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 62 毫秒
1.
The steam reforming of phenol towards H2 production was studied in the 650–800 °C range over a natural pre-calcined (air, 850 °C) calcite material. The effects of reaction temperature, water, hydrogen, and carbon dioxide feed concentrations, and gas hourly space velocity (GHSV, h−1) were investigated. The increase of reaction temperature in the 650–800 °C range and water feed concentration in the 40–50 vol% range were found to be beneficial for catalyst activity and H2-yield. A similar result was also obtained in the case of decreasing the GHSV from 85,000 to 30,000 h−1. The effect of concentration of carbon dioxide and hydrogen in the phenol/water feed stream was found to significantly decrease the rate of phenol steam reforming reaction. The latter was probed to be related to the reduction in the rate of water dissociation as evidenced by the significant decrease in the concentration of adsorbed bicarbonate and OH species on the surface of CaO according to in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS)-CO2 adsorption experiments in the presence of water and hydrogen in the feed stream. Details of the CO2 adsorption on the CaO surface at different reaction temperatures and gas atmospheres using in situ DRIFTS and transient isothermal adsorption experiments with mass spectrometry were obtained. Bridged, bicarbonate and unidentate carbonate species were formed under CO2/H2O/He gas mixtures at 600 °C with the latter being the most populated. A substantial decrease in the surface concentration of bicarbonate and OH species was observed when the CaO surface was exposed to CO2/H2O/H2/He gas mixtures at 600 °C, result that probes for the inhibiting effect of H2 on the phenol steam reforming activity. Phenol steam reforming reaction followed by isothermal oxygen titration allowed the measurement of accumulated “carbonaceous” species formed during phenol steam reforming as a function of reaction temperature and short time on stream. An increase in the amount of “carbonaceous” species with reaction time (650–800 °C range) was evidenced, in particular at 800 °C (4.7 vs. 6.7 mg C/g solid after 5 and 20 min on stream, respectively).  相似文献   

2.
Jerzy Zawadzki 《Carbon》1978,16(6):491-497
Spectral studies of oxidation process were carried out on carbon films prepared by carbonization of polyfurfuryl alcohol and of cellulose. Investigation of the oxygen role in the initial stages of carbonization shows that oxygen surface compounds formed during such process have different chemical structure from those produced during oxidation of carbonic film previously desorbed at 600°C. Chemisorption of oxygen at room temperature on the carbonic film (desorbed previously at 600°C) occurs with the participation of π electrons of condensed aromatic systems. Iono-radical structures formed during that process show absorption bands in the range of 1590cm−1. Oxidation of the carbon films carbonized at 800°C causes a decrease in the absorption coefficient, and oxygen surface compounds formed show absorption bands at about 1760, 1600 and 1260cm−1. The carbon film as a model substance gives possibilities for broader application of IR spectroscopy in studies of carbon, carbonization and activation processes, sorption effects and catalytic mechanisms.  相似文献   

3.
The characteristics of hydrogen production using immobilized cyanobacterium Microcystis aeruginosa were studied through a two-stage cyclic process. Immobilized cells were first grown photosynthetically under CO2 and light, followed by anaerobic H2 production in the absence of light and sulfur. M. aeruginosa was capable of generating H2 under immobilized conditions, and the use of immobilized cells allowed the maintenance of stable production and sped up the changes in culture conditions for cyclic two-stage operation. M. aeruginosa was also capable of utilizing exogenous glucose as a substrate to generate hydrogen and 30 mM concentration proved to be optimal. The externally added glucose improved H2 production rates, total produced volume and the lag time required for cell adaptation prior to H2 evolution. The rate of hydrogen evolution was increased as temperature increased, and the maximum evolution rate was 48 mL/h/L and 34.0 mL/h/L at 42 °C and 37 °C, respectively. The optimal temperature for hydrogen production was 37–40 °C because temperatures higher than 42 °C resulted in cell death. In order to continue repeated cycles of H2 production, at least two days of photosynthesis under conditions with light, CO2, and sulfur should be allowed for cells to recover H2 production potential and cell viability.  相似文献   

4.
The reduction of NO to N2/N2O in the presence of excess O2 has been successfully achieved at 70 °C using an electrochemical cell of the type, 0.1% NO, 0–10% O2, Pt | NAFION | Pt, H2O. An H+-conducting solid polymer electrolyte (SPE) plays a key role in evolving hydrogen on the Pt cathode, where the catalytic NO–H2 takes place. It was revealed that the competitive H2–O2 reaction is suppressed because the Pt surface was covered with stable nitrate (NO3) species, which blocks oxygen adsorption hereon. The inhibition of H2–O2 reaction becomes most efficient at 100 °C in agreement with the optimal operation temperature range of SPE. The reduction efficiency of NO in an excess O2 could be improved by packing 1 wt% Pt/ZSM-5 catalyst in the cathode room. The combination between the SPE cell and Pt catalysts can broadly be applied to novel low-temperature deNOx processes in a strongly oxidizing atmosphere.  相似文献   

5.
The structural change of diamond induced by hydrogen plasma exposure at room temperature, and its thermal stability, were investigated using electron spin resonance (ESR), X-ray photoelectron spectroscopy (XPS) and Fourier transform infrared spectroscopy (FTIR) techniques. ESR observation revealed that it gave rise to a highly defective structure (spin concentration of 1020 cm− 3), which is very similar to the structure of hydrogenated amorphous carbon (a-C : H) confirmed by XPS and FTIR. Post-annealing was also carried out to clarify the thermal stability of the defects. The number of spin centers decreased with increasing annealing temperature, and eventually, the defective structure changed to a graphitic one by annealing at 800 °C.  相似文献   

6.
Dense membrane with the composition of SrFe0.6Cu0.3Ti0.1O3-δ (SFCTO) was prepared by solid state reaction method. Oxygen permeation flux through this membrane was investigated at operating temperature ranging from 750℃ to 950℃ and different oxygen partial pressure. XRD measurements indicated that the compound was able to form single-phased perovskite structure in which part of Fe was replaced by Cu and Ti. The oxygen desorption and the reducibility of SFCTO powder were characterized by thermogravimetric analysis and temperature programmed reduction analysis, respectively. It was found that SFCTO had good structure stability under low oxygen pressure at high temperature. The addition of Ti increased the reduction temperature of Cu and Fe. Performance tests showed that the oxygen permeation flux through a 1.5 mm thick SFCTO membrane was 0.35-0.96 ml·min ^-1·cm^-2 under air/helium oxygen partial pressure gradient with activation energy of 53.2 kJ·mol^-1. The methane conversion of 85%, CO selectivity of 90% and comparatively higher oxygen permeation flux of 5 ml·min^-1·cm^- 2 were achieved at 850℃, when a SFCTO membrane reactor loaded with Ni-Ce/Al2O3 catalyst was applied for the partial oxidation of methane to syngas.  相似文献   

7.
The development of a nickel composite membrane with acceptable hydrogen permselectivity at high temperature in a membrane reactor for the highly endothermic dry reforming of methane reaction was the purpose of this work. A thin, catalytically inactive nickel layer, deposited by electroless plating on asymmetric porous alumina, behaved simply as a selective hydrogen extractor, shifting the equilibrium in the direction of a higher hydrogen production and methane conversion. The main advantage of such a nickel/ceramic membrane reactor is the elimination or limitation of the side reverse water gas shift reaction. For a Ni/Al2O3 catalyst, containing free Ni particles, normally sensitive to coking, the use of the membrane reactor allowed an important reduction of carbon deposition (nanotubes) due to restriction of the Boudouard reaction. For a Ni–Co/Al2O3 catalyst, where the metallic nickel phase was stabilized by the alumina, the selective removal of the hydrogen significantly enhanced both methane conversion (+67% at 450 °C, +22% at 500 °C and +18% at 550 °C) and hydrogen production (+42% at 450 °C, +32% at 500 °C and +22% at 550 °C) compared to the results obtained for a packed-bed reactor. The hydrogen selectivity during the catalytic tests at 550 °C, maintained with constant separation factors (7 for H2/CH4, 8 for H2/CO and 10 for H2/CO2), higher than Knudsen values, attested to the high thermal stability of the nickel composite membrane.  相似文献   

8.
Mesoporous silicas (MesoPSs) were hydrothermally synthesized from calcined and selectively acid-leached kaolinites with a range of crystallinity, using cetyltrimethyl ammonium bromide (CTABr), to investigate the effect of the kaolinite crystallinity on the porous properties of the resulting MesoPSs. Four kaolinites were used, with Hinckley indices ranging from 0.51 to 1.20 and (001) crystallite sizes ranging from 20 to 37 nm. After calcination at 600 °C for 24 h they were selectively leached with 2.5 M H2SO4 at 90 °C for 2 h to prepare microporous silica (MicroPSs). The Si/Al ratios of these MicroPSs varied from 21 to 82 and their specific surface areas (SBET) ranged from 169 to 370 m2/g, these parameters tending to increase with decreasing Hinckley index of the kaolinite. MesoPSs were synthesized by reacting the resulting MicroPSs with CTABr in NaOH solution under hydrothermal conditions. The MicroPS was mixed with CTABr, NaOH and water in the molar ratio (MicroPS):CTABr:NaOH:H2O = 1:0.1:0.3:150. The synthesis was carried out by stirring the suspension at room temperature for 24 h, aging for 24 h, hydrothermal treatment at 110 °C for 24 h and calcination at 560 °C for 6 h to remove the surfactants. The SBET values of the resulting MesoPSs ranged from 932 to 1240 m2/g, correlating with the SBET values of the precursor MicroPS and the crystallinity of the kaolinite starting materials.  相似文献   

9.
Mesoporous silicas with vesicular and onion-like morphologies were assembled through hydrogen-bonding pathway from sodium silicate as silica source and electrically neutral α,ω-diamine, Jeffamine D2000 surfactant (H2NCH(CH3)CH2[OCH2CH(CH3)]33NH2) as template in aqueous media at different synthesis temperatures (25, 60 and 100 °C). Assembling the material at 100 °C afforded onion-like core shell mesoporous silica, while at relatively lower temperature, e.g. 25 and 60 °C, multilamellar vesicles were obtained. Mesoporous silica with onion-like morphology was also obtained by a two-step synthesis involving an aging period of 20 h at room temperature followed by a hydrothermal stage (1–12 h) at 100 °C. The heavily cross-linked (Q4/Q3 ratio of 4.43) onion-like mesophase silica exhibited high hydrothermal stability. The BET surface area, pore volume and KJS (Kruk-Jaroniec-Sayari) pore diameter of the onion-like mesoporous silica were found to be 464 m2 g−1, 1.16 m3 g−1 and 7.2 nm, respectively.  相似文献   

10.
Thermal behaviour of synthetic Cu–Mg–Mn and Ni–Mg–Mn layered double hydroxides (LDHs) with MII/Mg/Mn molar ratio of 1:1:1 was studied in the temperature range 200–1100 °C by thermal analysis (TG/DTA/EGA), powder X-ray diffraction (XRD), Raman spectroscopy, and voltammetry of microparticles. Powder XRD patterns of prepared LDHs showed characteristic hydrotalcite-like phases, but further phases were indirectly found as admixtures. The Cu–Mg–Mn precipitate was decomposed at temperatures up to ca. 200 °C to form an XRD-amorphous mixture of oxides. The crystallization of CuO (tenorite) and a spinel type mixed oxide of varying composition CuxMgyMnzO4 with Mn4+ was detected at 300–500 °C. At high temperatures (900–1000 °C), tenorite disappeared and a consecutive crystallization of 2CuO·MgO (gueggonite) was observed. The high-temperature transformation of oxide phases led to a formation of CuI oxides accompanied by oxygen evolution. The DTA curve of Ni–Mg–Mn sample exhibited two endothermic effects characteristic for hydrotalcite-like compounds. The first one with minimum at 190 °C can be ascribed to a loss of interlayer water, the second one with minimum at 305 °C to the sample decomposition. Heating of the Ni–Mg–Mn sample at 300 °C led to the onset of crystallization of oxide phases identified as NixMgyMnzO4 spinel, (Ni,Mg)O oxide containing Mn4+ cations, and easily reducible XRD-amorphous species, probably free MnIII,IV oxides. At 600 °C (Raman spectroscopy) and 700 °C (XRD), the (Ni,Mg)6MnO8 oxide with murdochite structure together with spinel phase were detected. Only spinel and (Ni,Mg)O were found after heating at 900 °C and higher temperatures. Temperature-programmed reduction (TPR) profiles of calcined Cu–Mg–Mn samples exhibited a single reduction peak with maximum around 250 °C. The highest H2 consumption was observed for the sample calcined at 800 °C. The reduction of Ni–Mg–Mn samples proceeded by a more complex way and the TPR profiles reflected the phase composition changing depending on the calcination temperature.  相似文献   

11.
Fresh and stored coals from the United States (New Mexico, Colorado and Wyoming) and Czech Republic (North Bohemian and Upper Silesian Basins) were studied by the method of temperature programmed desorption. Desorbed gases were analyzed for their chemical and carbon isotope composition. Upon heating from room temperature with a constant rate of 40 °C/min, two desorption phases were observed: low temperature desorption of CH4 and CO2 with a maximum intensity between 50 and 80 °C and high temperature desorption of CO2 only between 150 and 210 °C. The desorption of (residual) primary coalbed gas was compared with the desorption of re-adsorbed gases. The δ13C values of the desorbed gases changed due to isotopic fractionation during coal degassing. Kinetic isotope effects were evaluated by comparing the gas desorption from fresh and stored coals from the same seams. Mean values of isotope enrichment during desorption are 2‰ and 1.9‰ for CO2 and CH4, respectively.  相似文献   

12.
Oxygen storage capacity (OSC) of CeO2–ZrO2 solid solution, CexZr(1−x)O4, is one of the most contributing factors to control the performance of an automotive catalyst. To improve the OSC, heat treatments were employed on a nanoscaled composite of Al2O3 and CeZrO4 (ACZ). Reductive treatments from 700 to 1000 °C significantly improved the complete oxygen storage capacity (OSC-c) of ACZ. In particular, the OSC-c measured at 300 °C reached the theoretical maximum with a sufficient specific surface area (SSA) (35 m2/g) after reductive treatment at 1000 °C. The introduced Al2O3 facilitated the regular rearrangement of Ce and Zr ions in CeZrO4 as well as helped in maintaining the sufficient SSA. Reductive treatments also enhanced the oxygen release rate (OSC-r); however, the OSC-r variation against the evaluation temperature and the reduction temperature differed from that of OSC-c. OSC-r measured below 200 °C reached its maximum against the reduction temperature at 800 °C, while those evaluated at 300 °C increased with the reduction temperature in the same manner as OSC-c.  相似文献   

13.
Turkish sepiolite–zirconium oxide mixtures were applied as a support for the silver catalyst in a soot combustion. Sepiolite–Zr–K–Ag–O catalyst was characterized by XRD, N2 adsorption, SEM, TPR-H2 and EGA-MS. The combustion of soot was studied with a thermobalance (TG-DTA). The modification resulted in a partial degradation of the sepiolite structure, however, the morphology was preserved. The adsorption of N2 of the modified sepiolite is a characteristic for mesoporous materials with a wide distribution of pores. The specific surface area SBET equals 83 m2/g and the pores volume is 0.23 cm3/g. The basic character of the surface centers of sepiolite is indicated by CO2 desorption (TPD-MS) at 170 °C and at about 620 °C due to a surface carbonates decomposition. The thermodesorption of oxygen at 650–850 °C indicates the decomposition of AgOx phases at the surface. The presence of AgOx phases is also confirmed by TPR-H2 spectrum (low temperature reduction peak at 130 and 180 °C). The high-temperature reduction at about 570 °C is probably related to Ag–O–M phases on the support.The soot combustion takes place at T50 = 575 °C. Without silver (sepiolite–Zr–K–O) T50 = 560 °C but sepiolite modified with silver (sepiolite–Zr–K–Ag–O) undergoes the same process at T50 = 490 °C.  相似文献   

14.
In this study sulfur pyrolysis behavior of two Chinese high sulfur coals and their treated coal samples was investigated by Py-MS at a heating rate of 5 °C/min from room temperature to 1025 °C under hydrogen, helium and 2% O2-He. It is found that the internal and external hydrogen do not show hydrogenation ability at temperature below 400 °C, due to no H2S formation at this temperature region for all the coal samples. At temperature higher than 400 °C, not only the indigenous hydrogen but also indigenous oxygen can react with sulfur-containing radicals to form H2S or SO2. The evolution of H2S and SO2 displays the same profiles in pyrolysis of ZY pyrite-free coal under He, further revealing that after the breakage of C-S bond in the organic sulfur structure in coal to form sulfur-containing radicals, which can equally react with indigenous hydrogen and oxygen. The similar tendency between evolution of CO2 and SO2 and the same ending temperature also shows that not only C-S but also C-C bond can be broken in pyrolysis of ZY coals under 2% O2-He atmosphere. However, unlike SO2 evolution, CO2 emission increases in the temperature ranging from 500 °C to 800 °C in LZ raw and deashed coals, implying the breakage of C-C bond at high temperature, which might be related to their low coal rank and high pyrite content.  相似文献   

15.
A. Dandekar  R.T.K. Baker  M.A. Vannice 《Carbon》1998,36(12):1821-1831
A high surface area activated carbon, graphitized carbon fibers and synthetic diamond powder were characterized by X-ray diffraction, temperature-programmed desorption and diffuse reflectance infrared (IR) spectroscopy (DRIFTS). The activated carbon was analyzed as received as well as after either a nitric acid treatment to introduce oxygen functional groups on its surface or a high temperature treatment (HTT) in H2 at 1223 K to remove surface groups. TPD evolution profiles of CO and CO2 were combined with DRIFTS spectra of these carbon surfaces before and after pretreatments in H2 at 723 and 1223 K to provide complementary information regarding the nature of these surface groups. Significant amounts of both low- and high-temperature CO2 desorption occurred from the HNO3-treated carbon, indicating that both strongly and weakly acidic groups were introduced on this surface and, in addition, comparable amounts of CO and CO2 were desorbed. With the graphitized carbon fibers and diamond powder, larger amounts of CO were desorbed compared to CO2, indicating the presence of predominantly weakly acidic or non-acidic groups on these surfaces. For the HNO3-treated carbon, IR peaks associated with surface carboxylic acid groups initially present disappeared after treatment at 723 K, while bands attributable to anhydride, quinone, ester and phenol species remained. Small amounts of ether, furan and phenol groups were detected on the graphitized fiber surface, while ketonic carbonyl groups were dominant on diamond. Significant amounts of chemisorbed hydrogen were also detected, presumably occurring on edge atoms made available by the decomposition of CO-yielding complexes at temperatures >873 K.  相似文献   

16.
The gasification of a very high purity natural graphite was studied at temperatures between 960 and 1120°C and at CO2 pressures below 108 millitorr. For CO2 depletion up to at least 90%, gasification rates were first order in CO2 pressure, that is with no inhibition by CO observed. The activation energy for the rate constant for the oxygen transfer step (103.5 ± 5.8 kcal/mole) agreed within experimental error with that found from kinetic studies at intermediate CO2 pressures where CO does inhibit the reaction. The rate of the oxygen transfer reaction is markedly inhibited by the presence of low pressures of H2. As H2 pressure is increased up to 3 millitorr, the gasification rate in CO2 at 1100°C monotonically decreases. Further increase in H2 pressure, has a negligible effect on rate. From measurements of hydrogen uptake at reaction temperature, it is clear that inhibition is caused by dissociative chemisorption of hydrogen on to active sites. Inhibition by hydrogen is even more marked for the Graphon-CO2 reaction and is attributed not only to its chemisorbing on carbon sites but also on to impurity catalyst sites. It is doubtful if true rate constants for the C-CO2 reaction, uninhibited by hydrogen, have ever been reported.  相似文献   

17.
N2O decomposition was investigated over a series of K-promoted Co-Al catalysts. The activity tests showed that doping with K greatly enhanced the catalytic activity of the Co-Al catalyst, and the enhancement was critically dependent on the amount of K and the calcination temperature. When the catalyst had a K/Co atomic ratio of 0.04 and was calcined at 700–800 °C, a full N2O conversion could be reached at a reaction temperature of 300 °C. Moreover, even under the simultaneous presence of 4% O2 and 2.6% water vapor, such high-temperature treated K/Co-Al catalyst exhibited high reactivity and stability, with the N2O conversion remaining at a constant value of 92% over 40 h run at 360 °C. In contrast, non-doped Co-Al catalyst showed a severe activity loss under such reaction conditions. A combination of characterization techniques was employed to reveal the promoting role of K and the effect of calcination temperature. The results suggest that doping with K increases the electron density of Co and weakens the Co–O bond, thus promoting the activation of N2O on the Co sites and facilitating the desorption of oxygen from the catalyst surface. High-temperature calcinations made the desorption of O2 proceed more readily.  相似文献   

18.
Surface oxygen complexes on Spheron 6, which thermally desorb as CO2, appear to be responsible for the acidity of the carbon. The base uptake of samples degassed at various temperatures has been related to the amounts of oxygen complex remaining on the surface. Two types of acidic oxides, both of which desorbed as CO2, were observed. At temperatures around 250°C an oxide which acts as a very weak monobasic acid is decomposed and at about 600°C a second oxide, which is stronger and dibasic is decomposed. The heat of neutralization of this second acid was found to be approximately 12 kcal-mole−1.  相似文献   

19.
A novel metal–organic framework (MOF) [Zn(TPT)(MPDCO)(H2O)·H2O]n (TPT = 1,2,3,4-tetra (4-pyridyl)thiophene, MPDCO = 6-methylpyridine-2,4-dicarboxylic acid N-oxide) (1) was obtained via hydrothermal synthesis and characterized by the elemental analysis, IR spectroscopy, TG analysis, luminescent spectroscopy and single-crystal X-ray diffraction analysis. Single-crystal structural analysis reveals that a one-dimensional (1-D) ladderlike MOF is constructed through the mixed ligands, and it is further connected by the intermolecular hydrogen bonds to form a 3-D supramolecular network. Complex 1 exhibits efficient blue luminescence at room temperature, and the framework is stable below 235 °C.  相似文献   

20.
Adsorption and desorption of hydrogen and coadsorption of H2 with H2O and CO over Ru(00l) surface have been studied under UHV conditions using the technique of TDS. Surface hydrogen interacts with adsorbed water resulting in an additional desorption state at 510 K which is not easily displaced by CO, but the total number of adsorption sites for hydrogen adsorption is independent of the amount of H2O predosed at room temperature. Hydrogen adsorption is blocked easily by CO dose (more than 0.5 L) or a small amount of O(a) formed from dissociative adsorption of water, and adsorbed hydrogen formed in the absence of significant water is easily displaced by CO dose at even room temperature.  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