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 15 毫秒
1.
Zhong Qin 《Thermochimica Acta》1995,260(1-2):125-136
The direct sulfation reaction of SO2 with CaCO3 has been investigated by thermogravimetry (TG) under the condition that the CaCO3 does not decompose to CaO prior to sulfation by controlling CO2 partial pressure. The direct sulfation process can be described by using a shrinking-core model for constant particle size. The model shows that the reaction rate and the diffusion rate of SO2 through the product layer are equally important. Temperature effects can be correlated by the activation energy of 35.9 kJ mol−1 for the sulfation reaction and 66.5 kJ mol−1 for the product layer diffusion. The sulfation reaction is found to be first order with respect to SO2. With larger pore volume and surface area of limestone samples, the sorbents have a stronger reactivity of SO2 removal. A 70% CaCO3 conversion can be achieved in 10 min at 800°C and 2000 ppm SO2.  相似文献   

2.
The hydrogen permeation and stability of tubular palladium alloy (Pd–23%Ag) composite membranes have been investigated at elevated temperatures and pressures. In our analysis we differentiate between dilution of hydrogen by other gas components, hydrogen depletion along the membrane length, concentration polarization adjacent to the membrane surface, and effects due to surface adsorption, on the hydrogen flux. A maximum H2 flux of 1223 mL cm−2 min−1 or 8.4 mol m−2 s−1 was obtained at 400 °C and 26 bar hydrogen feed pressure, corresponding to a permeance of 6.4 × 10−3 mol m−2 s−1 Pa−0.5. A good linear relationship was found between hydrogen flux and pressure as predicted for rate controlling bulk diffusion. In a mixture of 50% H2 + 50% N2 a maximum H2 flux of 230 mL cm−2 min−1 and separation factor of 1400 were achieved at 26 bar. The large reduction in hydrogen flux is mainly caused by the build-up of a hydrogen-depleted concentration polarization layer adjacent to the membrane due to insufficient mass transport in the gas phase. Substituting N2 with CO2 results in further reduction of flux, but not as large as for CO where adsorption prevail as the dominating flow controlling factor. In WGS conditions (57.5% H2, 18.7% CO2, 3.8% CO, 1.2% CH4 and 18.7% steam), a H2 permeance of 1.1 × 10−3 mol m−2 s−1 Pa−0.5 was found at 400 °C and 26 bar feed pressure. Operating the membrane for 500 h under various conditions (WGS and H2 + N2 mixtures) at 26 bars indicated no membrane failure, but a small decrease in flux. A peculiar flux inhibiting effect of long term exposure to high concentration of N2 was observed. The membrane surface was deformed and expanded after operation, mainly following the topography of the macroporous support.  相似文献   

3.
Mercuric 5-nitrotetrazole is a possible replacement for lead azide. The thermal decomposition peak maximum ranged from 185 to 270°C as the heating rate increased from 0.1 to 100°C min−1. The activation energy and frequency factor for thermal decomposition were determined from dynamic and isothermal DSC and isothermal TG data; the average values were 38.8 kcal mol−1 and 3.56×1014 s−1. A half-life experiment confirmed the kinetic constants and indicated that the decomposition reaction was first order. The heat of explosion was determined by a pressure DSC test and found to be 2587 J g−1. The linear coefficient of expansion was 37±2×10−6°C−1 from −60 to 160°C and indicated secondary transitions near −10 and 90°C. The specific heat was 0.0003154T+0.1339 in the region −40–90°C. The critical temperature for a slab with a half-thickness of 0.035 cm was calculated to be 232 °C.  相似文献   

4.
A series of γ-Al2O3 samples modified with various contents of sulfate (0–15 wt.%) and calcined at different temperatures (350–750 °C) were prepared by an impregnation method and physically admixed with CuO–ZnO–Al2O3 methanol synthesis catalyst to form hybrid catalysts. The direct synthesis of dimethyl ether (DME) from syngas was carried out over the prepared hybrid catalysts under pressurized fixed-bed continuous flow conditions. The results revealed that the catalytic activity of SO42−/γ-Al2O3 for methanol dehydration increased significantly when the content of sulfate increased to 10 wt.%, resulting in the increase in both DME selectivity and CO conversion. However, when the content of sulfate of SO42−/γ-Al2O3 was further increased to 15 wt.%, the activity for methanol dehydration was increased, and the selectivity for DME decreased slightly as reflected in the increased formation of byproducts like hydrocarbons and CO2. On the other hand, when the calcination temperature of SO42−/γ-Al2O3 increased from 350 °C to 550 °C, both the CO conversion and the DME selectivity increased gradually, accompanied with the decreased formation of CO2. Nevertheless, a further increase in calcination temperature to 750 °C remarkably decreased the catalytic activity of SO42−/γ-Al2O3 for methanol dehydration, resulting in the significant decline in both DME selectivity and CO conversion. The hybrid catalyst containing the SO42−/γ-Al2O3 with 10 wt.% sulfate and calcined at 550 °C exhibited the highest selectivity and yield for the synthesis of DME.  相似文献   

5.
A thermoanalyzer (Mettler) combined with a quadrupole mass spectrometer (Balzers) by a capillary inlet system allows simultaneous DTA, TG and evolved gas analysis in different atmospheres. Decomposition of CaC2O4·H2O in air and argon, respectively, demonstrates the usefulness of the mass spectrometer for the quantitative determination of H2O, CO2 and CO. Decomposition of NaHCO3 at a heating rate of 10°C min−1 reveals that H2O and CO2 evolved simultaneously at a relatively low temperature (159°C) can also be determined quantitatively and nearly without retardation compared with the weight loss step. In the investigation of clays an example will be given of the usefulness of the described DTA—TG—MS in the quantitative interpretation of overlapping reactions.  相似文献   

6.
在循环煅烧/碳酸化反应系统中,研究了SO2对钙基吸收剂CaCO3捕集CO2的影响,获得了SO2对钙基吸收剂碳酸化特性、煅烧特性以及循环稳定性的影响规律,并结合SEM分析结果,从循环煅烧/碳酸化反应角度,分析了可能存在的原因。结果表明,钙基吸收剂吸收CO2的能力随着循环反应次数的增加逐渐发生衰减,在SO2影响下,这种衰减会进一步加剧,且衰减程度随着SO2浓度的增加而增大,经过十次循环后,碳酸化转化率分别为25.5%(0%SO2)、16.9%(0.1%SO2)和5.2%(0.2%SO2)。造成这种衰减加剧的主要原因是反应产生较厚的硫酸化产物层,硫酸化产物层使颗粒表面孔隙发生堵塞,阻碍了CO2在吸收剂内部的扩散,降低了碳酸化转化率。  相似文献   

7.
The spectrum of SO (X3Σ) has been observed following the flash excitation of sulphur dioxide with radiation above 250 nm. Sulphur monoxide is produced via an excited molecule mechanism involving triplet SO2. The rate constant for the reaction 3SO2 + SO2 was measured as (3.1 ± 1) × 108 M−1 sec−1.  相似文献   

8.
The tridecameric aluminum polymer [AlO4Al12(OH)24(H2O)12]7+ was prepared by forced hydrolysis of Al3+ up to an OH/Al molar ratio of 2.2. Upon addition of sulfate, the tridecamer crystallized as the monoclinic basic aluminum sulfate Na0.1[AlO4Al12(OH)24(H2O)12](SO4)3.55. The dehydroxylation of the basic aluminum sulfate has been studied by Fourier transform in-situ infrared emission spectroscopy over a temperature range of 200° to 750°C at 50°C intervals. The spectrum is characterized by the sulfate ν1 (1024 cm−1), ν3 doublet (1117 and 1168 cm−1) and the ν4 doublet (568 and 611 cm−1) modes. Furthermore, minor bands assigned to nitrate are observed. Upon heating from ≈350° to 400°C major changes are observed, especially in the bandwidth and band intensities. The bands in the hydroxyl stretching region due to the Al13 group disappear, whereas the bands around 1050 cm−1 display various changes in bandwidths, intensities and positions associated with the dehydration and dehydroxylation of the basic sulfate and the changing of the structure into an aluminum oxosulfate. The nitrate bands diminish upon heating.  相似文献   

9.
The collisional quenching of electronically excited germanium atoms, Ge[4p2(1S0)], 2.029 eV above the 4p2(3P0) ground state, has been investigated by time-resolved atomic resonance absorption spectroscopy in the ultraviolet at λ = 274.04 nm [4d(1P10) ← 4p2(1S0)]. In contrast to previous investigations using the ‘single-shot mode’ at high energy, Ge(1S0) has been generated by the repetitive pulsed irradiation of Ge(CH3)4 in the presence of excess helium gas and added gases in a slow flow system, kinetically equivalent to a static system. This technique was originally developed for the study of Ge[4p2(1D2)] which had eluded direct quantitative kinetic study until recently. Absolute second-order rate constants obtained using signal averaging techniques from data capture of total digitised atomic decay profiles are reported for the removal of Ge(1S0) with the following gases (kR in cm3 molecule−1 s−1, 300 K): Xe, 7.1 ± 0.4 × 10−13; N2, 4.7 ± 0.6 × 10−12; O2, 3.6 ± 0.9 × 10−11; NO, 1.5 ± 0.3 × 10−11; CO, 3.4 ± 0.5 × 10−12; N2O, 4.5 ± 0.5 × 10−12; CO2, 1.1 ± 0.3 × 10−11; CH4, 1.7 ± 0.2 × 10−11; CF4, 4.8 ± 0.3 × 10−12; SF6, 9.5 ± 1.0 × 10−13; C2H4, 3.3 ± 0.1 × 10−10; C2H2, 2.9 ± 0.2 × 10−10; Ge(CH3)4, 5.4 ± 0.2 × 10−11. The results are compared with previous data for Ge(1S0) derived in the single-shot mode where there is general agreement though with some exceptions which are discussed. The present data are also compared with analogous quenching rate data for the collisional removal of the lower lying Ge[4p2(1D2)] state (0.883 eV), also characterized by signal averaging methods similar to that described here.  相似文献   

10.
This work presents chemical modeling of solubilities of metal sulfates in aqueous solutions of sulfuric acid at high temperatures. Calculations were compared with experimental solubility measurements of hematite (Fe2O3) in aqueous ternary and quaternary systems of H2SO4, MgSO4 and Al2(SO4)3 at high temperatures. A hybrid model of ion-association and electrolyte non-random two liquid (ENRTL) theory was employed to fit solubility data in three ternary systems H2SO4–MgSO4–H2O, H2SO4–Al2(SO4)3–H2O at 235–270 °C and H2SO4–Fe2(SO4)3–H2O at 150–270 °C. Employing the Aspen Plus™ property program, the electrolyte NRTL local composition model was used for calculating activity coefficients of the ions Al3+, Mg2+ Fe3+ and SO42−, HSO4, OH, H3O+, respectively, as well as molecular species. The solid phases were hydronium alunite (H3O)Al3(SO4)2(OH)6, hematite Fe2O3 and magnesium sulfate monohydrate (MgSO4)·H2O which were employed as constraint precipitation solids in calculating the metal sulfate solubilities. A correlation for the equilibrium constants of the association reactions of complex species versus temperature was implemented. Based on the maximum-likelihood principle, the binary interaction energy parameters for the ionic species as well as the coefficients for equilibrium constants of the reactions were obtained simultaneously using the solubility data of the ternary systems. Following that, the solubilities of metal sulfates in the quaternary systems H2SO4–Fe2(SO4)3–MgSO4–H2O, H2SO4–Fe2(SO4)3–Al2(SO4)3–H2O at 250 °C and H2SO4–Al2(SO4)3–MgSO4–H2O at 230–270 °C were predicted. The calculated results were in excellent agreement with the experimental data.  相似文献   

11.
The reaction: F + HCl→ HF (v 3) + Cl (1), has been initiated by photolysing F2 using the fourth-harmonic output at 266 nm from a repetitively pulsed Nd: YAG laser By analysing the time-dependence of the HF(3,0) vibrational chemiluminescence, rate constants have been determined at (296 ± 5) K for reaction (1), k1 = (7.0 ± 0.5) × 10−12 cm3 molecule−1 s−1, and for the relaxation of HF(v = 3) by HCl, CO2, N2O, CO, N2 and O2: kHCl = (1.18 ±0.14) × 10−11 kCO2 = (1.04 ± 0. 13) × 10−12, kN2O = (1.41 ± 0.13) × 10−11 kCO = (2.9 ± 0.3) × (10−12, kN2 = (7.1 ± 0.6) × 10−14 and kO2 = (1.9 ± 0.6) × 10−14 cm3molecule−1s−1.  相似文献   

12.
The second-order rate constants of gas-phase Lu(2D3/2) with O2, N2O and CO2 from 348 to 573 K are reported. In all cases, the reactions are relatively fast with small barriers. The disappearance rates are independent of total pressure indicating bimolecular abstraction processes. The bimolecular rate constants (in molecule−1 cm3 s−1) are described in Arrhenius form by k(O2)=(2.3±0.4)×10−10exp(−3.1±0.7 kJmol−1/RT), k(N2O)=(2.2±0.4)×10−10exp(−7.1±0.8 kJmol−1/RT), k(CO2)=(2.0±0.6)×10−10exp(−7.6±1.3 kJmol−1/RT), where the uncertainties are ±2σ.  相似文献   

13.
A high-resolution emission spectrum of a low-pressure Ar-diluted CO + N2O → CO2 + N2 flame catalyzed by Na metal vapor has been obtained and examined for vibrational disequilibrium. Emission in the 1900-2400 cm−1 spectral region, which includes the fundamental and “hot” bands of CO, CO23), and N2O(ν3), was recorded with high resolution and the CO emission was analyzed in detail to determine vibrational and rotational temperatures which were found to be unequal, Tv = 2050°K and TR = 1100°K. An examination of vib-vib and vib-trans energy transfer mechanisms results in the conclusion that an excess of 14% of the chemical energy is preferentially deposited in the resonantly-coupled N2, CO, CO23), and N2O(ν3) vibrational modes. It is further observed that CO vibrational levels for ν > 4 are excessively populated, presumably due to quenching of Na*(3p) by CO; the flame is accompanied by intense Na D-line chemiluminescence.  相似文献   

14.
电石制备过程中不同含钙化合物与焦炭的反应行为研究   总被引:1,自引:0,他引:1  
针对粉状原料-氧热法电石生产新工艺的需求,通过程序升温法研究了多种粉状含钙原料与粉状焦炭直接反应制备电石的过程。结果表明,升温过程中CaCO3释放的CO2、Ca(OH)2和电石渣释放的H2O对焦炭的量和质的影响很小,CaCO3、Ca(OH)2和电石渣均可直接用于电石生产,电石生成反应自1 450℃开始,1 740℃左右达到峰值。CaSO4与焦炭在920℃左右反应形成CaS,在研究的温度范围内CaS不与焦炭反应。  相似文献   

15.
An efficient photocatalytic reduction of carbon dioxide to HCOOH and HCHO is reported using K[Ru(H-EDTA)Cl] · 2H2O (1) as homogeneous catalyst and particulate Pt—CdS—RuO2 as photon absorber at 505 nm. This system produces 0.22 M of HCOOH and 0.10 M of HCHO in 6 h of photolysis at rates of 3.05 × 10−2 M h−1 and 2.0 × 10−2 M h−1 respectively. Trace amounts of CH3OH, CH4 and CO are detected in the reaction vessel. The rates of formation of HCOOH and HCHO exhibit a first-order dependence on the catalyst and dissolved CO2 concentrations. The reaction shows deuterium isotope effects (kH/kD) of 1.5 and 2.00 for the formation of HCOOH and HCHO respectively. Under identical experimental conditions, the rate of decomposition of formate was studied. The rate of decomposition of formate is slower (by two orders of magnitude compared with the formation of formate) even at high formate concentrations. A mechanism for the formation of HCOOH and HCHO is proposed.  相似文献   

16.
The gas-phase Boudouard disproportionation reaction between two highly vibrationally excited CO molecules in the ground electronic state has been studied in optically pumped CO. The gas temperature and the CO vibrational level populations in the reaction region, as well as the CO2 concentration in the reaction products have been measured using FTIR emission and absorption spectroscopy. The results demonstrate that CO2 formation in the optically pumped reactor is controlled by the high CO vibrational level populations, rather than by CO partial pressure or by flow temperature. The disproportionation reaction rate constant has been determined from the measured CO2 and CO concentrations using the perfectly stirred reactor (PSR) approximation. The reaction activation energy, 11.6 ± 0.3 eV (close to the CO dissociation energy of 11.09 eV), was evaluated using the statistical transition state theory, by comparing the dependence of the measured CO2 concentration and of the calculated reaction rate constant on helium partial pressure. The disproportionation reaction rate constant measured at the present conditions is kf = (9 ± 4) × 10−18 cm3/s. The reaction rate constants obtained from the experimental measurements and from the transition state theory are in good agreement.  相似文献   

17.
The rate constant for the reaction between the sulphate radical (SO4√−) and the ruthenium (II) tris-bipyridyl dication (Ru(bipy)32+) is (3.3±0.2)×109 mol−1 dm3 s−1 in 1 mol dm−3 H2SO4 and (4.9±0.5)×109 mol−1 dm3 s−1 in 0.1 mol dm−3, pH 4.7 acetate buffer. The SO4√−radical produced by the electron transfer quenching of Ru(bipy)32+* by S2O82− reacts rapidly with both acetate buffer and chloride ions. These side reactions result in a reduction in the overall quantum yield of Ru(bipy)33+ production and reduced reaction selectivity when Ru(bipy)32+* is quenched by persulphate.  相似文献   

18.
Time profiles of weight change of coal samples and the evolution of low molecular weight gases (H2, CH4, CO and CO2) in both steam gasification and pyrolysis of Yallourn brown coal and Taiheiyo subbituminous coal were measured using a thermobalance reactor with a micro GC and a mass spectrometer, in order to examine the reaction mechanism of steam gasification with rapid heating (100 K s−1). It was found that, in the case of slow heating, steam reacted with metaplast and promoted the evolution of tar above 623 K and that a water shift reaction took place above 873 K. Steam gasification of produced char occurred above 1023 K, increasing the evolution of CO, CO2 and H2. When the heating rate was high, steam reforming of volatile matter and steam gasification of metaplast took place parallel to metaplast formation and condensation. The char produced by pyrolysis was almost completely gasified and converted into H2 and CO2 by steam. The chemical energy of coal was mainly converted into hydrogen energy and the gasification efficiency was slightly increased by rapid heating (i.e. 100 K s−1).  相似文献   

19.
During thermolysis of the 7-germanorbornadiene 1 in chlorobenzene at 70°C in the presence of concentrated hydrochloric acid, besides the well-known formation of free germylene Me2Ge and its consecutive product dimethylchlorogermane 2, the polar splitting of only one Ge---C bond in 1 has been observed for the first time. It does not yield Me2Ge, but instead it rapidly forms the 1-germyl-1,4-dihydronaphthalene 3. The kinetics of this reaction at 53°C are of 2nd order, t1/2 = 20 min, k = 0.22 l mol−1 min−1. At room temperature 3 is formed quantitatively. Also, at 70°C the slower formation of the germylene Me2Ge from 1 can be suppressed completely if HCl gas is bubbled through the reaction mixture, thus favouring the rapid formation of 3. As a by-product the 1,2-dihydronaphthalene 5 is generated.  相似文献   

20.
Solid acids – NiSO4/Al2O3, Fe2(SO4)3/Al2O3 and TiO2/SO42− – appeared to be effective catalysts for the acid catalyzed synthesis of methyl ester of trifluoropyruvic acid. They are active at 150–180 °C.  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