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 406 毫秒
1.
The adsorption of chloridazon (5‐amino‐4‐chloro‐2‐phenylpyridazin‐3(2H)‐one) on a new type of material formed by kerolite/stevensite bearing lithofacies and on a bentonite, desiccated at 110 °C from aqueous solution at 25 °C has been studied by using batch experiments. In addition, column experiments were carried out with these samples using aqueous solutions of chloridazon at a concentration (C) of 45 × 10−4 cmol dm−3. The experimental data points have been fitted to the Langmuir equation to calculate the adsorption capacities (Xm). Values for Xm ranged from 0.072 cmol kg−1 for bentonite up to 1.30 cmol kg−1 for kerolite. The removal efficiency, R, ranged from 17.1% for bentonite up to 85.1% for kerolite. The batch experiments show that the kerolite is more effective than bentonite in adsorbing chloridazon. The column experiments show that kerolite might be reasonably used in removing chloridazon, the data indicating that a readily available and inexpensive Spanish kerolite can be employed as a filter for contaminated waters with chloridazon, so controlling its release to the environment. © 2000 Society of Chemical Industry  相似文献   

2.
Sepiolite is a hydrated magnesium silicate clay with a fibrous structure and binder properties. To calculate the potential use of sepiolite in removing atrazine [2‐chloro‐4‐ethylamino‐6‐isopropilamino‐1,3,5,‐triazine], isoproturon [3‐(4‐isopropylphenyl)‐1,1‐dimethylurea] and imidacloprid [1‐(6‐chloro‐3‐pyridinylmethyl)‐N‐nitroimidazolin‐2‐ylideneamine] from water, the adsorption of atrazine, isoproturon and imidacloprid on sepiolite desiccated at 110 °C from aqueous solution at 25 °C has been studied by using batch experiments. In addition, column experiments were carried out with the sepiolite sample using aqueous solutions of atrazine, isoproturon and imidacloprid at a concentration of 20.0 × 10−4 cmol dm−3. The experimental data points have been fitted to the Langmuir equation to calculate the adsorption capacities (Xm). Values for Xm ranged from 2.70 × 10−1 cmol kg−1 for isoproturon up to 3.97 × 10−1 cmol kg−1 for atrazine. The removal efficiency (R) ranged from 36.7% for isoproturon up to 74.3% for atrazine. The batch experiments show that the sepiolite is more effective in adsorbing atrazine than imidacloprid and isoproturon. The column experiments show that sepiolite might be reasonably used in removing atrazine, the column efficiency being 46%. The data indicate that a readily available and inexpensive Spanish sepiolite can be employed as a filter for contaminated waters with these pesticides, controlling their release to the environment. © 1999 Society of Chemical Industry  相似文献   

3.
Increasing public concerns and EPA air regulations in non‐attainment zones necessitate the remediation of volatile organic compounds (VOCs) generated in the poultry‐rendering industry. Wet scrubbers using chlorine dioxide (ClO2) have low overall removal efficiencies due to lack of reactivity with aldehydes. Contrary to wet scrubbers, a biofilter system successfully treated the aldehyde fraction, based on GC/MS analysis of inlet and outlet streams. Total VOC removal efficiencies ranged from 40 to 100% for the biofilter, kinetic analysis indicated that the overall removal capacity approached 25 g m−3 h−1, and aldehyde removal efficiency was significantly higher compared with chemical wet scrubbers. Process temperatures monitored in critical unit operations upstream from the biofilter varied significantly during operation, rising as much as 30 °C within a few minutes. However, the outlet air temperature of a high intensity scrubber remained relatively constant at 40 °C, although the inlet air temperature fluctuated from 50 to 65 °C during monitoring. These data suggest a hybrid process combining a wet scrubber and biofilter in series could be used to improve overall VOC removal efficiencies and process stability. Copyright © 2005 Society of Chemical Industry  相似文献   

4.
The sorption of linuron on bentonite desiccated at 110°C untreated, and acid‐treated with H2SO4 solutions over a concentration range between 0.25 M and 1.00 M from aqueous solution at 25°C has been studied by using batch experiments. In addition, column experiments were carried out with the bentonite sample treated with the 1.00 M H 2SO4 solution [B‐A(1.00)] by using two aqueous solutions of linuron of different concentrations (C=4.97 mg dm−3 and C=7.63 mg dm−3 ). The experimental data points have been fitted to the Langmuir equation in order to calculate the sorption capacities (Xm) of the samples; Xm values range from 0.02 g kg−1 for the untreated bentonite [B‐N] up to 0.20 g kg−1 for the sample acid‐treated with the 1.00 M H2 SO4 solution. The removal efficiency (R ) has also been calculated; R values ranging from 15.86% for the [B‐N] sample up to 41.54% for [B‐A(1.00)]. The batch experiments show that the acid‐treated bentonite is more effective than the natural bentonite in relation to sorption of linuron. The column experiments show that the B‐A(1.00) sample might be reasonably used in removing linuron, the column efficiency increasing from 61.8% for the C=7.63 mg dm−3 aqueous solution of linuron up to 77.6% for the C=4.97 mg dm−3 one. © 1999 Society of Chemical Industry  相似文献   

5.
The effect of a decrease in operating temperature on the performance of two 10 dm3 anaerobic baffled reactors (ABR) was examined in terms of steady state chemical oxygen demand (COD) removal efficiency. To minimise variations, and have a totally biodegradable feed, a synthetic carbohydrate (sucrose)–protein (meat extract) substrate was used. The reactors were operated at 20 h hydraulic retention time (HRT), 4 g dm−3 COD, and 35°C as a base-line condition. Because of their different histories, the reactors responded differently to a decrease in operating temperature to 25°C. Reactor 1 remained stable at 97% COD removal, whereas Reactor 2 decreased to 93% removal, but rose to 97% after adding an effluent recycle of 0·25. At 15°C, the efficiency of Reactor 1 dropped to 75%, while the removal of Reactor 2 declined to 83%, and no improvement in efficiency occurred with an effluent recycle at 0·25. At 25°C, the decreased rate of catabolism of the slow-growing syntrophs and methanogens resulted in a shift of the volatile fatty acids (VFA) peak to the second compartment. However, the biomass present in the reactor prevented VFAs breaking through in the effluent. Nevertheless, at 15°C VFAs were present in the effluent, perhaps due to the lower rates of metabolism and an increase in the Ks for VFAs. Finally, at 15°C part of the increase in the effluent COD was due to the enhanced production of soluble microbial products (SMP), or a decrease in their metabolism, with these compounds constituting some 10% of the inlet COD. © 1997 SCI.  相似文献   

6.
A novel trickling fibrous-bed bioreactor was developed for biofiltration to remove pollutants present in contaminated air. Air containing benzene as the sole carbon source was effectively treated with a coculture of Pseudomonas putida and Pseudomonas fluorescens immobilized in the trickling biofilter, which was wetted with a liquid medium containing only inorganic mineral salts. When the inlet benzene concentration (Cgi) was 0·37 g m−3, the benzene removal efficiency in the biofilter was greater than 90% at an empty bed retention time (EBRT) of 8 min or a superficial air flow rate of 1·8 m3 m−2 h−1. In general, the removal efficiency decreased but the elimination capacity of the biofilter increased with increasing the inlet benzene concentration and the air (feed) flow rate. It was also found that the removal efficiency decreased but the elimination capacity increased with an increase in the loading capacity, which is equal to the inlet concentration divided by EBRT. The maximum elimination capacity achieved in this study was ∽11·5 g m−3 h−1 when the inlet benzene concentration was 1·7 g m−3 and the superficial air flow rate was 3·62 m3 m−2 h−1. A simple mathematical model based on the first-order reaction kinetics was developed to simulate the biofiltration performance. The apparent first order parameter Kl in this model was found to be linearly related to the inlet benzene concentration (Kl=4·64−1·38 Cgi). The model can be used to predict the benzene removal efficiency and elimination capacity of the biofilter for benzene loading capacity up to ∽30 g m−3 h−1. Using this model, the maximum elimination capacity for the biofilter was estimated to be 12·3 g m−3 h−1, and the critical loading capacity was found to be 14 g m−3 h−1. The biofilter had a fast response to process condition changes and was stable for long-term operation; no degeneration or clogging of the biofilter was encountered during the 3-month period studied. The biofilter also had a relatively low pressure drop of 750 Pa m−1 at a high superficial air flow rate of 7·21 m3 m−2 h−1, indicating a good potential for further scale up for industrial applications. © 1998 Society of Chemical Industry  相似文献   

7.
The sorption of 1,1′-dimethyl-4,4′bipyridilium dichloride (paraquat) on bentonite desiccated at 110°C untreated, and acid-treated with H2SO4 solutions over a concentration range between 0·25 M and 1·00 M , from aqueous solution at 30°C has been studied by using batch experiments. In addition, column experiments were carried out with the bentonite sample treated with the 1·00 M H2SO4 solution [B-A(1·00)] by using two aqueous solutions of paraquat of different concentrations (C = 29·40 mg dm−3 and C = 65·38 mg dm−3). The experimental data points have been fitted to the Langmuir equation in order to calculate the sorption capacities (Xm) of the samples; Xm values range from 1·35×105 mg kg−1 for the sample acid-treated with 0·375 M H2SO4 [B-A(0·375)] up to 1·96×105 mg kg−1 for the untreated bentonite [B-N]. The removal efficiency (R) has also been calculated; R values ranging from 44·61% for the [B-A(0·375)] sample up to 67·23% for B-N. The batch experiments show that the natural bentonite is more effective than the acid-treated bentonite in relation to sorption of paraquat. The column experiments show that the B-A(1·00) sample might be reasonably used in removing paraquat, the column efficiency increasing from 37·55% for the C = 65·38 mg dm−3 aqueous solution of paraquat up to 66·58% for the C = 29·40 mg dm−3 one. © 1997 SCI.  相似文献   

8.
The diffusion coefficient at zero penetrant concentration D0 of dichloromethane, chloroform, carbon tetrachloride, cyclohexane, benzene, o-xylene, m-xylene, and p-xylene, and n-hexane in LDPE were measured at 25°C, using the desorption method. The D0 values obtained in this way are correlated with the size, shape, and chemical nature of the penetrant molecules. The temperature dependence of the diffusion coefficients of toluene and n-hexane in LDPE are also reported in the limited temperature range of 25–45°C. It indicates that, in spite of a size larger than that of toluene, n-hexane has a lower activation energy of diffusion.  相似文献   

9.
《分离科学与技术》2012,47(11):1699-1708
Hydroxy terminated polybutadiene (HTPB)-based polyurethaneurea membranes with and without cross-linkage were synthesized and first used as membrane material to separate p-/o-xylene mixtures by pervaporation. Compared with HTPB-PU (without cross-linkage) membranes, HTPB-DVB-PU (cross-linked HTPB-PU with divinyl benzene) membranes demonstrated a lower degree of swelling in xylene isomer solutions and noticeable improved separation factor of p-/o-xylene. On the other hand, the amount of p-xylene adsorbed in HTPB-DVB-PU membranes increased significantly rather than that of o-xylene. While the separation factor of p-/o-xylene increased but the total flux decreased with increasing DVB content, which can be ascribed to the improved chemical structure and more homogeneous chain structures of the HTPB-DVB-PU membranes. The p-xylene normalized permeation rate and separation factor of p-/o-xylene of HTPB-DVB-PU membrane reached 2.70 kgµm/m2h and 2.23, respectively, at a feed concentration of 10 wt% p-xylene at 30°C.  相似文献   

10.
BACKGROUND: In this work, the feasibility of biodegradation and the removal performance of sole and mixed odorous vapors, such as dimethyl disulfide (DMDS), methyl phenyl sulfide (MPS), and ethanethiol (EtSH) in an EtSH‐acclimated biotrickling filter seeded with commercially available B350 microorganisms, were investigated. RESULTS: Removal efficiencies (REs) for DMDS as a sole substrate were evaluated under different inlet concentrations and empty bed residence times (EBRT), 100% RE was achieved at concentration below 0.4 g m?3 at EBRT 110 s. In addition, 100% RE was obtained for binary EtSH and DMDS (3:2) at the same EBRT. According to the Michaelis–Menten type kinetic equation, the maximum removal rates (Vmax) were calculated as 28.7 and 13.9 g m?3 h?1 for DMDS and MPS as sole substrate, respectively, while Vmax was 22.1 and 10.1 g m?3 h?1 for DMDS and MPS in the presence of EtSH and EtSH‐DMDS mixture, respectively. After 5 and 20 days starvation, the re‐acclimation times were only 2 and 8 days, respectively, for the binary system. An EtSH:DMDS:MPS (3:2:1) ternary mixture was removed efficiently by the rebooted system after starvation. CONCLUSION: The proposed system can be applied to cost‐effectively decompose a mixture of volatile organic sulfide compounds at pilot scale. Copyright © 2011 Society of Chemical Industry  相似文献   

11.
BACKGROUND: To investigate the microbial degradation performance of organic pollutants in the atmosphere using a biotrickling filter, two microorganism strains, Bacillus cereus S1 and Bacillus cereus S2, were selected, identified and inoculated into a twin biotrickling filter for comparison. RESULTS: Both strains showed good performance towards the degradation of model organic pollutants when gas flow rates ranged from 100 to 600 L h−1. For S1, the total maximum removal efficiency (RE) of toluene was maintained nearly 100% not only at gas flow rates of 100 L h−1 corresponding to empty bed residence time (EBRT) 199.44 s, but also at gas flow rates of 200 L h−1 (EBRT = 99.72 s) and 300 L h−1 (EBRT = 66.48 s). However, S2 had a much lower degradation capability; near 100% removal efficiency was obtained only at the gas flow rate of 100 L h−1 although both bacteria belong to the same Bacillus cereus. With further increase in gas flow rate, the total REs for both S1 and S2 decreased slightly at first and then dropped sharply to 46% and 35%, respectively, at an EBRT of 33.24 s, corresponding to a gas flow rate of 600 L h−1. Starvation for between 2 and 10 days resulted in the re‐acclimation times of both strains ranging between 1.0 and 15.5 h. CONCLUSION: Strain S1 would be a better choice for inoculation into a biotrickling filter than strain S2, because of its much higher toluene removal capacity and rapid recovery to full performance. Copyright © 2008 Society of Chemical Industry  相似文献   

12.
The catalytic oxidation of volatile organic compounds (VOCs) into mainly CO2 and H2O appears promising in the context of the abatement of atmospheric gaseous pollutants and is the subject of this paper. The catalytic oxidation of toluene and m-xylene was carried out in a tubular reactor on a phenolic resin based activated carbon fiber (ACF) impregnated with nitrates of Co, Cr, Ni, and Cu at the reaction temperatures below 300 °C. The extent of the removal (i.e. conversion) of toluene and m-xylene was examined through the breakthrough curves and was found to strongly depend on the types and loading of the metal precursors. The experimental results showed that the reaction rate was significantly affected by O2 concentration below 3% (v/v). The performance of the ACFs impregnated with 5% (w/w) of Ni oxide was found to be superior to that of other metal oxides at reaction temperature between 170 and 290 °C. A surface kinetic mechanism for the catalytic oxidation of volatile organic compound was proposed and incorporated in a transport model developed to explain the experimental breakthrough data.  相似文献   

13.
Styrene was polymerized with β-pinene at 30°C and ?50°C in methylene dichloride and in m-xylene solvents. The styrene-to-terpene ratio on pyrolysis suggests that the polymers made at 30°C are copolymers, whereas most of the polymers made at ?50°C are not copolymers. On the other hand, ethylene, ethane, and propylene analysis suggests that more of the polymers made at ?50°C are copolymers.  相似文献   

14.
Oligomers of o-xylene (molecular weight ~400) have been shown to be novel heat stabilizers for isotatic polypropylene film. In contrast to unstabilized films which failed after exposure of 5 hr in a 150°C air oven, a film containing 0.5 and 1.0 wt-% of the o-xylene oligomer survived 100 and 226 hr, respectively. Furthermore if 0.25% dilauryl thiodipropionate (DLTP) was also present, only 0.1 wt-% of the o-xylene oligomer was necessary to maintain a film intact up to 124 hr. In absence of DLTP, this low level of o-xylene oligomer was not effective. Oligomers of toluene, m-xylene, p-xylene, and 1,2,4-trimethyl benzene were found less effective.  相似文献   

15.
Semi-continuous column fractionation of canola oil deodorizer distillate using supercritical CO2 (SCCO2) was carried out to determine the feasibility of value-added processing of this feed material for the recovery of bioactive components such as sterols and tocopherols and to determine the effect of operating conditions [pressure (20, 25 MPa using a temperature gradient of 70–100 °C), temperature (70, 100 °C) and a linear temperature gradient (70–100 °C at 25 MPa)] on extract yield and separation efficiency. Total extract yield increased significantly (p ≤ 0.05) with pressure, whereas at isobaric conditions (25 MPa) the highest yield was obtained at the lowest temperature tested (70 °C). Fractionation efficiency was reflected in the composition of fractions and was affected by operating conditions. Residue composition was determined by extract yield in addition to selectivity. Use of the thermal gradient (70–100 °C) decreased the content of volatiles, free fatty acids and tocopherols while increasing sterol content significantly (p ≤ 0.05) to a level of 40% (GC area %) in the residue obtained at 25 MPa. The findings indicate the potential of canola oil deodorizer distillate as a source of sterols and warrant further research on the countercurrent column fractionation to improve the separation efficiency.  相似文献   

16.
The physical parameters of the xylene isomers (the positional isomers o-, m-, and p-xylenes and the skeletal isomer ethyl benzene) responsible for the differing permeation behavior of the isomers through lined unsupported 0.41 mm thick nitrile glove material were investigated. An ASTM type permeation cell at 30°C, constant mixing conditions, hexane liquid collection, and capillary column gas chromatography/mass spectrometry of samples taken from the collection side every ten minutes allowed break through times tb and steady-state sections to be defined. While pure isomers had distinct break through times tb(m-xylene = p-xylene < ethyl benzene = o-xylene), steady-state permeation rates Ps(p-xylene > m-xylene > ethyl benzene = o-xylene), lag times tl(m-xylene < p-xylene = ethyl benzene < o-xylene), and diffusion coefficients Dp(m-xylene < p-xylene = ethyl benzene < o-xylene), such behavior was lost in a equal volume mixture (tb, tl, Ps, and Dp were equivalent). The average Ps of the mixture isomers of equal volumes did not differ from that expected from the individual pure isomer Ps values. The results for the pure isomers were attributed to o-xylene and ethyl benzene being similarly sterically hindered, the p-xylene being the flattest and most symmetrical molecule and having no dipole moment, and m-xylene being intermediate in steric structure. The pure isomer tl were directly related to viscosity divided by the log octanol-water coefficient, while their log Ps was inversely related to dipole moment times the logarithm of the capacity factor for water for a reversed-phase high-performance liquid chromatography column. In an equivolume mixture of the isomers, isomer interactions caused equivalence for all permeation kinetic parameters, indicating that the kinetics of mixture constituents is not predictable from the behavior of the pure constituents, although mass transfer appears additive. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 63: 1713–1721, 1997  相似文献   

17.
The relative rates of toluene transalkylation, disproportionation and benzene transalkylation of m-xylene over a catalyst consisting of 72% HY zeolite, 18% β-AlF3 and 10% (all w/w) Cu have been studied at a temperature of 723K, reciprocal space velocity of 352 g h?1 mol?1 and at total pressures varying between 101 and 911.93 kPa. The catalysed disproportionation of m-xylene is fastest and this has been explained in terms of the greater basicity of m-xylene over either benzene or toluene, whereas the catalyst is acidic. Thus m-xylene would be expected to adsorb more strongly on the catalyst.  相似文献   

18.
Effect of temperature on the performance of microbial fuel cells   总被引:1,自引:0,他引:1  
Single and double chamber microbial fuel cells (MFCs) were tested in batch mode at different temperatures ranging from 4 to 35 °C; results were analysed in terms of efficiency in soluble organic matter removal and capability of energy generation. Brewery wastewater diluted in domestic wastewater (initial soluble chemical oxygen demand of 1200 and 492 mg L−1 of volatile suspended solids) was the source of carbon and inoculum for the experiments. Control reactors (sealed container with support for biofilm formation) as well as baseline reactors (sealed container with no support) were run in parallel to the MFCs at each temperature to assess the differences between water treatment including electrochemical processes and conventional anaerobic digestion (in the presence of a biofilm, or by planktonic cells). MFCs showed improvements regarding rate and extent of COD removal in comparison to control and baseline reactors at low temperatures (4, 8 and 15 °C), whilst differences became negligible at higher temperatures (20, 25, 30 and 35 °C). Temperature was a crucial factor in the yield of MFCs both, for COD removal and electricity production, with results that ranged from 58% final COD removal and maximum power of 15.1 mW m−3 reactor (8.1 mW m−2 cathode) during polarization at 4 °C, to 94% final COD removal and maximum power of 174.0 mW m−3 reactor (92.8 mW m−2 cathode) at 35 °C for single chamber MFCs with carbon cloth-based cathodes. Bioelectrochemical processes in these MFCs were found to have a temperature coefficient, Q10 of 1.6.A membrane-based cathode configuration was tested and gave promising results at 4 °C, where a maximum power output of 294.6 mW m−3 reactor (98.1 mW m−2 cathode) was obtained during polarization and a maximum Coulombic efficiency (YQ) of 25% was achieved. This exceeded the performance at 35 °C with cloth-based cathodes (174.0 mW m−3; YQ 1.76%).  相似文献   

19.
High-density polyethylene (HDPE) was compounded with untreated and surface-treated mica (10, 20, 40 wt %) and composites were injection-molded. The composites were radiation crosslinked (100, 300, 700 kGy) and hydrocarbon permeability, tensile impact strength, and tensile strength at 25 and 80°C of the composites were examined. The permeability of HDPE decreased from 7 to 3.6 g/(d × m2) by compounding the polymer with 20 wt % mica, and the permeability was additionally reduced to 1.3 g/(d × m2) by irradiation of the compounds (700 kGy). When surface-treated mica was used, the permeability of the composite furthermore decreased to about 1.0 g/(d × m2). Upon irradiation, the E modulus measured at 25°C increased 5% when the dose was 300 kGy. At 80°C, the corresponding increase was 40%. The tensile impact strength of an unfilled polymer increased more than three times by an irradiation dose of 700 kGy, and for a polymer with 10 wt % mica, the tensile impact strength was twice the level of an unirradiated composite. © 1996 John Wiley & Sons, Inc.  相似文献   

20.
Polyacrylonitrile was grafted onto kenaf fibers in aqueous media by ceric ion–p-xylene redox system. The graft yield dependence on p-xylene concentration in the range 1.8–45.0 × 10−4M showed a minimum accompanied by an enhanced yield. This suggested the existence of two kinetically distinct grafting reactions associated with two precursor-initiating species, a p-xylyl radical and a diradical. The frequency of graft Fg and the average molecular weight of grafted polymer Mv were inversely related at varying concentrations of p-xylene with values for Mv of up to 5.18 × 104. The effect of ceric ion concentration on Mv showed that the ion is nonterminating at low concentrations in the range 8.3–33.3 × 10−3M. The graft yield showed positive temperature dependence in the region of 30–40°C and a negative one at higher temperatures, resulting in a decrease in the initial rate of graft at 70°C by a factor of 8 compared to its value at 40°C. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 1751–1755, 1999  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