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.
BACKGROUND: Simultaneous removal of sulfur, nitrogen and carbon compounds from wastewaters is a commercially important biological process. The objective was to evaluate the influence of the CH3COO?/NO3? molar ratio on the sulfide oxidation process using an inverse fluidized bed reactor (IFBR). RESULTS: Three molar ratios of CH3COO?/NO3? (0.85, 0.72 and 0.62) with a constant S2?/NO3? molar ratio of 0.13 were evaluated. At a CH3COO?/NO3? molar ratio of 0.85, the nitrate, acetate and sulfide removal efficiencies were approximately 100%. The N2 yield (g N2 g?1 NO3?‐N consumed) was 0.81. Acetate was mineralized, resulting in a yield of 0.65 g inorganic‐C g?1 CH3COO?‐C consumed. Sulfide was partially oxidized to S0, and 71% of the S2? consumed was recovered as elemental sulfur by a settler installed in the IFBR. At a CH3COO?/NO3? molar ratio of 0.72, the efficiencies of nitrate, acetate and sulfide consumption were of 100%, with N2 and inorganic‐C yields of 0.84 and 0.69, respectively. The sulfide was recovered as sulfate instead of S0, with a yield of 0.92 g SO42?‐S g?1 S2? consumed. CONCLUSIONS: The CH3COO?/NO3? molar ratio was shown to be an important parameter that can be used to control the fate of sulfide oxidation to either S0 or sulfate. In this study, the potential of denitrification for the simultaneous removal of organic matter, sulfide and nitrate from wastewaters was demonstrated, obtaining CO2, S0 and N2 as the major end products. Copyright © 2008 Society of Chemical Industry  相似文献   

2.
Silver ions are bound on polystyrene containing ? CH2Cl groups with a distribution coefficient of about 100 ml/g, but they are not bound on polystyrene itself. This indicates direct interaction between Ag+ ions and the chlorine atoms of the ? CH2Cl groups in the sense of formation of a 1:1 complex. In presence of 0.1 mol/l of ammonium ions somewhat smaller distribution coefficients are found in the range from pH 4 to pH 6. In presence of 10?3 mol/l of Na2S2O3 the distribution coefficients are only about 6 ml/g. From solutions, however, containing 1 mol/l of NaCl or 0.1 mol/l of Na2S2O3 or 10?3 mol/l of NaCN, silver is not sorbed in measurable amounts, because complexation in solution predominates.  相似文献   

3.
A new lauryl amidopropyl trimethyl ammonium methyl carbonate with the formula CH3(CH2)10CONH(CH2)3N+(CH3)3CH3CO3 ? was synthesized via a high pressure process with tertiary amines and dimethyl carbonate, and its chemical structure was confirmed using 1H-NMR spectra, mass spectral fragmentation, and FTIR spectroscopic analysis. In addition, several quaternary ammonium salts with new counterions X? (X?=HCO3 ?, HCOO?, CH3COO?, CH3CH(OH)COO?) were also synthesized by the ion exchange reaction of methyl carbonate quaternary ammoniums with corresponding acids. The surface activities of these compounds were measured, including surface tension (??), critical micelle concentration and minimum surface area (A min) at 25?°C. Adsorption and micellization free energies of these quaternary ammonium salts in their solutions showed a good tendency towards adsorption at interfaces. The antimicrobial activities are reported for the first time against representative bacteria and fungi for lauryl amidopropyl trimethyl ammoniums. It was found that the antimicrobial potency was Gram-positive bacteria?>?fungi?>?Gram-negative bacteria.  相似文献   

4.
Ozone/hydrogen peroxide batch treatment was utilized to study the degradation of the steroidal hormone estrone (E1). The competition kinetics method was used to determine the rate constants of reaction for direct ozone and E1, and for hydroxyl radicals and E1 at three pH levels (4, 7, and 8.5), three different molar O3/H2O2 ratios (1:2, 2:1, and 4:1) and a temperature about 20°C. The average second-order rate constants for direct ozone-E1 reaction were determined as 6.2?×?103?±?3.2?×?103 M?1s?1, 9.4?×?105?±?2.7?×?105 M?1s?1, and 2.1?×?107?±?3.1?×?106 M?1s?1 at pH 4, 7, and 8.5, respectively. It was found that pH had the greatest influence on the reaction rate, whereas O3/H2O2 ratio was found to be slightly statistically significant. For the hydroxyl radical-E1 reaction, apparent rate constants ranged from 1.1?×?1010 M?1s?1 to 7.0?×?1010 M?1s?1 with an average value of 2.6?×?1010 M?1s?1. Overall, O3/H2O2 is shown to be an effective treatment for E1.  相似文献   

5.
One way to improve the rate of hydrogen production from water-splitting reactions is by the separation of photogenerated carriers. This separation process can be achieved with narrow bandgap semiconductors. ZnSe has a 2.7?eV bandgap, but its photocatalytic activity is very low due to a high recombination rate of the photogenerated carriers. Therefore, a combination of Bi2S3 and ZnSe may potentially produce a visible-light-active photocatalyst, utilizing bandgap engineering and the p-n junction effect. ZnSe, Bi2S3 and Bi2S3-ZnSe nanocomposites were prepared by a hydrothermal method. Bi2S3 at different weight percentages (3–15?wt%) was decorated with ZnSe nanoparticles. The hydrogen evolution reaction was conducted in the investigation of ZnSe, Bi2S3 and Bi2S3-ZnSe photocatalytic efficiency. The results demonstrate that photocatalytic efficiency was highly affected by the Bi2S3 weight percent. The optimal weight percent for Bi2S3 was 15?wt%, at which the rate of hydrogen evolution was 2600?μmol?g?1 h?1 within 240?min in the presence of 1.2?g/L photocatalyst.  相似文献   

6.
The air‐aging process at 120°C and the thermooxidative degradation of peroxide prevulcanized natural rubber latex (PPVL) film were studied with FTIR and thermal gravity (TG) and differential thermal gravity (DTG) analysis, respectively. The result of FTIR shows that the ? OH and ? COOH absorption of the rubber molecules at IR spectrum 3600–3200 cm?1, the ? C?O absorption at 1708 cm?1, and the ? C? OH absorption of alcohol at 1105 and 1060 cm?1 increased continuously with extension of the aging time, but the ? CH3 absorption of saturated hydrocarbon at 2966 and 2868 cm?1, the ? CH3 absorption at 1447 and 1378 cm?1, and the C?C absorption at 835 cm?1 decreased gradually. The result of TG‐DTG shows that the thermal degradation reaction of PPVL film in air atmosphere is a two‐stage reaction. The reaction order (n) of the first stage of thermooxidation reaction is 1.5; the activation energy of reaction (E) increases linearly with the increment of the heating rate, and the apparent activation energy (E0) is 191.6 kJ mol?1. The temperature at 5% weight loss (T0.05), the temperature at maximum rate of weight loss (Tp), and the temperature at final weight loss (Tf) in the first stage of degradation reaction move toward the high temperature side as the heating rate quickened. The weight loss rate increases significantly with increment of heating rate; the correlation between the weight loss rate (αp) of DTG peak and the heating rate is not obvious. The weight loss rate in the first stage (αf1) rises as the heating rate increases. The final weight loss rate in second stage (αf2) has no reference to heating rate; the weight loss rate of the rubber film is 99.9% at that time. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 92: 3196–3200, 2004  相似文献   

7.
Block copolymers having a pendant trichlorogermyl group as a part of polyamide segment? (CO? R′? CO? NH? Ar? NH? )xCO? R′? CO? and polydimethylsiloxane of general formula [(? CO? R′? CO? HN? Ar? NH)x? CO? R′? CO? NH(CH2)3SiO(CH3)2 ((CH3)2SiO)ySi(CH3)2(CH2)3 NH? ]n (where R′ = CH2CH(GeCl3), CH(CH3)CH(GeCl3), CH(GeCl3)CH(CH3); Ar = C6H4, (? C6H3? CH3)2, (? C6H3? OCH3)2, 2,5‐(CH3)2? C6H2, C6H4? O? C6H4) were prepared by a polycondensation reaction and characterized using CHN and Ge analysis, Fourier transform infrared (FTIR) and 1H NMR spectroscopy, thermogravimetric analysis (TGA) and molecular weight determination. They have a lamellar structure with weight‐average molecular weight in the range 1.21 × 105–4.79 × 105 g mol?1. These copolymers display two glass transition temperatures and have an average decomposition temperature of 489 °C. TGA, FTIR and gas chromatography/mass spectrometry studies indicate that degradation of these block copolymers results in carbon monoxide, oligomeric siloxanes and polyamide fragments. They are thermally stable due to the hydrogen bonded interlinked chains of polyamide, while they absorb water due to the presence of Ge? Cl bonding. Copyright © 2010 Society of Chemical Industry  相似文献   

8.
A mechanism is developed for the initiated nonbranched-chain formation of ethylene glycol in methanol-formaldehyde solutions at formaldehyde concentrations of 0.1–3.1 mol dm?3 and temperatures of 373–473 K. At a formaldehyde concentration of 1.4 mol dm?3 and T = 473 K, the radiation-chemical yield of ethylene glycol is 139 molecules per 100 eV. The effective activation energy of ethylene glycol formation is 25 ± 3 kJ mol?1. The quasi-steady-state treatment of the reaction network suggested here led to a rate equation accounting for the nonmonotonic dependence of the ethylene glycol formation rate on the concentration of the free (unsolvated) form of dissolved formaldehyde. It is demonstrated that the peak in this dependence is due to the competition between methanol and CH2=O for reacting with the adduct radical HOCH2CH2O?.  相似文献   

9.
In this work, Co3S4/CNTs three-dimensional hierarchical microstructure has been synthesized through a facile sulfidation method with Co3O4/CNTs as precursor. Particularly, by controlling hydrothermal time, the optimized Co3S4/CNTs-9 composite consists of ultrathin and rich mesoporous Co3S4 nanosheets (<?10?nm) with CNTs firmly intertwined inside microstructures. Benefitting from the porous and robust secondary-structure, as well as the intertwined CNTs among ultrathin active materials Co3S4 nanosheets, Co3S4/CNTs-9 electrode provides fast efficient ways of ion and electron transportation. As a result, it exhibits an excellent rate performance. Charge capacity retention at 1?A?g?1 is up to 80% compared with 0.1?A?g?1. The enhanced extrinsic pseudocapacitive contribution is confirmed at higher rate through the cyclic voltammetry analysis at different scan rates of Co3S4/CNTs-9 electrode. Importantly, Co3S4 electrode also exhibits a high initial coulombic efficiency of 82.89%, which suggests that Co3S4 is a promising anode candidate for sodium-ion batteries.  相似文献   

10.
The catalysts such as Al2O3/ZrO2 with 2–10?wt% of Al2O3 were coated on honeycomb monoliths by dip-and-dry technique. These catalysts were also prepared in their powder form. All the catalysts (honeycomb and powder form) were characterized for their surface acidity, crystallinity, functionality, elemental analysis, and morphology. The catalytic activity of all the catalysts was performed in the transesterification of methyl salicylate with benzyl alcohol to synthesize benzyl salicylate. Reaction conditions like reaction time, reaction temperature, and the molar ratio of the reactants were varied to obtain the highest yield of benzyl salicylate. The 6% Al2O3/ZrO2 coated on honeycomb exhibited the highest conversion of methyl salicylate at 383?K in 60?min. Kinetic studies were conducted to determine the energy of activation and temperature coefficient. The rate constants in the case of 6AZ (HCM) was found to be 5.0?×?10?3?min?1 (373?K); 6.4?×?10?3?min?1 (383?K) and 2.2?×?10?3?min?1 (373?K); 3.2?×?10?3?min?1 (383?K) in the case of 6AZ (PF) catalyst, while the energy of activation (Ea) values were found to be 35.12 and 39.93 kJ mol?1 for 6AZ (HCM) and 6AZ (PF), respectively. The reactant preadsorption study discloses that the transesterification follows the Eley–Rideal mechanism. Reactivation and recyclability of the catalysts were also examined and the results clearly indicate that Al2O3/ZrO2 coated on the honeycomb is efficient and green catalytic system.  相似文献   

11.
Soil salinization adversely affects sustainable land use and limitation of greenhouse gas emission. Methane (CH4) uptake and the specific activity of methanotrophs in three saline–alkaline soils—S1, electrical conductivity (EC) 4.80?dS?m?1; S2, EC 2.60?dS?m?1; and S3, EC 0.74?dS?m?1—were observed and measured across crop phenological development in the Hetao Irrigation District of Inner Mongolia, China. There were significant differences in CH4 uptake between the three soil types. The cumulative uptake of CH4 was 97.97 mg m?2, 109.49 mg m?2, and 150.0 mg m?2 in S1, S2, and S3, respectively. Cumulative CH4 uptake was 35%, 35%, and 53% lower in S1 than in S3, and was 27%, 28%, and 19% lower in S2 than in S3 in 2014, 2015, and 2016, respectively. Differences in CH4 uptake were driven by the different specific activities of the methanotrophs in the three soils, of which the key controlling factor was soil EC. The findings demonstrate that saline–alkaline soils with high EC led to low CH4 uptake and thereby significantly increased the total greenhouse effect of CH4.  相似文献   

12.
BACKGROUND: Loss of the active CuCl component occurs during acetylene dimerisation to monovinylacetylene (MVA) catalysed by an anhydrous catalyst with the formation of a dark red precipitate. Acidic species can reduce the loss of CuCl but have an unfavorable influence on acetylene dimerisation. This study aims to determine the precipitate composition and regulate the acidity of the catalyst to find a balance between reaction rate, MVA selectivity and catalyst life. RESULT: The precipitate composition was 2CuCl·3C2H2·1/3CH3CH2NH2·1/7C3H7NO, formed by the combination of DMF, CH3CH2NH2, C2H2 and the [Cu]‐acetylene π‐complex, which is an intermediate in the reaction. From an overall consideration of the loss of CuCl, conversion of acetylene, and selectivity of MVA, the reaction temperature and acetylene space velocity were optimized at 65 °C and 200 h?1, respectively. The introduction of HCl into the catalyst with a rate of 3.2 h?1 could reduce CuCl loss by 73.5%, whereas conversion of acetylene was only lowered by 9.0%. CONCLUSION: Acidity regulation of the anhydrous catalyst by optimising the reaction temperature, acetylene space velocity, and rate of addition of HCl shows little negative effect on acetylene conversion and selectivity to MVA but can reduce CuCl loss significantly. © 2012 Society of Chemical Industry  相似文献   

13.
Oxotris(dimethyl dithiocarbamato) vanadium(V) [VO(S2CN(CH3)2)3] sensitizes the polymerization of styrene when irradiated by light of λ = 365 nm at 25°C. Under the experimental conditions employed, no retardation occurs, and the rate of initiation is independent of monomer concentration. The mean values of the quantum yield of iniiation (?i) and polymerization (?o) are 2.85 × 10?3 and 6.72 respectively. Spectroscopic analysis shows that initiation occurs predominatly through scission of the N,N-dimethyl dithiocarbamate ligand (—SC(S)N(CH3)2) with reduction of vanadium(V) to (IV), and VO (S2CN(CH3)2)2 is the final photolytic product. A reaction mechanism is proposed based on an intramolecular photoredox reaction which leads to the primary formation of SC(S)N(CH3)2 radicals and a vanadium(IV) chelate complex. The rellevant kinetic parameters are evaluated. The polystyrene produced shows a photoactivity when irradiated with UV-light.  相似文献   

14.
This paper presents the luminescence properties and potential of red SrAl12O19:Mn4+ (SAO:Mn4+) phosphor for optical thermometry application. The SAO crystal consisted of a spinel block along with two mirror-like blocks. The Al3+/Sr2+ molar ratio of the precursor solution affected the crystalline-phases, morphology, and photoluminescence of the phosphor. The addition of flux H3BO3 promoted the growth of hexagonal-nanoflakes and enhanced the external quantum efficiency of phosphor 2.6-fold. The absolute sensitivity Sa and relative sensitivity Sr of SAO:Mn4+ showed a linear function of the temperature. The value of Sa was 4.17?×?10?3 K?1, and the maximum Sr was 2.70?×?10?3 K?1 at 393?K. A stable emission color was observed even with a change in temperature and a bright red light was seen in both daylight and a darkroom.  相似文献   

15.
The reaction rate of enolization of p-chloroacetophenone has been studied in the presence of amino acids, namely, β-alanine, DL-alanine, L-alanine, and glycine. Maximum reaction rate was observed in the case of β-alanine at 22M concentration. The effects of parameters such as effect of ketone concentration, effect of dielectric constant, effect of catalysts, etc., have been found to exert a significant effect on the reaction rate. The effect of temperature was studied in the range of 313–328?K, and several thermodynamic parameters such as entropy (ΔS), enthalpy (ΔH), energy of activation (ΔEa), and Gibbs free energy (ΔF) were found to be ?10.32?e.u., 17.87?cal?mol?1, 19.24?kcal?mol?1, and 21.20?cal?mol?1, respectively. The reaction rate increased from 1.9 to 5.8?min?1 on increasing the percentage of dimethylformamide from 10 to 50% (v/v).  相似文献   

16.
Electrochemical reduction of acetone in aqueous H2SO4 solution was studied on Hg electrode by galvanostatic, potentiodynamic and polarographic methods. Electrocapillary curves show that acetone molecule adsorbs to a monolayer at concentrations larger than 0·5 M in a potential range of 0 > V > ? 1·0 V (nhe). It orients with the positive end of the dipole toward electrode surface and extends an attractive interaction to its neighbours.Kinetics oberved by galvanostatic pulse method and polarography leads the following conclusions; (1) adsorbed acetone molecule undergoes elecrochemical reduction at potentials more negative than ?1·0 V (nhe), (2) reaction rate is the first and second order with respect to the amount of adsorbed acetone and proton activity, and (3) one electron transfer step determines the reaction rate (Tafel slope is 0·12 V). Reaction intermediate is concluded from the potentiodynamic study to be isopropanol radical, (CH3)2??OH, whose amount is proportional to that of adsorbed acetone and to aH+. From the above results, the rate of the electrochemical reduction of acetone is H2SO4 solution is concluded as determined by the step, (CH3)2?OH(a) + H+ + e? → (CH3)2CHOH.  相似文献   

17.
PVC, which was polymerized at atmospheric pressure (so called U-PVC) contains relatively high concentrations of defects contrary to normal PVC. The number of chain scissions in U-PVC determined by ozonolytic cleavage resulted in values between 0.026 and 0.058 per 100 monomer units (100 VC). The determination of allylic and tertiary chlorine was done by selective reaction of U-PVC with phenol and NMR-spectroscopic investigations of the phenolized polymers. The average ‘labile chlorine’ content amounts to 0.65/100 VC. Hydroxyl radicals formed during the decomposition of the initiator (K2S2O8) resulted in alcoholic endgroups in U-PVC, which were detectable in the IR-spectrum at 3580 cm?1. The termination with hydroxyradicals also led to structures at the chain ends changing into ß-chloraldehyde groups accompanied by HCl-elimination. The corresponding signal in the IR-spectrum appeared at 1720 cm?1. U-PVC raw material contained about two branch points per 100 VC. The CCl4 extracts of the same polymers revealed the ten-fold content of branching. The olefinic structures ? CH?CH? CHCl? and ? CHCl? CH?CH2 were determined by NMR-spectroscopy. The concentrations of each ranged from 0.25 to 0.3/100 VC. A typical double bond for U-PVC at the chain ends represented the structure ? CH?CH? CH2Cl, which was preferably present in the low molecular weight material.  相似文献   

18.
A sensitive and convenient method is described for the determination of trace amounts of S2? ion in aqueous solution based on the classical Belousov-Zhabotinskii (B-Z) oscillating reaction. The results showed that the change in the oscillating period was linearly proportional to the concentration of the S2? ion with a wider range from 7.94×10?8 to 3.16×10?5 mol·L?1. Various influencing factors are studied and a possible reaction mechanism is also discussed by means of the well-known FKN model. A comparison with the other methods used to determine the S2? ion is given, too.  相似文献   

19.
Linear asymmetrical poly(propylene oxide) was synthesized through four‐step reactions: selective benzylation, alcohol exchange reaction, propylene oxide anionic polymerization, debenzylation. One terminal of the asymmetrical polymer chains is alcohol hydroxyl and the other is phenol hydroxyl. It was characterized with infrared (IR) and 1H Nuclear Magnetic Resonance (1H‐NMR). Peaks at 1.11, 3.38, and 3.53 ppm were attributed to side groups (? OCH2CH(CH3)? ), backbone units (? OCH2CH(CH3)? ) and (? OCH2CH(CH3)? ) of poly(propylene oxide), respectively. Molecular weight and molecular weight distribution were measured with 1H‐NMR and laser light scattering (LLS), which showed that the linear asymmetrical poly(propylene oxide) was mono‐disperse (PDI = 1.02–1.07). Then, its carbamate reaction with phenyl isocyanate was studied; the reaction rate constants for phenol hydroxyl and alcohol hydroxyl of poly(propylene oxide) were k1 = 0.209 mol L?1 min?1 and k2 = 0.051 mol L?1 min?1. There was a great reactivity difference for two types of hydroxyls in asymmetrical poly(propylene oxide), contrasting to the single carbamate reaction rate constant of symmetrical poly(propylene oxide) (k3 = 0.049 mol L?1 min?1). © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

20.
Enhancement of the air‐water interfacial transfer of I2 by reaction with I? was investigated in order to evaluate a mechanistic model. Separate systems for evaporation and absorption of I2 were studied. The overall mass transfer coefficient increased from 2 × 10?4 to 2.5 × 10?3 cm/s as I? concentration increased from 10?4 to 10?1 M. At I? concentrations greater than 10?1 M, the interfacial transfer of I2 was gas‐side limited. The model gave a good representation of these results.  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