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.
The microwave-hydrothermal decomposition of persistent and bioaccumulative perfluorooctanoic acid (PFOA) in water with persulfate (S2O82−) at 60, 90, and 130 °C was examined to develop an effective technology for treating PFOA pollution. S2O82− is an efficient oxidant for degrading PFOA even at the room temperature of 27 °C. Higher temperature accelerates the PFOA decomposition rate, but an extremely high temperature (130 °C) will lead to the formation of significant amounts of radical oxidants that are released rapidly to consume most remaining persulfate thus causing a lower mineralization efficiency. The solution pH value is another important factor to influence the degradation rate; there is almost no PFOA decomposition reaction under alkaline conditions. The decomposition rate in acidic conditions is 1.1-7.4 times faster than in alkaline condition. Additionally, the proposed method is also effective in decomposing other PFCA species such as the C2-C7 perfluoroalkyl groups.  相似文献   

2.
Chen YC  Lo SL  Kuo J 《Water research》2011,45(14):4131-4140
Titanate nanotubes (TNTs) were used to remove perfluorooctanoic acid (PFOA) from aqueous solutions in this study. Direct photolysis of PFOA by a 254-nm UV light (400 W) was found effective to decompose PFOA without presence of photocatalysts. Shorter-chain perfluorocarboxylic acids (PFCAs) and fluoride ions were formed during photodecomposition. Addition of TNTs as photocatalysts did not greatly enhance photocatalytic decomposition of PFOA. TNTs mainly act as adsorbents to adsorb PFOA and form TNT-PFOA complexes. It suggested that sodium ions and oxygen atoms on the surfaces of TNTs play important roles in PFOA adsorption. X-ray Photoelectron Spectroscopy (XPS) and Fourier Transform Infrared Spectroscopy (FTIR) analyses indicated that ion-exchange, electrostatic interaction, and hydrophobic interaction all participated in the photocatalytic reaction of PFOA by TNTs.  相似文献   

3.
The oxidation of 2,4-dinitrotoluene (DNT) by persulfate (S2O82−) activated with zero-valent iron (Feo) was studied through a series of batch experiments. The mechanism for Feo activation was investigated by comparing with Fe2+, and the effects of persulfate-to-iron ratio and pre-reduction on DNT oxidation were examined. DNT was stable in the presence of persulfate and transformed only when Feo was added. Most DNT was degraded oxidatively by Feo-activated persulfate, whereas direct reduction of DNT by Feo was unimportant. The rate of DNT degradation increased with higher Feo dose, presumably due to increasing activation of persulfate by Feo and Fe2+. In contrast to the Feo-persulfate system, where complete oxidation DNT was achieved, only ≤ 20% of DNT was degraded and the reaction was terminated rapidly when Feo was replaced with equimolar Fe2+. This indicates that Feo is more effective than Fe2+ as activating agent and potentially more suitable for environmental applications. The reduction products of DNT were more rapidly oxidized by persulfate than DNT, suggesting that converting the nitro groups of NACs to amino groups prior to oxidation can greatly enhance their oxidation. This suggests that a sequential Feo reduction-persulfate oxidation process may be an effective strategy to promote NAC degradation.  相似文献   

4.
Potential of perfluorooctanoic acid (PFOA) to degrade via indirect photolysis in aquatic solution under conditions representing surface water was studied. Globally distributed and bioaccumulative PFOA does not absorb solar radiation by itself, but may be potentially photochemically transformed by the natural sensitizers such as dissolved organic matter (DOM), nitrate or ferric iron. Reaction solutions containing purified water, fulvic acid (representing DOM), nitrate, ferric iron or sea water from the Baltic Sea were spiked with PFOA and irradiated with an artificial sun (290-800 nm). In comparison similar samples were also irradiated under UV radiation at 254 nm in order to study the direct photolysis. UV radiation at 254 nm decomposed PFOA to perfluoroheptanoic-, perfluorohexanoic- and perfluoropentanoic acids. The samples irradiated with an artificial sun contained no decomposition products and no decrease in PFOA concentration was observed. According to the detection limit of the products and typical solar radiation at the surface of ocean, the photochemical half-life for PFOA was estimated to be at least 256 years at the depth of 0 m, > 5000 years in the mixing layer of open ocean and > 25,000 years in coastal ocean. This is significantly more than the previously reported photochemical half-life of PFOA (> 0.96 years).  相似文献   

5.
Photo-reductive defluorination of perfluorooctanoic acid in water   总被引:4,自引:0,他引:4  
Yan Qu  Fei Li  Jing Chen  Qi Zhou 《Water research》2010,44(9):2939-2947
Globally distributed and highly stable, perfluorooctanoic acid (PFOA) has prompted much concern regarding its accumulation in the natural environment and its threats to ecosystems. Therefore, it is desirable to develop an effective treatment against PFOA pollution. In this study, a photo-reduction method is developed and evaluated for the decomposition of perfluorooctanoic acid (PFOA) in aqueous phase with potassium iodide (KI) as a mediator. The experiment was conducted under 254 nm irradiation at room temperature and pH 9 under anaerobic conditions. Ultraviolet photolysis of iodide solutions led to the generation of hydrated electrons (eaq, Eaq/e°= −2.9 V), which contributed to the defluorination of PFOA. Defluorination was confirmed by fluoride release of 98%, indicating almost complete defluorination of PFOA. Kinetic analysis indicated that the PFOA decomposition fit the first-order model with a rate constant of 7.3 × 10−3 min−1. Besides fluoride ions, additional intermediates identified and quantified include formic acid, acetic acid, and six short-chain perfluorocarboxylic acids (C1-C6). Furthermore, small amounts of CF3H and C2F6 were also detected as reaction products by using GC/MS. With observation of the degradation products and verification via an isotopic labeling method, two major defluorination pathways of PFOA are proposed: direct cleavage of C-F bonds attacked by hydrated electrons as the nucleophile; and stepwise removal of CF2 by UV irradiation and hydrolysis. This method was applied to the decomposition of PFOA in wastewater issued from a fluorochemical plant and proved to be effective.  相似文献   

6.
Lin H  Niu J  Ding S  Zhang L 《Water research》2012,46(7):2281-2289
Electrochemical decomposition of environmentally persistent perfluorooctanoic acid (PFOA) in aqueous solution was investigated over Ti/SnO2-Sb, Ti/SnO2-Sb/PbO2, and Ti/SnO2-Sb/MnO2 anodes. The degradation of PFOA followed pseudo-first-order kinetics. The degradation ratios on Ti/SnO2-Sb, Ti/SnO2-Sb/PbO2, and Ti/SnO2-Sb/MnO2 anodes achieved 90.3%, 91.1%, and 31.7%, respectively, after 90 min electrolysis at an initial 100 mg/L PFOA concentration at a constant current density of 10 mA/cm2 with a 10 mmol/L NaClO4 supporting electrolyte solution. The defluorination rates of PFOA on these three anodes were 72.9%, 77.4%, 45.6%, respectively. The main influencing factors on electrochemical decomposition of PFOA over Ti/SnO2-Sb anode were evaluated, including current density (5-40 mA/cm2), initial pH value (3-11), plate distance (0.5-2.0 cm), and initial concentration (5-500 mg/L). The results indicated that PFOA (100 mL of 100 mg/L) degradation ratio and defluorination ratio achieved 98.8% and 73.9%, respectively, at the optimal conditions after 90 min electrolysis. Under this optimal condition, the degradation rate constant and the degradation half-life were 0.064 min−1 and 10.8 min, respectively. The intermediate products including short-chain perfluorinated carboxylic acids (PFCAs, C2∼C6) and perfluorocarbons (C2∼C7) were detected by electrospray ionization (ESI) mass spectrum. A possible electrochemical degradation mechanism of PFOA including electron transfer, Kolbe decarboxylation, radical reaction, decomposition, and hydrolysis was proposed. The electrochemical technique could be employed to degrade PFOA from contaminated wastewater as well as to reduce the toxicity of PFOA.  相似文献   

7.
The persulfate (S(2)O(8)(2-))-induced photochemical decomposition of C(3)F(7)CF=CHCOOH in water was investigated to develop a method to neutralize stationary sources of fluorotelomer unsaturated carboxylic acids (FTUCAs), which have recently been detected in the environment, and are considered to be more toxic than the environmentally persistent perfluorocarboxylic acids (PFCAs). Photolysis of S(2)O(8)(2-) produced highly oxidative sulfate radical anions (SO(4)(-)), which efficiently decomposed C(3)F(7)CF=CHCOOH to F(-) and CO(2) via C(3)F(7)COOH. With an initial S(2)O(8)(2-) concentration of 12.5mM and irradiation from a 200-W xenon-mercury lamp, C(3)F(7)CF=CHCOOH at a concentration of 680muM was completely decomposed within 5min. When 8.00mM S(2)O(8)(2-) was used, the initial rate of C(3)F(7)CF=CHCOOH decomposition induced by 254-nm light irradiation was 45 times as high as that with photolysis alone. The apparent quantum yield for the C(3)F(7)CF=CHCOOH decomposition with 6.25mM S(2)O(8)(2-) and 254-nm light was 2.4, indicating that virtually all SO(4)(-) anions produced by the photolysis of S(2)O(8)(2-) contribute to the decomposition of C(3)F(7)CF=CHCOOH.  相似文献   

8.
Perfluoroalkyl substances (PFAS) have unique properties that limit their degradability in the environment. One of these PFAS is an acid (PFOA). Electrochemical oxidation is a promising method for remediation, but energy costs are high. To limit the energy consumption, this study used a boron‐doped diamond (BDD) electrode stack and a combined current density technique that employed 50 mA/cm2 for the first 0.25 hours then lowered the current density to 1, 5, or 10 mA/cm2. This technique is similar to one developed previously; however, that method was only developed for compounds comprising of carbon, oxygen and nitrogen, whereas PFAS have the addition of fluorine. For the degradation of PFOA, the combined current density of 50 and 5 mA/cm2 (50&5) allowed for a 37% reduction in energy usage to obtain 75% defluorination compared to using 50 mA/cm2 alone. Further investigation into remediating an ion‐exchange regeneration solution shows great promise.  相似文献   

9.
The interactions of co-present Cr(VI) and As(V), and the influences of humic acid and bicarbonate in the process of Cr(VI) and As(V) removal by Fe0 were investigated in a batch setting using simulated groundwater with 5 mM NaCl, 1 mM Na2SO4, and 0.8 mM CaCl2 as background electrolytes at an initial pH value of 7. Cr(VI) and As(V) were observed to be subject to different impacts induced by co-existing As(V) or Cr(VI), humic acid and bicarbonate, originating from their distinct removal mechanisms by Fe0. Cr(VI) removal is a reduction-dominated process, whereas As(V) removal principally involves adsorption onto iron corrosion products. Experimental results showed that Cr(VI) removal was not affected by the presence of As(V) and humic acid. However, As(V) removal appeared to be inhibited by co-present Cr(VI). When the Cr(VI) concentration was 2, 5, and 10 mg/L, in the absence of humic acid and bicarbonate, As(V) removal rate constants were decreased by 27.9%, 49.0%, and 61.2%, respectively, which probably resulted from competition between Cr(VI) and As(V) for adsorption sites of the iron corrosion products. Furthermore, the presence of humic acid significantly varied As(V) removal kinetics by delaying the formation and aggregation of iron hydroxides due to the formation of soluble Fe-humate complexes and stably dispersed fine iron hydroxides colloids. In the presence of bicarbonate, both Cr(VI) and As(V) removal was increased and the inhibitory effect of Cr(VI) on As(V) removal was suppressed, resulting from the buffering effects and the promoted iron corrosion induced by bicarbonate, and the formation of CaCO3 in solution, which enhanced As(V) adsorption.  相似文献   

10.
Xu Y  Zhao D 《Water research》2007,41(10):2101-2108
Laboratory batch and column experiments were conducted to investigate the feasibility of using a new class of stabilized zero-valent iron (ZVI) nanoparticles for in situ reductive immobilization of Cr(VI) in water and in a sandy loam soil. Batch kinetic tests indicated that 0.08g/L of the ZVI nanoparticles were able to rapidly reduce 34mg/L of Cr(VI) in water at an initial pseudo first-order rate constant of 0.08h(-1). The extent of Cr(VI) reduction was increased from 24% to 90% as the ZVI dosage was increased from 0.04 to 0.12g/L. The leachability of Cr preloaded in a Cr-loaded sandy soil was reduced by nearly 50% when the soil was amended with 0.08g/L of the ZVI nanoparticles in batch tests at a soil-to-solution ratio of 1g: 10mL. Column experiments indicated that the stabilized ZVI nanoparticles are highly deliverable in the soil column. When the soil column was treated with 5.7 bed volumes of 0.06g/L of the nanoparticles at pH 5.60, only 4.9% of the total Cr was eluted compared to 12% for untreated soil under otherwise identical conditions. The ZVI treatment reduced the TCLP leachability of Cr in the soil by 90%, and the California WET (Waste Extraction Test) leachability by 76%. The stabilized ZVI nanoparticles may serve as a highly soil-dispersible and effective agent for in situ reductive immobilization of chromium in soils, groundwater, or industrial wastes.  相似文献   

11.
The influences of various geochemical constituents, such as humic acid, HCO3, and Ca2+, on Cr(VI) removal by zero-valent iron (Fe0) were investigated in a batch setting. The collective impacts of humic acid, HCO3, and Ca2+ on the Cr(VI) reduction process by Fe0 appeared to significantly differ from their individual impacts. Humic acid introduced a marginal influence on Fe0 reactivity toward Cr(VI) reduction, whereas HCO3 greatly enhanced Cr(VI) removal by maintaining the solution pH near neutral. The Cr(VI) reduction rate constants (kobs) were increased by 37.8% and 78.3%, respectively, with 2 mM and 6 mM HCO3 in solutions where humic acid and Ca2+ were absent. Singly present Ca2+ did not show a significant impact to Cr(VI) reduction. However, probably due to the formation of passivating CaCO3, further addition of Ca2+ to HCO3 containing solutions resulted in a decrease of kobs compared to solutions containing HCO3 alone. Ca2+ enhanced humic acid adsorption led to a minor decrease of Cr(VI) reduction rates. In Ca2+-free solutions, humic acid increased the amount of total dissolved iron to 25 mg/l due to the formation of soluble Fe-humate complexes and stably dispersed fine Fe (oxy)hydroxide colloids, which appeared to suppress iron precipitation. In contrast, the coexistence of humic acid and Ca2+ significantly promoted the aggregation of Fe (oxy)hydroxides, with which humic acid co-aggregated and co-precipitated. These aggregates would progressively be deposited on Fe0 surfaces and impose long-term impacts on the permeability of PRBs.  相似文献   

12.
13.
Applications of perfluorinated compounds (PFCs) have led to a PFC exposure of the general population worldwide. Most PFC human biomonitoring data are available from developed countries. Here we report for the first time PFC levels in serum from children and adults living in the low developed country of Afghanistan. Among a health cooperation project we had the chance to collect blood samples from 12 children (age 2.5-9 years) and 43 adults (age 20-65 years). 25 participants were from Kabul and 30 lived in a rural area. Drinking water samples were collected from 10 tap water and 16 well water sources. PFC levels were determined by HPLC and MS/MS detection after offline protein precipitation with acetonitrile. PFOS could be quantified in all blood samples (limit of quantification, LOQ: 0.1 µg/l). Median (range) was 1.2 µg/l (0.21-11.8 µg/l). Most PFOA (n = 43) and PFHxS levels (n = 42) were below LOQ of 0.5 µg/l. Maximum levels were 1.5 (PFOA) and 3.0 µg/l (PFHxS). All PFOS and PFOA concentrations in drinking water were below LOQ (PFOA 0.03 µg/l and PFOS 0.015 µg/l). It is concluded that exposure to PFCs also occurs in Afghanistan but on a very low level.  相似文献   

14.
Xiong Z  Zhao D  Pan G 《Water research》2007,41(15):3497-3505
Perchlorate has emerged as a widespread contaminant in groundwater and surface water. Because of the unique chemistry of perchlorate, it has been challenging to destroy perchlorate. This study tested the feasibility of using a new class of stabilized zero-valent iron (ZVI) nanoparticles for complete transformation of perchlorate in water or ion-exchange brine. Batch kinetic tests showed that at an iron dosage of 1.8 g L(-1) and at moderately elevated temperatures (90-95 degrees C), approximately 90% of perchlorate in both fresh water and a simulated ion-exchange brine (NaCl=6% (w/w)) was destroyed within 7h. An activation energy (Ea) of 52.59+/-8.41 kJ mol(-1) was determined for the reaction. Kinetic tests suggested that Cl(VII) in perchlorate was rapidly reduced to chloride without accumulation of any intermediate products. Based on the surface-area-normalized rate constant k(SA), starch- and CMC-stabilized ZVI nanoparticles degraded perchlorate 1.8 and 3.3 times, respectively, faster than non-stabilized ZVI particles. Addition of a metal catalyst (Al, Cu, Co, Ni, Pd, or Re) did not show any reaction improvement. This technology provides an effective method for complete destruction of perchlorate in both contaminated water and brine.  相似文献   

15.
Mak MS  Lo IM  Liu T 《Water research》2011,45(19):6575-6584
A column study was conducted using a combination of zero-valent iron (Fe0) and iron oxide-coated sand (IOCS) for removing Cr(VI) and As(V) from groundwater. The removal efficiency and mechanism of Cr(VI) and As(V), the effects of humic acid (HA), and the various configurations of Fe0 and IOCS were investigated. The results showed that the use of an Fe0 and IOCS mixture in a completely mixed configuration can achieve the highest removal of both Cr(VI) and As(V), whilst the effects of HA were marginal in using these reactive materials. The solid phase analysis revealed the occurrence of the synergistic effect in these reactive materials as Fe2+ can be adsorbed onto the IOCS and transform the iron oxides to magnetite, providing more reactive surface area for Cr(VI) reduction and reducing the passivation on the Fe0. As(V) can then be removed by adsorption onto these iron corrosion products. HA can be adsorbed onto the IOCS so that the impacts of the deposition of HA aggregates on the Fe0 surface can be reduced, thus enhancing the Fe0 corrosion.  相似文献   

16.
采用电化学/过硫酸盐耦合体系(E-PS过程)降解水中的有机药物卡马西平(CBZ)。实验采用了分批模式进行,研究了温度、过硫酸钠浓度、初始pH值、电压等因素对E-PS过程降解CBZ的影响。反应100min后,单独过硫酸钠、电解和E-PS过程对卡马西平的降解率分别为25.5%、59.3%、78.1%,TOC去除率分别为8.25%、23.48%、26.68%。升高温度可以有效提高CBZ的降解率。反应100min后,在288K,CBZ降解率为60.2%;在298K,CBZ降解率达到78.1%;而在308K,CBZ降解率为90.1%。CBZ的降解率随着过硫酸盐浓度的增加而提高。当过硫酸盐浓度为40g/L时,反应100min,CBZ降解率达94.7%。初始pH值对CBZ降解率的影响为pH 3.0pH 5.0pH 7.0;电压对CBZ降解率的影响为6V5V4V。  相似文献   

17.
In this study, the Fe(0)/CO(2) process was investigated for removing nitrate from aqueous solution under different operating conditions such as CO(2) bubbling rate (0-400 mL/min), Fe(0) dosage (1-6g/L), initial nitrate concentration (6-23 mgN/L), batch mode, and fresh Fe(0) supplementing (0-1g/L). Results show that the bubbling of CO(2) flow rate at 200 mL/min was sufficient for supplying H(+) into solution to create an acidic environment favorable to nitrate reduction reaction. It was found that sigmoidal model equation describes the S-curve behaviors of nitrate reduction, ferrous accumulation and ammonium formation satisfactorily, and the parameter t(1/2) of the proposed model equation serves as a powerful tool for the comparison of nitrate reduction rate. Sustainability test demonstrates that Fe(0) powder began to deteriorate after three batches operation. Concerning the operating modes, the batch mode with the treated solution emptied and freshly refilled outperforms the one, which was operated by retaining the treated solution and spiking concentrated nitrate into it for the next batch treatment. To guarantee satisfactory nitrate removal using the former mode, supplement of appropriate amount of Fe(0) needs to be optimized.  相似文献   

18.
Effective and economical removal of selenium (Se) in agricultural drainage water is very important in Se bioremediation. Zero-valent iron (ZVI) and a redox mediator [anthraquinone-2,6-disulfonate (AQDS)] were assessed for their ability to enhance the removal of Se(VI) or Se(IV) (500 µg/L) in synthetic drainage water by Enterobacter taylorae. The results showed that E. taylorae was capable of using inexpensive sucrose to remove Se from the drainage water. During a 7-day experiment, Se(VI) was almost entirely reduced to Se(0) and transformed to organic Se in the drainage water with sucrose levels of 500 to 1000 mg/L. Addition of ZVI to the drainage water increased the removal of total soluble Se to 94.5-96.5% and limited the production of organic Se. Addition of AQDS to the drainage water with or without ZVI decreased Se(VI) removal, but enhanced the removal of Se(IV), suggesting that E. taylorae only can use anthrahydroquinone-2,6-disulfonate (AHQDS, a reduced form of AQDS) to respire Se(IV), and not Se(VI). These results show that ZVI has promising application potential in the bioremediation of Se in Se-contaminated water.  相似文献   

19.
Exploring the cause‐and‐effect relationship between economic sectors and water resources is important to China. This study implemented a factor decomposition analysis by weighted average decomposition (WAD) model on the changes of Chinese virtual water (VW) consumption between 2002 and 2007, which includes both direct water consumption (consumed to produce final products) and indirect water consumption (consumed to produce intermediate products). The change in VW consumption is decomposed into three determinant factors: technological effect, economic structural effect and the products' scale effect. The results show that the volume of VW consumption in China has decreased from 5.92 × 1011 m3 in 2002 to 5.17 × 1011 m3 in 2007, which is mainly because of the technological effect (?5.48 × 1011 m3). The increase in net VW exports is mainly due to the economic structure effect (6.19 × 109 m3) and the fast growth of exports (3.49 × 1010 m3).  相似文献   

20.
Wang YH  Wong PK 《Water research》2005,39(9):1844-1848
A simple and rapid headspace method for gas chromatographic determination of dichloroacetic acid (DCAA) and trichloroacetic acid (TCAA) in drinking water was developed. Acidic methanol esterification followed by a headspace technique using a capillary column gas chromatograph (GC) equipped with an electron capture detector (ECD) was applied to determine the levels of DCAA and TCAA in drinking water. The major advantages of this method are the use of acidic methanol as the derivatization agent instead of the hazardous diazomethane, and esterification is carried out in water instead of organic solvent. DCAA and TCAA methyl esters produced in the reaction were determined directly by a headspace GC/ECD method. The linear correlation coefficients at concentrations ranging from 0 to 60 microg/L were 0.992 and 0.996 for DCAA and TCAA, respectively. The relative standard deviations (RSD, %) for the determination of DCAA and TCAA in drinking water were 15 and 21.3%, respectively (n=3). The detection limits of this method were 3 and 0.5 microg/L for DCAA and TCAA, respectively, and the recovery was 68-103.2% for DCAA and TCAA.  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