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 46 毫秒
1.
Ambient levels of chlorinated gases and aerosol components were measured by online chemical ionization and aerosol mass spectrometers after an indoor floor were repeatedly washed with a commercial bleach solution. Gaseous chlorine (Cl2, 10's of ppbv) and hypochlorous acid (HOCl, 100's of ppbv) arise after floor washing, along with nitryl chloride (ClNO2), dichlorine monoxide (Cl2O), and chloramines (NHCl2, NCl3). Much higher mixing ratios would prevail in a room with lower and more commonly encountered air exchange rates than that observed in the study (12.7 h?1). Coincident with the formation of gas‐phase species, particulate chlorine levels also rise. Cl2, ClNO2, NHCl2, and NCl3 exist in the headspace of the bleach solution, whereas HOCl was only observed after floor washing. HOCl decays away 1.4 times faster than the air exchange rate, indicative of uptake onto room surfaces, and consistent with the well‐known chlorinating ability of HOCl. Photochemical box modeling captures the temporal profiles of Cl2 and HOCl very well and indicates that the OH, Cl, and ClO gas‐phase radical concentrations in the indoor environment could be greatly enhanced (>106 and 105 cm?3 for OH and Cl, respectively) in such washing conditions, dependent on the amount of indoor illumination.  相似文献   

2.
Carbamazepine reactivity and fate during chlorination was investigated in this study. From a kinetic standpoint, a third-order reaction (first-order relative to the CBZ concentration and second-order relative to the free chlorine concentration) was observed at neutral and slightly acidic pH, whereas a second-order reaction (first order relative to the CBZ concentration and first order relative to the free chlorine concentration) was noted under alkaline conditions. In order to gain insight into the observed pH-dependence of the reaction order, elementary reactions (i.e. reactions of Cl2, Cl2O, HOCl with CBZ and of ClO with CBZ or of HOCl with the ionized form of CBZ) were highlighted and second order rate constants of each of them were calculated. Close correlations between the experimental and modeled values were obtained under these conditions. Cl2 and Cl2O were the main chlorination agents at neutral and acidic pH. These results indicate that, for a 1 mg/L free chlorine concentration and 1–10 mg/L chloride concentration at pH 7, halflives about 52–69 days can be expected. A low reactivity of chlorine with CBZ could thus occur under the chlorination steps used during water treatment. From a mechanistic viewpoint, several transformation products were observed during carbamazepine chlorination. As previously described for the chlorination of polynuclear aromatic or unsaturated compounds, we proposed monohydroxylated, epoxide, diols or chlorinated alcohol derivatives of CBZ for the chemical structures of these degradation products. Most of these compounds seem to accumulate in solution in the presence of excess chlorine.  相似文献   

3.
Chlorine and chloramines are volatile compounds which are stripped (“flashed off”) from recirculating cooling water systems by the large volumes of air which flow through the water in the cooling tower. The fraction of a volatile gas, such as hypochlorous acid (HOCl), which is removed by stripping is determined by Henry's constant H for that gas: H = XG/XL, where XG is the mole fraction of the gas in the air and XL is the mole fraction of the gas in the water. We have measured H for HOCl, OCl, NH3, NH2Cl, NHCl2 and NCl3 at 20 and 40°C. We found H = 0.076 for HOCl, compared to 0.71 for NH3, at 20°C. At 40°C, H was about 2.5-fold larger for HOCl. This means that 10–15% of the HOCl is stripped from cooling water on each passage through a typical cooling tower. The measured flashoff of free available chlorine (HOCl + OCl) was markedly pH-sensitive with a pK of 7.5, exactly as expected if HOCl is volatile but OCl is not. The data permit a quantitative understanding of the fate of chlorine in cooling systems. The values of H at 40°C for NH2Cl, NHCl2 and NCl3 were 1.28, 3.76 and 1067. This means that all of the chloramines are quickly stripped in a cooling tower.  相似文献   

4.
Chlorine and chloramines are volatile compounds which are stripped (“flashed off”) from recirculating cooling water systems by the large volumes of air which flow through the water in the cooling tower. The fraction of a volatile gas, such as hypochlorous acid (HOCl), which is removed by stripping is determined by Henry's constant H for that gas: H = XG/XL, where XG is the mole fraction of the gas in the air and XL is the mole fraction of the gas in the water. We have measured H for HOCl, OCl?, NH3, NH2Cl, NHCl2 and NCl3 at 20 and 40°C. We found H = 0.076 for HOCl, compared to 0.71 for NH3, at 20°C. At 40°C, H was about 2.5-fold larger for HOCl. This means that 10–15% of the HOCl is stripped from cooling water on each passage through a typical cooling tower. The measured flashoff of free available chlorine (HOCl + OCl?) was markedly pH-sensitive with a pK of 7.5, exactly as expected if HOCl is volatile but OCl? is not. The data permit a quantitative understanding of the fate of chlorine in cooling systems. The values of H at 40°C for NH2Cl, NHCl2 and NCl3 were 1.28, 3.76 and 1067. This means that all of the chloramines are quickly stripped in a cooling tower.  相似文献   

5.
The fate of pharmaceuticals in the aquatic environment is significantly affected by the presence of humic acids (HA). In this work, the synergistic effect of electrochemical oxidation (EO) and ultraviolet irradiation (UVI) was evaluated for HA removal and for the simultaneous degradation of three pharmaceuticals (carbamazepine, propranolol and sulfamethoxazole) in presence of HA. The effectiveness of EO, UVI and their combination for HA removal was assessed using different operating parameters, such as type of anode (Nb/BDD and Ti/IrO2), supporting electrolyte (NaCl, NaBr and Na2SO4), current density (8.1, 16.1, 28.2, 40.3, and 48.4 mA/cm2), pH (3, 7 and 9) and NaCl electrolyte concentration (7, 14 and 21 mM). The use of non‐active anode Nb/BDD, NaCl electrolyte and combination EO‐UVI was the most efficacious option for HA removal, due to the production of hydroxyl radicals as well as active chlorine species (HClO, Cl and ClO?) generated by anodic oxidation and by UVI. The effectiveness of the EO process was enhanced coupling EO with UVI, however the energetic consumption increased. The composition of the electrolyte was the pivotal parameter since a complete degradation of the pharmaceuticals was achieved by both processes EO and EO‐UVI using NaCl as electrolyte; this is attributed to the indirect oxidation by electrogenerated active chlorine which dominates the pharmaceuticals degradation.  相似文献   

6.
The aim of our work was to study the reaction between chlorine dioxide (ClO2) and organic substances. In the first part of our survey, chlorine dioxide demands were measured in diluted aqueous solutions of various kinds of organic compounds (5 × 10−5 −10 M) at pH 7. In the second part, the study of the action of chlorine dioxide on phenols (phenol, di and triphenols) was undertaken by observing the change of the organic substance through global parameter controls (COT, u.v. spectrum) and by trying to identify a certain number of oxidation products by means of chromatographic analysis (HPLC and GC), of mass spectrometry and of Nuclear Magnetic Resonance spectrometry (NMR).  相似文献   

7.
Ozone reacts with free aqueous chlorine when present as hypochlorite ion (OCl) with a second order rate constant of 120 ± 15 M−1 s−1 at 20°C. About 77% of the chlorine reacts to produce Cl and 23% is oxidized to ClO3. No ClO4 is formed. Conversion of chlorine to monochloramine reduces the ozone reaction rate to 26 ± 4 M−1 s−1, independent of pH, NH2Cl is transformed quantitatively to NO3 and Cl by O3. Rate data for other chloramines are also presented. The direct reaction of ozone with chlorine accounts for a significant amount of the chlorine and ozone demand found when the two oxidants are used in combination under water works conditions.  相似文献   

8.
Second-order rate constants for reactions of ozone with 40 inorganic aqueous solutes are reported. Included are compounds of sulfur (e.g. H2S, H2SO3, HOCH2SO3H), chlorine (e.g. Cl, HOCl, NH2Cl, HClO2, ClO2), bromine (e.g. Br, HOBr), nitrogen (e.g. NH3, NH2OH, N2O, HNO2) and oxygen (e.g. H2O2), as well as free radicals (e.g. O2, OH). Most of these compounds exhibit an increase in rate constant with increasing pH corresponding to their degree of dissociation. Rate constants are based on ozone consumption rates measured by conventional batch-type or continuous-flow methods (10−3-10+6 M−1 s−1 range) and determinations of stoichiometric factors. Also listed are data determined by pulse-irradiation techniques using kinetic spectroscopy (1010 M−1 s−1 range). Additional literature data are reviewed for completeness. Results are discussed with respect to water treatment and environmental processes.  相似文献   

9.
Edmondo Canelli 《Water research》1980,14(10):1533-1540
Chlorine in the forms of HOCl or ClO was determined rapidly and precisely in the range from 0.10 to 3.0 ppm chlorine, without titration, using a bare-electrode portable amperometric analyzer. The instrument was calibrated with a 1.00 ppm chlorine standard solution or an equivalent permanganate solution which is stable for at least 6 months. The detection limit was 0.10 ppm chlorine, and relative standard deviations (N = 10) were <6.0% in potable or swimming-pool waters containing 0.25, 1.00 and 2.50 ppm or 0.35, 1.40 and 2.70 ppm chlorine respectively. Bromine-, triiodide ion-, and Mn (VII)-generated signals were stoichiometrically equivalent to that of hypochlorite, but MnO2 suspensions (1.5, 5.0 or 15 ppm) did not produce detectable amperometric signals.Analysis of solutions of hypochlorite with ammonium chloride or selected organic nitrogen compounds indicated that various N-chloro compounds may interfere. In the presence of N-chloroglycine (2.70 ppm chlorine) the amperometric signal was about 5% of that for the equivalent concentration of hypochlorite, but higher relative responses were obtained with NH2Cl, NHCl2 or NCl3 (19, 42 and 70% at 2.60, 1.20 or 1.00 ppm chlorine respectively). Chlorinated urea (2.2 ppm N, 2.90 ppm chlorine), chlorinated bovine albumin (10 ppm albumin, 2.00 ppm chlorine) or monochloroisocyanurate (1.30 ppm chlorine) produced amperometric signals (76, <5 and 50% respectively) which are lower than those obtained by the N,N-diethyl-p-phenylenediamine (DPD) method (100, 12 and 92%).Twenty potable water samples were analyzed for free chlorine by the DPD and amperometric procedures. Statistical analysis showed no significant differences between the two sets of results (P < 0.1). Swimming-pool-water samples were also analyzed by the two methods in the field (22 samples) and in the laboratory (24 samples). In each set of results the mean free-chlorine value by the DPD procedure was significantly higher than that obtained amperometrically (P > 0.005). This discrepancy was associated with the probable presence of chlorinated urea, whose signal by amperometry is lower than that by the DPD method.The advantages of this simple procedure must be weighed against possible inaccurate results in the presence of NH2Cl, NHCl2 or NCl3.  相似文献   

10.
Reverse osmosis (RO) membranes have been successfully applied around the world for wastewater reuse applications. However, RO is a physical separation process, and besides the clean water stream (permeate) a reverse osmosis concentrate (ROC) is produced, usually representing 15-25% of the feed water flow and containing the organic and inorganic contaminants at higher concentrations. In this study, electrochemical oxidation was investigated for the treatment of ROC generated during the reclamation of municipal wastewater effluent. Using laboratory-scale two-compartment electrochemical systems, five electrode materials (i.e. titanium coated with IrO2-Ta2O5, RuO2-IrO2, Pt-IrO2, PbO2, and SnO2-Sb) were tested as anodes in batch mode experiments, using ROC from an advanced water treatment plant. The best oxidation performance was observed for Ti/Pt-IrO2 anodes, followed by the Ti/SnO2-Sb and Ti/PbO2 anodes. The effectiveness of the treatment appears to correlate with the formation of oxidants such as active chlorine (i.e. Cl2/HClO/ClO). As a result, electro-generated chlorine led to the abundant formation of harmful by-products such as trihalomethanes (THMs) and haloacetic acids (HAAs), particularly at Ti/SnO2-Sb and Ti/Pt-IrO2 anodes. The highest concentration of total HAAs (i.e. 2.7 mg L−1) was measured for the Ti/SnO2-Sb electrode, after 0.55 Ah L−1 of supplied specific electrical charge. Irrespective of the used material, electrochemical oxidation of ROC needs to be complemented by a polishing treatment to alleviate the release of halogenated by-products.  相似文献   

11.
C.H. Rav-Acha  R. Blits 《Water research》1985,19(10):1273-1281
The removal of PAH from surface water by disinfectants like chlorine or chlorine dioxide is important where contamination by these compounds is concerned and no other water treatment processes are available. Our particular interest in these reactions arise from the fact that PAH can be used as an excellent model for the investigation of the different mechanisms by which the two oxidants react with aquatic organics. The vast differences between the rates of Cl2 and ClO2 reactions with various PAH, as well as the physical and chemical factors influencing those reactions indicate that chlorine reacts with PAH by several possible mechanisms, e.g. addition, substitution and oxidation. Chlorine dioxide on the other hand reacts mainly as a pure oxidant and a one-electron acceptor. As a consequence, chlorine dioxide reacts much more specifically with those PAH that undergo facile oxidation. Therefore, some PAH that react quite easily with Cl2, do not react at all with ClO2, while other PAH react with ClO2 much more rapidly than with Cl2. The widespread and highly carcinogenic benzo(a)pyrene and benzo(a)anthracene for example react with ClO2 much faster than with Cl2.  相似文献   

12.
《Water research》1987,21(7):849-857
Resorcinol reacted with chlorine and chlorine dioxide at pH 7 in dilute aqueous solution and in the presence of granular activated carbon (GAC) to produce numerous chlorinated compounds. Major products included chlororesorcinols and various chlorinated cyclopentenediones. Few reaction products were found in common to both Cl2 and ClO2 when reacted with resorcinol in the absence of GAC. However, both oxidants produced the same major recoverable chlorinated product, a dichlorocyclopentenedione, when reacted with resorcinol preadsorbed on GAC columns. No products were found when aqueous resorcinol was applied to GAC columns following contact of the carbon with Cl2 or ClO2 solutions.  相似文献   

13.
With an increasing use of indoor disinfectants such as chlorine (Cl2) and hypochlorous acid, a convenient sampler for estimating exposure to oxidants, such as effective chlorine, is necessary. Here, we developed a personal passive air sampler (PPAS) composed of a redox dye, o-dianisidine, in a polydimethylsiloxane (PDMS) sheet. o-Dianisidine readily reacts with gaseous oxidants generated by bleach usage, and its color changes as the reaction progresses; hence, personal exposure to effective chlorine could be easily detected by the naked eye, while cumulative exposure could be determined by measuring concentrations of o-dianisidine reacting with it. The PPAS was calibrated, and a sampling rate of 0.00253 m3/h was obtained using a small test chamber. The PPAS was tested with the help of ten volunteers whose personal exposure to Cl2-equivalent gas was estimated after bathrooms were cleaned using spray and liquid-type household disinfection products, and the accumulated exposure-gas concentrations ranged from 69 to 408 ppbv and 148 to 435 ppbv, respectively. These PPAS-derived exposure concentrations were approximately two orders lower than those estimated using ConsExpo, suggesting a significant overestimation by prevailing screening models, possibly due to the ignorance of transformation reactions.  相似文献   

14.
Modeling monochloramine loss in the presence of natural organic matter   总被引:2,自引:0,他引:2  
A comprehensive model describing monochloramine loss in the presence of natural organic matter (NOM) is presented. The model incorporates simultaneous monochloramine autodecomposition and reaction pathways resulting in NOM oxidation. These competing pathways were resolved numerically using an iterative process evaluating hypothesized reactions describing NOM oxidation by monochloramine under various experimental conditions. The reaction of monochloramine with NOM was described as biphasic using four NOM specific reaction parameters. NOM pathway 1 involves a direct reaction of monochloramine with NOM (kdoc1=1.05×104-3.45×104 M−1 h−1). NOM pathway 2 is slower in terms of monochloramine loss and attributable to free chorine (HOCl) derived from monochloramine hydrolysis (kdoc2=5.72×105-6.98×105 M−1 h−1), which accounted for the majority of monochloramine loss. Also, the free chlorine reactive site fraction in the NOM structure was found to correlate to specific ultraviolet absorbance at 280 nm (SUVA280). Modeling monochloramine loss allowed for insight into disinfectant reaction pathways involving NOM oxidation. This knowledge is of value in assessing monochloramine stability in distribution systems and reaction pathways leading to disinfection by-product (DBP) formation.  相似文献   

15.
The effectiveness of chlorine (Cl2) and chlorine dioxide (ClO2) in controlling biofouling of 304L stainless steel heat exchanger tubing was compared using an experimental trough system. Three combinations of dose and contact time were evaluated. Chlorination coupled with a dispersant was also tested. Three criteria were used to assess the degree of fouling; organic carbon and dry weight of the fouling material accumulated on metal specimens and the visual appearance of this material on the specimens. These parameters correlated well with one another and therefore, collectively provided an effective means of evaluating biocide efficacy.Metal specimens in all troughs receiving biocide treatment were much less fouled than those in the trough receiving no biocide. Continuous application of Cl2 at about 0.15 ppm was more effective than four 15-min 1 ppm Cl2 applications per day. Both of these treatment regimes were more effective than a dose of about 1 ppm for 1 h day−1. Use of a dispersant in combination with Cl2 showed no significant reduction in the amount of biofouling material accumulation, although a difference in the texture of this material was observed. Unlike the Cl2 results, low-level continuous ClO2 treatment at 0.15 ppm resulted in biofouling similar to that when 1 ppm of ClO2 was used for 1 h day−1. Overall, ClO2 was significantly (P < 0.05) more effective in controlling biofouling than Cl2.  相似文献   

16.
This research studied As(III) and As(V) removal during electrocoagulation (EC) in comparison with FeCl3 chemical coagulation (CC). The study also attempted to verify chlorine production and the reported oxidation of As(III) during EC. Results showed that As(V) removal during batch EC was erratic at pH 6.5 and the removal was higher-than-expected based on the generation of ferrous iron (Fe2+) during EC. As(V) removal by batch EC was equal to or better than CC at pH 7.5 and 8.5, however soluble Fe2+ was observed in the 0.2-μm membrane filtrate at pH 7.5 (10-45%), and is a cause for concern. Continuous steady-state operation of the EC unit confirmed the deleterious presence of soluble Fe2+ in the treated water. The higher-than-expected As(V) removals during batch mode were presumed due to As(V) adsorption onto the iron rod oxyhydroxides surfaces prior to the attainment of steady-state operation. As(V) removal increased with decreasing pH during both CC and EC, however EC at pH 6.5 was anomalous because of erratic Fe2+ oxidation. The best adsorption capacity was observed with CC at pH 6.5, while lower but similar adsorption capacities were observed at pH 7.5 and 8.5 with CC and EC. A comparison of As(III) adsorption showed better removals during EC compared with CC possibly due to a temporary pH increase during EC. In contrast to literature reports, As(III) oxidation was not observed during EC, and As(III) adsorption onto iron hydroxides during EC was only 5-30% that of As(V) adsorption. Also in contrast to literature, significant Cl2 was not generated during EC, in fact, the rods actually produced a significant chlorine demand due to reduced iron oxides on the rod. Although Cl2 generation and As(III) oxidation are possible using a graphite anode, a combination of graphite and iron rods in the same EC unit did not produce As(III) oxidation. However, a two-stage process (graphite anode followed by iron anode in separate chambers) was effective in As(III) oxidation and removal. The competing ions, silica and phosphate interfered with As(V) adsorption during both CC and EC. However, the degree of interference depends on the concentration and presence of other competing ions. In particular, the presence of silica lowered the effect of phosphate with increasing pH due to silica’s own significant effect at high pHs.  相似文献   

17.
Watts MJ  Linden KG 《Water research》2007,41(13):2871-2878
The photodegradation of chlorine-based disinfectants NH(2)Cl, HOCl, and OCl(-) under UV irradiation from low- (LP) and medium-pressure (MP) Hg lamps was studied. The quantum yields of aqueous chlorine and chloramine under 254nm (LP UV) irradiation were greater than 1.2molEs(-1) for free chlorine in the pH range of 4-10 and 0.4molEs(-1) for monochloramine at pH 9. Quantum yields for MP (200-350nm) ranged from 1.2 to 1.7molEs(-1) at neutral and basic pH to 3.7molEs(-1) at pH 4 for free chlorine, and 0.8molEs(-1) for monochloramine. Degradation of free chlorine was enhanced under acidic water conditions, but water quality negatively impacted the MP Hg lamp degradation of free chlorine, compared to the LP UV source. The production of hydroxyl radical via chlorine photolysis was assessed along with the rate of reaction between ()OH and HOCl using radical scavengers (parachlorobenzoic acid and nitrobenzene) in chlorinated solutions at pH 4. The quantum yield of OH radical production from HOCl at 254nm was found to be 1.4molEs(-1), while the reaction of HOCl with OH radical was measured as 8.5x10(4)M(-1)s(-1). NH(2)Cl was relatively stable in all irradiated solutions, with <0.3mgL(-1) increase in nitrate following a UV dose of 1000mJcm(-2). For water treatment plants, no significant changes in chlorine concentration would be expected under typical pH levels and UV doses; however, the formation of ()OH could have implications for chlorinated byproducts or decay of unwanted chemical contaminants.  相似文献   

18.
Yeon Jung Jung 《Water research》2010,44(18):5345-5355
The characteristics of chlorate (ClO3) and perchlorate (ClO4) formation were studied during the electrolysis of water containing chloride ions (Cl). The experiments were performed using an undivided Pt/Ti plate electrode under different pH conditions (pH 3.6, 5.5, 7.2, 8.0 and 9.0). ClO3 and ClO4 were formed during electrolysis in proportion to the Cl concentration. The generation rates of ClO3 and ClO4 under acidic conditions (pH 3.6 and 5.5) were lower than in basic pH conditions (pH 7.2, 8.0 and 9.0). However, the pH of the solution did not influence the conversion of ClO3 to ClO4. The effects of intermediately formed oxidants on the production of ClO3 and ClO4 were observed using sodium thiosulfate (Na2S2O3) as the active chlorine scavenger and tertiary butyl alcohol (t-BuOH) as the hydroxyl radical (OH) scavenger. The results revealed that electrolysis reactions that involved active chlorine contributed dominantly to ClO3 production. The direct oxidation reaction rate of Cl to ClO3 was 13%. The OH species that were intermediately formed during electrolysis were also found to significantly affect ClO3 and ClO4 production. The key formation pathways of ClO3 and ClO4 were studied using kinetic model development.  相似文献   

19.
The reaction between chlorite (CIO2) and vanillic acid, at pH 6.0 in the presence of granular activated carbon (GAC), yielded several reaction products identifiable by GC/MS; no products were found in the absence of GAC. Indan and C1O2 or C1O2 reacted in aqueous solution and gave similar products in the presence and absence of GAC. Carbon exposed to C1O2 appears to become capable of promoting hydroxylation, decarboxylation, ring cleavage and CO2 addition reactions with vanillic acid.Chlorite in aqueous solution was reduced by GAC predominantly to Cl. The capacity of one type of virgin GAC for C1O2 reduction to Cl was about 80–90 mg C1O2 g−1 GAC before the rate of reaction was sharply reduced.  相似文献   

20.
Soufan M  Deborde M  Legube B 《Water research》2012,46(10):3377-3386
Diclofenac reactivity and fate during water chlorination was investigated in this work. In the first step, chlorination kinetic of diclofenac (DCF) was studied in the pH range of 4-10 at 20 ± 2 °C and in the presence of an excess of total chlorine. A second-order reaction (first-order relative to DCF concentration and first-order relative to free chlorine concentration) was shown with rate constant about 3.89 ± 1.17 M−1 s−1 at pH 7. The elementary reactions (i.e. reactions of hypochlorous acid (HOCl) with neutral and ionized forms of DCF, and acid-catalysed reaction of HOCl with neutral and ionized forms of DCF) were proposed to explain the pH-dependence of the rate constants and intrinsic constant of each of them were calculated. In the second step, several degradation products formed during chlorination of DCF were identified. These compounds could come from an initial chlorine electrophilic attack on aromatic ring or amine function of DCF. Some of these chlorinated derivatives seem to accumulate in solution in the presence of an excess of chlorine.  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