of 24 min. The oxygen stoichiometry (~6 mol of O2 mol?1 SDHAPI) and the kobs(on+off) of 0.084 s?1 and kobs(off) of 0.009 s?1 were estimated with an oxygen electrode. Copyright © 2005 Society of Chemical Industry  相似文献   

5.
M.A. Haleem  D. Billaud  A. Pron 《Polymer》1982,23(10):1409-1411
The kinetics of doping and degradation of all trans-polyacetylene by oxygen, have been studied over the temperature range 70°–110°C. Doping and degradation have first-order kinetics with the observed Arrhenius parameters E = 11.52, E = 13.45 kcal mol?1 and log A (s?1 = 4.19, log A (s?1) = 4.49 respectively. The observations support the view that the reactions proceed via an intermediate-complex mechanism. Initially oxygen reduced the resistivity of (CH)x by formation of a charge transfer complex and subsequently reacted with (CH)x with an increase in resistivity.  相似文献   

6.
SrCo0.9Sc0.1O3 (SCSc) perovskite powders with sub-micron particle size were synthesized by a modified Pechini method combined with a post-treatment of sintering and ball-milling. From the prepared powders, the SCSc hollow fibre membranes with asymmetric structure and gas-tight property were fabricated by spinning a polymer solution containing 58.4 wt% SCSc followed by sintering at 1200 °C for 5 h. The oxygen permeation properties of the obtained SCSc fibres were measured under air/He gradients at 500–800 °C. This showed the oxygen flux of 1 mL cm?2 min?1 at 750 °C and 4.41 mL cm?2 min?1 at 900 °C. Modeling analysis reveals that the oxygen permeation process is predominated by oxygen surface exchange kinetics with an activation energy of 95.0 kJ mol?1. The SCSc membranes showed excellent oxygen permeation performance while exhibiting high structural and permeating stability at intermediate temperatures (500–800 °C).  相似文献   

7.
BACKGROUND: Xylitol is a sugar alcohol (polyalcohol) with many interesting properties for pharmaceutical and food products. It is currently produced by a chemical process, which has some disadvantages such as high energy requirement. Therefore microbiological production of xylitol has been studied as an alternative, but its viability is dependent on optimisation of the fermentation variables. Among these, aeration is fundamental, because xylitol is produced only under adequate oxygen availability. In most experiments with xylitol‐producing yeasts, low oxygen transfer volumetric coefficient (KLa) values are used to maintain microaerobic conditions. However, in the present study the use of relatively high KLa values resulted in high xylitol production. The effect of aeration was also evaluated via the profiles of xylose reductase (XR) and xylitol dehydrogenase (XD) activities during the experiments. RESULTS: The highest XR specific activity (1.45 ± 0.21 U mgprotein?1) was achieved during the experiment with the lowest KLa value (12 h?1), while the highest XD specific activity (0.19 ± 0.03 U mgprotein?1) was observed with a KLa value of 25 h?1. Xylitol production was enhanced when KLa was increased from 12 to 50 h?1, which resulted in the best condition observed, corresponding to a xylitol volumetric productivity of 1.50 ± 0.08 gxylitol L?1 h?1 and an efficiency of 71 ± 6.0%. CONCLUSION: The results showed that the enzyme activities during xylitol bioproduction depend greatly on the initial KLa value (oxygen availability). This finding supplies important information for further studies in molecular biology and genetic engineering aimed at improving xylitol bioproduction. Copyright © 2008 Society of Chemical Industry  相似文献   

8.
Oxygen reduction electrocatalysts based on the monoethanolmine complexes {[CoEtm]2(μ-Etm)4Ni(NO3)2} and {[CoEtm]2(μ-Etm)4Ni(NO3)2} + activated carbon AG-3 have been obtained by high-temperature synthesis. The nature of active centers on the synthesized electrocatalysts was described. Using potentiostatic and cyclic potentiodynamic voltammetry, the kinetic characteristics of catalysts in the oxygen electroreduction reaction were determined. Thermal decomposition of the thermally unstable complexes was described and character of the active centers formed was discussed. The optimal synthesis temperature of electrocatalysts is 600 °C in an inert atmosphere. The calculated exchange current densities for the oxygen electroreduction reaction at the catalysts in 1 M KOH at 20 °C was j 0  = 1.01 × 10?3 A g?1–3.3 × 10?3 A g?1. The Tafel slopes of stationary polarization curves are 0.054–0.063 V for b 1 and 0.106–0.125 V for b 2 . The prepared electrocatalysts can be recommended only for electrochemical systems with alkaline electrolyte.  相似文献   

9.
In a study of the enlargement of pores of coals it has been found that treatment of a bituminous coal (PSOC No. 371, from the Pennsylvania State University Coal Section) with a 5:95 O2:N2 stream 4 h at 400 °C increases the surface area as measured by nitrogen adsorption at 77K by a factor of at least 50 to a value 52 m2 g?1. The increase in pore size was accompanied by a 9.7% weight loss. Simultaneously, the area as measured by carbon dioxide at 195K increased from 61 to 136 m2 g?1 and that measured by carbon dioxide at room temperature increased from 125 to 237 m2 g?1. Attempts to enlarge the pores by oxidation with hydrogen peroxide or ozone were unsuccessful. A Pittsburgh coal subject to a small percentage of oxygen in nitrogen or steam at 300 to 400 °C showed a surface area as measured by nitrogen adsorption of less than 1 m2 g?1 both before and after such pretreatment. This same coal with a 5:95 O2:N2 stream for 4 h at 450 °C showed a surface area of 110 m2 g?1 measured by nitrogen adsorption at 77K.  相似文献   

10.
Glucose oxidase was immobilized onto poly(2-hydroxyethyl methacrylate) (pHEMA) membranes by two methods: by covalent bonding through epichlorohydrin and by entrapment between pHEMA membranes. The highest immobilization efficiency was found to be 17.4% and 93.7% for the covalent bonding and entrapment, respectively. The Km values were 5.9 mmol dm?3, 8.8 mmol dm?3 and 12.4 mmol dm?3 for free, bound and entrapped enzyme, respectively. The Vmax values were 0.071 mmol dm?3 min?1, 0.067 mmol dm?3 min?1 and 0.056 mmol dm?3 min?1 for free, bound and entrapped enzyme. When the medium was saturated with oxygen, Km was not significantly altered but Vmax was. The optimum pH values for the free, covalently-bound and entrapped enzyme were determined to be 5, 6, and 7, respectively. The optimum temperature was 30°C for free or covalently-bound enzyme but 35°C for entrapped enzyme. The deactivation constant for bound enzyme was determined as 1.7 × 10?4 min?1 and 6.9 × 10?4 min?1 for the entrapped enzyme.  相似文献   

11.
Facilitated transport of oxygen was performed through chelate membranes containing cobalt with selective oxygen binding ability as a fixed oxygen carrier. Chelate membranes were obtained from Schiff base membranes after treating a poly(allyl amine) (PAAm) and poly(vinyl alcohol) (PVA) blend with salicylaldehyde. It is confirmed that the O? O stretching peak through a frequency change in FTIR could be seen at 1150 cm?1 between cobalt in the membrane and incoming oxygen. The permeability of oxygen through Schiff base membranes was 2.01?2.98 × 10?13 [cm3 (STP) cm2/cm s cmHg] and oxygen permselectivity was in the range of 1.83?3.27. For chelate membranes, both the permeability of oxygen and oxygen selectivity increased to 2.15?2.82 × 10?12 [cm3 (STP) cm2/cm s cmHg] and around 8, respectively. Permselectivity of chelate increased as a result of facilitation of O2 and inhibition of N2 transport. Detailed results and the mechanism of facilitation of oxygen are discussed on the basis of molecular interactions. © 1995 John Wiley & Sons, Inc.  相似文献   

12.
Diffusion coefficients for oxygen in 0·1 M sodium hydroxide solution have been determined from 0° to 65°C using a rotating platinum electrode. The results may be represented with a standard deviation of approximately 1% by the expression:
where Do = 8·03 × 10?3cm2s?1 and the apparent activation energy, QD, is 3·49 kcal mol?1. Between 25° and 65°C the product of the diffusivity and the solution viscosity is essentially constant:
Values for diffusion coefficients for oxygen in pure water have been derived from these data with an absolute error probably less than 4%.Difficulties in obtaining reproducible results at a rotating platinum electrode are attributed to deactivation of the electrode by reduction of a surface oxygen phase. Results at the higher temperatures indicate that methods of determining diffusivities by diffusion through a stagnant layer of solution involve an increasing indeterminate error as the temperature rises, due to convective mass transport.  相似文献   

13.
The ionic conduction in sintered Bi2O2-Y2O3 was investigated by measuring the conductivity and the emf of an oxygen concentration cell using the specimen tablet as electrolyte. The face centred cubic phase in this system was found to show high oxide ion conduction accompanied by a little electronic conduction when exposed to air. This phase was stable with a composition of 25 ~ 43 mol % Y2O3 over a wide range of temperatures, and the oxide ion conductivity increased with decrease in Y2O3. The conductivities of (Bi2O3)0.75 (Y2O3)0.25 were 1.6×10?1 Ω?1 cm?1 at 700°C and 1.2×10?2 Ω?1 cm?1 at 500°C values which are many times higher than those of stabilized zirconia (ZrO2)0.90(Y2O3)0.10 at corresponding temperatures. Specimens containing less than 25 mol % Y2O3 showed a phase transition at 700 ~ 580°C and the conductivities decreased remarkably below these temperatures. High oxide ion conduction in the fcc phase is attributed to the migration of oxide ion vacancies which were present in an appreciable amount.  相似文献   

14.
In this study, the crystal structure, thermal, oxygen transport, electrical conductivity and electrochemical properties of the perovskite NdBa0.5Sr0.5Co2O5+δ (NBSC55) are investigated. In the temperature range of 250 °C–350 °C, the weight loss upon heating was due to a partial loss of lattice oxygen and along with a reduction of Co4+ to Co3+. The tend of weight-loss slows down as temperature increased above 350 °C indicating a reduction of Co3+ to Co2+ during this stage. The oxygen migration is dominated by surface exchange process at high temperature range (650-800 °C); however, the bulk diffusion process prevails at low temperature range (500–600 °C). For long-term testing, the polarization resistance of NBSC55 increases gradually form 3.13 Ω cm2 for 2 h to 3.34 Ω cm2 for 96 h at 600 °C and an increasing-rate for polarization resistance is around 0.22% h?1. The power density of the single cell with NBSC55 cathode reached 341 mW cm?2 at 800 °C.  相似文献   

15.
Wet oxidation (WO) pretreatment of sugarcane bagasse, rice hulls, cassava stalks and peanut shells was investigated. WO was performed at 195 °C for 10 min, with 2 g kg?1 of Na2CO3 and under either 3 or 12 bar of oxygen. Oxygen pressure and the type of raw material used had a major effect on the fractionation of the materials, formation of sugars and by‐products, and cellulose enzymatic convertibility. Cellulose content in the solid fraction increased after pretreatment of all materials, except rice hulls. The greatest increase, from 361 g kg?1 to almost 600 g kg?1, occurred for bagasse. The solubilisation of individual components was different for each material. Bagasse xylan was solubilised to a large extent, and 45.2% of it was recovered as xylose and xylo‐oligosaccharides in the liquid fraction. In the prehydrolysates of rice hulls around 40% of the original glucan was recovered as gluco‐oligosaccharides, due to hydrolysis of starch contained in grain remains. The formation of by‐products was modest for all the materials, but increased with increasing oxygen pressure. The highest yield of acetic acid (34–36 g kg?1 of raw material) and furfural (0.7–1.8 g kg?1) occurred for bagasse. The pretreatment enhanced the enzymatic convertibility of cellulose giving the best result (670.2 g kg?1) for bagasse pretreated at the highest oxygen pressure. However, for the other materials the pretreatment conditions were not effective in achieving cellulose conversions above 450 g kg?1. Some enzymatic conversion of xylan was observed. Copyright © 2007 Society of Chemical Industry  相似文献   

16.
Ammonium fumarate production from glucose‐based media by Rhizopus arrhizus NRRL 1526 with mycelial growth controlled by phosphorus limitation exhibited mixed‐growth‐associated product formation kinetics, with growth‐associated production related to secondary mycelial growth only. The contribution of the primary mycelial growth phase was minimised by resorting to prolonged batch production using free mycelia under intermittent glucose feeding or repeated batch production using immobilised mycelia. The metabolic activity of free or immobilised mycelia was limited by fumarate accumulation or by oxygen diffusion phenomena, respectively. For batch cultures in a 15 dm3 stirred bioreactor the peripheral impeller speed (vI) was increased from 1.88 to 3.3 m s?1, and the fumarate yield coefficient on glucose increased from 0.25 ± 0.01 to 0.42 ± 0.02 g g?1, while the malate yield coefficient on fumarate (YM/F) reduced from 0.46 ± 0.01 to 0.14 ± 0.01 g g?1. With a net increase in the fumarate‐to‐malate ratio from 2 to 6.5, a vI value of 3.3 m s?1 gave the best fermentation performance and provided a basis for further scale‐up studies. © 2002 Society of Chemical Industry  相似文献   

17.
In this review article we want to give information about low molecular and polymer organic semiconductors, which were recently synthesized in our institute. Specific electric conductivities up to σ298°K = 9.0 · 10?5Ω?1 · cm?1 and thermic activation energies of E = 0.30 eV of polyenearylenes, respectively -heteroarylenes were measured. Polyazomethines have a maximum σ298°K = 3.3 · 10?9Ω?1 · cm?1 and E = 0.35 eV. Polymers with indophenine units have conductivities up to σ298°K = 1.1 · 10?4Ω?1 · cm?1 and E = 0.39 eV. A maximum of σ298°K = 5.0 · 10?2Ω?1 · cm?1 and E = 0.05 eV was found for bis-(1.2-dicyanoethylenedithiolo)-metal salts. Polymers with a phthalocyanine- or hemiporphyrazine-like structure achieve a conductivity of σ298°K = 2.3 · 10?2Ω?1 · cm?1 and E = 0.15eV. Coordination polymers of dimercaptomaleic acid, respectively their monoamide show a maximum of σ298°K = 3.2 · l0?lΩ?l · cm?1 and E = 0.20 ev. Polymers with σ298°K ≤1.5 · 10?5 Ω?l · cm?l and E ≥ 0.5 eV were obtained by the polymerization of succinonitrile. All the investigated substances show an electronic conductivity. The existence of an ionic conductivity could, in all cases, be excluded by using direct current measurements over a long period of time.  相似文献   

18.
The ozone demand to oxidize HS?/H2S [pKa(H2S) = 6.9, k(HS? + O3) = 3 × 109 M?1 s?1, k(H2S + O3) = 3 × 104 M?1 s?1] to SO4 2? is only 2.4 mol ozone per mol SO4 2? formed, much lower than stoichiometric 4.0 mol/mol if a series of O-transfer reactions would occur. As primary step, the formation of an ozone adduct to HS?, HSOOO, is suggested that decomposes into HSO and singlet oxygen (16%) or rearranges into peroxysulfinate ion, HS(O)OO (84%). Potential reactions of the above intermediates are discussed. Some of these can account for the low ozone demand.  相似文献   

19.
Cu1.5Mg0.5V2O7 was prepared by a solid state method. Its phase, microstructure, thermal expansion property, and Raman spectra were analyzed in detail. Results show that Cu1.5Mg0.5V2O7 maintains a monoclinic crystal structure and exhibits an excellent linear negative thermal-expansion property with coefficient of thermal expansion of ?8.72?×?10?6?K?1 over a wide temperature range of 153–673?K. The mechanism underlying the negative thermal expansion of Cu1.5Mg0.5V2O7 involves the coupling effect of the tetrahedron caused by the lateral vibration of the bridge oxygen atom and the tensile effect of the tetrahedron, The partial collapse caused by the loss of the oxygen atoms also plays an important role in the mechanism.  相似文献   

20.
SiOx films were deposited from a mixture of tetramethoxysilane (TMOS) and oxygen on poly(ethylene 2,6‐naphthalate) film using ion‐assisted plasma polymerization technique (Method II) and conventional plasma polymerization technique (Method I), and were compared in chemical composition and gas barrier properties. Methods I and II were different in electrical circuit between electrodes (anode and cathode) and electric power supply. In Method I, the anode electrode was grounded, and the cathode electrode was coupled to the discharge power supply. In Method II, the anode electrode was connected with the discharge power supply, and the cathode electrode was grounded. There was not large difference in SiOx deposition rate between the plasma polymerizations by Methods I and II. Plasma polymers deposited from TMOS/O2 mixtures by Method II possessed smaller C/Si and O/Si atomic ratios than those deposited by Method I and showed advantage in gas barrier properties. The oxygen and water vapor permeation rates were 0.08–0.13 cm3 m?2 day?1 atm?1 at 30°C at 90% RH and 0.244–0.276 g m?2 day?1 at 40°C at 90% RH, respectively. From these results, it can be concluded that the ion‐assisted plasma polymerization is a useful technique for deposition of gas barrier SiOx thin films. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 104: 915–925, 2007  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