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 312 毫秒
1.
Fine particles of needle‐like lepidocrocite (γ‐FeOOH) were synthesized by the oxidation of aqueous suspensions of ferrous hydroxide using a bubble column with draft tube at a constant temperature ranging from 20°C to 30°C. The oxidation steps leading to green rust (an intermediate) and lepidocrocite (final product), termed Step I and II, respectively, could be described apparently as first order, with respect to oxygen, and zero order, with respect to total ferrous species. When the concentration of oxygen in the feed stream was varied under a constant gas velocity, the mean size based on the major axis of needle‐like particle decreased from 0.60 to 0.35 μm with increasing oxidation rate. When the gas velocity was varied under a constant oxygen concentration, the particle size was almost independent of the oxidation rate and was equal to ca. 0.6 μm. By the addition of a small amount of sodium dihydrogenphosphate (NaH2PO4), the major axis could be reduced to 0.2 μm with the minor axis and the oxidation rate almost unchanged.  相似文献   

2.
The temperature, supersaturation, seeding procedure, stirring speed and other parameters were varied in crystallization experiments of calcium carbonate performed in aqueous solutions to control size, particle size distribution and morphology of the particles. Particle size information was obtained by focused beam reflectance measurements and the Coulter Counter Multisizer. Crystals of CaCO3 could be crystallized as spherical polycrystalline particles of the vaterite polymorph, needle‐like crystals of aragonite and both cube‐like and novel plate‐like crystals of calcite. Filtration experiments for calcium carbonate, performed at a constant pressure difference of 2 bar, show that spherical particles with a larger size show better filterability and that spheres with a wider size distribution, as a result of high supersaturation and nucleation, give higher average cake resistance values. Comparing different particle morphologies, plate‐like crystals and needle‐like crystals show worse filterability than spherical particles and cube‐like particles. © 2011 American Institute of Chemical Engineers AIChE J, 2012  相似文献   

3.
This work proposed a synthesis route of ZSM‐5 via the hydrothermal method with premixing in a stirred tank reactor (STR). Effects of various operating conditions, including pre‐mixing time, molar ratio of SiO2/Al2O3, TPAOH (organic template agents) concentration, NaCl (alkali metal cations) concentration, crystallization temperature, and crystallization reaction time, on the average particle size (PS) and particle size distribution (PSD) were investigated. It was found that the pre‐mixing time in the STR significantly affect the formation of proto‐nuclei in premixing process and crystal growth in hydrothermal reaction process, and consequently influence the PS and PSD of the prepared ZSM‐5. ZSM‐5 with good thermal stability, a PS of 380 nm, PSD of 0.17–0.9 µm, pore diameter of 2.31 nm, pore volume of 0.19 cm3 · g?1 and specific surface area of 337.25 m2 · g?1 were obtained under the optimal conditions of a crystallization reaction time of 24 h, a crystallization temperature of 130 °C, a molar ratio of SiO2/Al2O3 of 200, a TPAOH concentration of 3.5 mol · L?1, NaCl concentration of 0.3 mol · L?1, and a pre‐mixing time of 5 h. This work indicated that the operating conditions including premixing time have a significant effect on its PS and PSD.  相似文献   

4.
Observed pseudo‐first‐order rate constants (ko) for the reaction between CO2 and 2‐((2‐aminoethyl) amino) ethanol (AEEA) were measured using the stopped‐flow technique in an aqueous system at 298, 303, 308 and 313 K, and in non‐aqueous systems of methanol and ethanol at 293, 298, 303 and 308 K. Alkanolamine concentrations ranged from 9.93 to 80.29 mol m?3 for the aqueous system, 29.99–88.3 mol m?3 for methanol and 44.17–99.28 mol m?3 for ethanol. Experimentally obtained rate constants were correlated with two mechanisms. For both the aqueous‐ and non‐aqueous‐AEEA systems, the zwitterion mechanism with a fast deprotonation step correlated the data well as assessed by the reported statistical analysis. As expected, the reaction rate of CO2 in the aqueous‐AEEA system was found to be much faster than in methanol or ethanol. Compared to other promising amines and diamines studied using the stopped‐flow apparatus, the pseudo‐first‐order reaction rate constants were found to obey the following order: PZ (cyclic‐diamine) > EDA (diamine) > AEEA (diamine) > 3‐AP (primary amine) > MEA (primary amine) > EEA (primary amine) > MO (cyclic‐amine). The reaction rate constant of CO2 in aqueous‐AEEA was double that in aqueous‐MEA, and the difference increased with an increase in concentration. All reaction orders were practically unity. With a higher capacity for carbon dioxide and a higher reaction rate, AEEA could have been a good substitute to MEA if not for its high thermal degradation. AEEA kinetic behaviour is still of interest as a degradation product of MEA. © 2012 Canadian Society for Chemical Engineering  相似文献   

5.
The purpose of this study was to determine major factors impacting the size of simvastatin (SIM)‐loaded poly(d , l ‐lactic‐co‐glycolide) (PLGA) nanoparticles (NPs) that was prepared using electrospraying. Three variables including concentration of polymer and salt as well as solvent flow rate were used as input variables. Size of NPs was considered as output variable. For the first time, our findings using a systematic and experimental approach, showed the importance of salt concentration as the dominant factor determining the size with a sharp and reverse effect. Optimum formulation (i.e., flow rate 0.08 mL h?1, polymer concentration 0.7 w/v %, and salt concentration 0.8 mM) was then evaluated for aqueous solubility, encapsulation efficiency, particle size, in vitro drug release pattern and cytotoxicity. A very appreciable encapsulation efficiency (90.3%) as well as sustained release profile, considerable enhancement in aqueous solubility (~5.8 fold) and high IC50 (>600 µM of SIM‐loaded PLGA NPs) indicated PLGA as a promising nanocarrier for SIM. The optimum formulation had particle size, zeta potential value, polydispersity index (PDI) and drug loading of 166 nm, +3 mV, 0.62 and 9%, respectively. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 43602.  相似文献   

6.
The electro‐rheological (ER) behaviour of suspensions in silicone oil of phosphoric ester cellulose particles (average particle size 17.77 µm) was investigated at room temperature with electric fields up to 2.5 kV mm−1 with the aim of developing anhydrous ER suspensions applicable over a wide temperature range. Anhydrous ER suspensions dispersing cellulose particles which were treated by phosphoric ester reaction of 2 mol phosphoric acid and 4 mol urea were measured; not only were the electrical characteristics such as dielectric constant, current density and electrical conductivity studied but also the rheological properties on the electric fields and volume fraction of dispersing particles. Anhydrous ER suspensions dispersing phosphoric ester cellulose particles showed very low current density, conductivity and very high performance of ER effect (τA/ τ0 ≈ 1030) on the 2.5 kV mm−1 electric field, and the yield stress (τy) had a quadratic dependence on electric fields. © 2000 Society of Chemical Industry  相似文献   

7.
Experimental data on the kinetics of methane gas hydrate decomposition are reported. The isothermal/isobaric semi‐batch stirred‐tank reactor, used by Kim et al. (1987), was modified to include an on‐line particle size analyzer. The experiments were conducted at temperatures ranging from 274.65 K to 281.15 K and at pressures between 3.1 and 6.1 MPa. The model of Clarke and Bishnoi (1999, 2000) was used to determine the intrinsic rate constant. It was found that the activation energy for methane hydrate decomposition is 81 kJ/mol and the intrinsic rate constant of decomposition is 3.6 × 104 mol/m2 Pa.s.  相似文献   

8.
Methane oxidation in soils with different textures and land use   总被引:5,自引:0,他引:5  
Intact core samples from soils with different textures and land use were tested for their capacity to oxidise methane. The soil cores were taken from arable land, grassland and forest. It was found that coarse textured soils (6.74–16.38 μg CH4 m-2 h-1) showed a higher methane uptake rate than fine textured soils (4.66–5.34 μg CH4 m-2 h-1). Increasing soil tortuosity was thought to reduce the methane oxidation rate in fine textured soils. The oxidation rate of forest soils (16.32–16.38 μg CH4 m-2 h-1), even with a pH below 4.5, was very pronounced and higher than arable land (11.40–14.47 μg CH4 m-2 h-1) and grassland (6.74–9.30 μg CH4 m-2 h-1). Within the same textural class arable land showed a faster methane uptake rate than grassland. In grassland with a fine texture, even methane production was observed. Nitrogen availability and turnover in these land use systems were thought to cause the different oxidation rates. Decreasing the moisture content slowed down the oxidation rate in all soils. This could be caused by an increased N turnover and a starvation of the methanotrophic bacteria. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

9.
Abstract

The preparation of ground low‐density poly(ethylene) (LDPE), high‐density poly(ethylene) (HDPE), and poly(ethylene terephthalate) (PET) materials suitable for use by inverse gas chromatography (IGC) experiments is described. Materials prepared showed a smaller and more uniform particle size, 83–86 µm (200 mesh), than commercially available Chromosorb® W solid chromatography support. In addition, the BET surface area of polymeric materials, 0.367–0.559 m2/cm3, was much greater than the surface area of 150‐ to 212‐µm‐diameter glass beads. The use of polymeric materials with uniform particle sizes and defined surface areas have several advantages over polymer coated Chromosorb® W and glass beads in IGC experiments.  相似文献   

10.
A new approach to kinetic studies of fat hydrogenation is discussed. An experimental setup is described in detail. An example of reactor performance in hydrogenation of fatty acid methyl esters is given. aPresent address: AB Karlshamns Oljefabriker, S-292 00 Karlshamn, Sweden. Notation: c, outlet concentration, mol/m3 ; co, inlet concentration, mol/m3: cb, concentration in bulk fluid, mol/m3 ; cs concentration at catalyst surface, mol/m3 ; d,pore diameter, m; De, effective diffusivity, m2/s; E, activation energy of reaction, J/mol; h, heat transfer coefficient, J/m2 s K; ΔH, heat of reaction, J/mol; kc, mass transfer coefficient, m/s; p, partial pressure, Pa; po, saturated vapor pressure, Pa; qf, total flow to reactor, m3/s; qrec, recycle flow, m3/s; R, observed reaction rate per unit particle volume, mol/s m3 ; Rg, gas constant, J/mol K; r, observed reaction rate per unit mass of catalyst, mol/kg s; rp particle radius, m; Tb, temperature of gas in bulk flow, K; Ts, temperature of gas at catalyst surface, K;-v, molar volume, m3/mol; V, volume of catalyst bed, m3 ; W, catalyst mass, kg. Greek symbols: σ, surface tension, N/m; λe, effective thermal conductivity of catalyst particle, J/m s K.  相似文献   

11.
Halogenated phenols and bisphenols are recognized as being recalcitrant in conventional biological treatments. The current research evaluated lignin peroxidase‐catalyzed oxidation and polymerization as a potential alternative for their detoxification. Gel permeation–HPLC analysis demonstrated the formation of dimers, trimers and tetramers upon oxidation of the target substrates. Polymerization was accompanied by effective detoxification of the aqueous phase during oxidation of 2,4‐dibromophenol, the extent of which correlated with the extent of oxidation and polymerization. Steady state kinetic measurements at a saturating concentration of H2O2 revealed high Km values (270–1100 µmol dm?3) for the target substrates, reflecting the strong electron‐withdrawing properties of halogen substituents, which increase the oxidation potential of the phenols, resulting in thermodynamically less favorable reactions. However, kcat values were not dissimilar from non‐halogenated phenols and the rapid oxidation and polymerization suggests that low retention times could be expected in a continuous process for their treatment, in contrast to conventional biological methods. The operational stability of lignin peroxidase was significantly improved by inclusion of redox mediators, which resulted in enhanced oxidation and more rapid reaction rates. However, due to their inherent toxicity, the use of redox mediators impeded toxicity assays. The findings highlight the potential of lignin peroxidase as a possible alternative for the high‐rate treatment of industrial wastewater when conventional methods are ineffective. Copyright © 2003 Society of Chemical Industry  相似文献   

12.
Char gasification has been studied in different ways with different feedstock; however, the fundamental studies about the variation of char reactivity with different combination of parameters are still required in order to design the biomass char gasification process in large scale unit. In this work, char from flax straw pyrolysis was used for gasification with different partial pressure of CO2, different temperature and particle size. Results showed that 375‐µm particle sizes of char has higher reactivity compared to other particle sizes and the inhibition effect was also less at 375‐µm particle size. Kinetic parameters varied for gasification reaction with different particle sizes and the average activation energy was 196 kJ/mol and the order of the reaction was approximately 1. Inhibition studies with the addition of CO in the gasifying agent proved that CO molecules interfere significantly and reduced the reactivity. ANOVA test showed that the temperature plays a vital role in char reactivity. The particle sizes of 375 and 800 µm along with the CO2 partial pressure of 0.35 bars are the best combination for achieving the maximum reactivity. © 2012 Canadian Society for Chemical Engineering  相似文献   

13.
BACKGROUND: Supercritical antisolvent (SAS) micronisation of synthetic trans ‐β‐carotene was studied using tetrahydrofuran (THF) as solvent and supercritical carbon dioxide (CO2) as antisolvent, with the objective of increasing its bioavailability and facilitating its dispersion in oil and emulsion formulations as a result of its smaller particle size. The micronised powder was analysed by scanning electron microscopy and high‐performance liquid chromatography. Micronisation experiments were performed in order to evaluate the effects of temperature (308.15–333.15 K), pressure (6.5–13 MPa) and concentration of the liquid solution (6–9 g L?1). The effect of the supercritical CO2/THF flow ratio in the range between 4 and 44 (on a mass basis) was also analysed. Determinations of equilibrium concentrations of β‐carotene in the CO2/THF mixture were also performed. RESULTS: The particle size obtained ranged from 1 to 500 µm, with mean particle diameters around 100 µm. Three types of morphology were found in the precipitated powder: crystalline with superficial pores and leaf‐like appearance; crystalline with regular shapes and blade‐like edges; and crystalline without superficial pores and leaf‐like apearance. The Peng–Robinson equation of state was used to calculate the density of the CO2/THF binary mixture, and the solubility of β‐carotene in this mixture was correlated with its density. CONCLUSION: The use of the SAS technique to micronise β‐carotene proved to be efficient, and the absence of degradation in the micronised powder allows the industrial application of this technique. Copyright © 2008 Society of Chemical Industry  相似文献   

14.
Adsorption equilibrium constants for methyl oleate and methyl linoleate in vapor phase on supported copper and nickel catalysts have been determined using the technique of pulse gas chromatography. The results are discussed in relation to selectivity in fat hydrogenation. Notation: A, column cross-section, m2 ; an,bn, nth Fourier coefficients; c, concentration of adsorbate in bulk flow, mol/m3 ; c* = c/ ∫ 0 cdt, normalized concentration of adsorbate in bulk flow; Ci, concentration of adsorbate in catalyst pores, mol/m3 ; ca, concentration of adsorbate on catalyst surface, mol/kg; cTOT, active area of catalyst as measured by hydrogen adsorption, mol/kg; De, effective diffusion coefficient of adsorbate in catalyst, m2/s; Dea, axial dispersion coefficient based on void cross-section, m2/s; hn, nth coefficient in Hermite polynomial expansion; Hn nth Hermite polynomial; ΔHA, adsorption enthalpy, kJ/mol; ΔHvap , heat of vaporization, kJ/moll; ka, adsorption rate constant, m3/kgs;KA, adsorption equilibrium constant, m3/kg; K0 , preexponential factor defined in Eqn. 8, m3/kg; kf, mass transfer coefficient, m/s; L, bed length, m; q, flow rate, m3/s; R, particle radius, m; Rg, gas constant; t, time, s; T, temperature, K; TF, period of Fourier expansion, s; u = q/A, linear velocity, m/s; z, length coordinate in packed column, m. Greek symbols: δ(t), Dirac delta function; ∈B, void fraction of bed; ∈-p, particle void fraction, ρrp, particle density, kg/m3 ; ξ, radial coordinate in particle, m; μ1, first absolute moment, μ2, second central moment.  相似文献   

15.
A kinetic mathematical model has been applied to investigate for the first time the effects of Pd particle size on the rates of oxygen back-spillover and CO oxidation during Oxygen Storage Capacity (OSC) measurements under dynamic conditions over Pd/CeO2 catalysts in the 500–700 °C range. The dependence of the intrinsic rate constant k1 of the CO oxidation reaction on PdO, and that of k 2 app of the oxygen back-spillover from ceria to Pd/PdO on the palladium particle size was estimated by performing curve-fitting of the experimental CO and CO2 pulse transient responses obtained. Activation energies of 8.0, 9.5 and 21.1 kJ/mol were calculated for the Eley–Rideal step of CO oxidation for the 1.3, 1.8 and 16.4 nm Pd particles, respectively, supported on CeO2. The transient rates of CO oxidation and oxygen back-spillover were found to decrease with increasing Pd particle size.  相似文献   

16.
The absorption rates of NO into aqueous solutions of NaClO2 blended with Na2CO3 were investigated in a stirred reactor. The experimental results showed that the reaction rate could be expressed as . An equation for a second‐order reaction rate constant between NO and NaClO2, kmn = 5.79 · 1015 exp(–5557.26/T), was obtained. The addition of Na2CO3 into the solution of NaClO2 decreased the reaction rate constant. The optimal absorption conditions involved a NO concentration of 540 ppm, NaClO2 concentration of 50 mol m–3, Na2CO3 concentration of 10 mol m–3, temperature of 333 K and O2 concentration of 4 %, which were determined by an orthogonal experiment. Under these optimal conditions, it was possible for the absorption rate to reach up to 1.9271 · 10–5 mol/(s m2).  相似文献   

17.
《分离科学与技术》2012,47(12):2645-2660
Abstract

Three sulfa drugs (sulfathiazole, sulfamethizole, and sulfabenzamide) were crystallized using carbon dioxide and water as antisolvents, and the effects of the type of solvent, the crystallization temperature, and the antisolvent injection rate were investigated. Sulfathiazole crystallized in granulate form from acetone, but it was crystallized in acicular form from methanol. Sulfamethizole was crystallized in tabulate form from acetone and as plates from DMF. Sulfabenzamide was precipitated in the form of prisms from acetone and of aciculates from ethyl acetate. As the crystallization temperature increased from 30 to 50°C, the average particle size increased from 6.5 to 10.5 µm for sulfathiazole, 29.5 to 53.1 µm for sulfamethizole, and 33.0 to 59.8 µm for sulfabenzamide. The crystal habit tended to become more needle‐like as the antisolvent injection rate increased. Larger particles were produced when the antisolvent was changed from carbon dioxide to water.  相似文献   

18.
Granular activated carbons were obtained from grape seeds by pyrolysis at 600°C and subsequent physical activation with CO2 (750–900°C, 1–3 h, 25–74% burn-off). The carbon and ash content increased during the activation, reaching values of 79.0% and 11.4%, respectively. Essentially microporous materials with BET surface areas between 380 and 714 m2/g were obtained. The performance of the activated carbon in the adsorption of diuron in aqueous phase was studied within the 15–45°C temperature range. Equilibrium data showed that the maximum uptake increased with temperature from 120 to 470 µmol/g, also evidencing some dependence of the adsorption mechanism on temperature. Data were fitted to five isotherm models [Langmuir, Freundlich, Dubinin–Radushkevich, BET, and GAB (Guggenheim, Anderson, and de Boer)]. Kinetic data were analyzed using first- and second-order rate equations and intraparticle diffusion model. The second-order rate constant values obtained (2.8–13.5 × 10?3 g/µmol min) showed that the hollow core morphology of the material favors the adsorption kinetics.  相似文献   

19.
Hydrodesulfurization of Athabasca (Syncrude) fluid coke was studied for particle sizes of ?74 + 53, ?53 + 44 and ?44 μm using a quartz reactor. Five flow rates of hydrogen from 1.2 × 10?6 to 2.5 × 10?6 m3/s were investigated with from 0.45 to 1.0 g of coke. Gas production — time profiles for H2S and CH4 were obtained at temperatures from 973 to 1073 K for each particle size range. Desulfurization rates were functions of particle size and temperature. Results agree with predictions of the shrinking core model, the rate being controlled initially by the gas film and chemical reaction resistances followed by control due to diffusion of hydrogen through the increasing ash layer. Below 998 K, the apparent activation energy was determined to be 293 kJ/mol · K, while at temperatures between 998 and 1073 K it was 29 kJ/mol · K.  相似文献   

20.
《分离科学与技术》2012,47(4):739-768
Abstract

Experimental results for the separation of S,R‐α‐Tetralol enantiomers obtained on preparative columns packed with particle size 20 µm of chiral adsorbent CHIRALPAK AD are presented. The total porosity was measured by using the non‐retained compound 1,3,5‐Tri‐tert‐butylbenzene and was 0.61. The permeability of the bed packed with CHIRALPAK AD was calculated as 4.4×10?13 m2. The efficiency of columns was characterized by the height equivalent to a theoretical plate (HETP) and a linear dependency has been found over tested flow rates. The HETP of S‐α‐Tetralol and R‐α‐Tetralol calculated at the flow rate 5.0 cm3/min were 320 µm and 340 µm, respectively. Thermodynamic adsorption parameters enthalpy, ΔH and entropy, ΔS, have been calculated from van't Hoff plot. Equilibrium and kinetics of adsorption of single enantiomers and racemic mixture of α‐Tetralol on CHIRALPAK AD were evaluated as well. The parameters for multicomponent isotherm linear‐Langmuir model are presented. The breakthrough curves of α‐Tetralol enantiomers are simulated with a mathematical model that accounts for axial dispersion and linear driving force for the intraparticle mass transfer.  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