Вы находитесь на странице: 1из 11

WATER REUSE: REMOVAL OF TRACE ORGANIC COMPOUNDS

_____________________________________ Brian Bolto and Manh Hoang Abstract 29/8/11

The recycling of wastewater and stormwater is often cited as a viable alternative to seawater desalination. However, many undesirable organic compounds have been detected in waterways that receive wastewater effluents that are used as drinking water sources. The effects of ingestion of low levels of these compounds is not known. At present, no single treatment can remove all trace organic micropollutants. The various approaches are reviewed. Optimal treatment depends very much on the specific micropollutants present in the original wastewater. More needs to be done on the prediction of membrane treatment efficiency where the classes of compounds most likely to be rejected can be defined. Introduction There are often calls for alternative options to desalination such as the recycling of wastewater and stormwater to receive more attention (Kaempf, 2011; Khan, 2011; Malawaraarachchi et al., 2011). Unfortunately, many undesirable biologically active organic compounds have been detected in waterways that receive wastewater effluents and are used as drinking water sources, via indirect potable reuse (Asano et al., 2007; Williams et al., 2007; Schfer et al., 2011). The effect of ingestion of low levels of these compounds is not clear, but they represent known or suspected health effects for consumers and the environment, and are therefore of increasing concern. Certainly they have reproductive impacts on animals (Eertmans, et al., 2003), two examples being the feminising of fish and the declining fertility of young alligators (Nghiem and Schfer, 2002; Snyder et al., 2006; Kimura et al., 2009; Yangali-Quintanilla et al., 2009). The compounds occur at bioactive concentrations in the ng/L and low g/L range (Kolpin et al., 2002), and include: endocrine disruptors pharmaceutically active compounds from human and animal use insecticides and herbicides personal care products (surfactants, chelating agents) illicit drugs (Zuccato et al., 2008). A study has been carried out on wastewater plants utilising two different wastewater treatment technologies, activated sludge and trickling filter beds (Kasprzyk-Hordern et al., 2009). The impact of treated wastewater on the quality of receiving waters has also been assessed, and bioactive compounds found to be present at high loads, reaching 10 kg per day in the raw sewage. The efficiency of their removal was found to be strongly dependent on the technology used. Thus a trickling filter plant resulted in less than 70% removal of the compounds, while activated sludge treatment gave a much higher removal efficiency of over 85%. A monitoring program revealed that treated wastewater effluents were the main contributors to bioactive compounds in rivers, at up to 3 kg per day. As the effluents were major contributors to river flows, with dilution factors not exceeding 23, the effect of the effluent on the quality of the river water was significant.

Conventional treatment of river and other source waters when they are the feed for municipal water treatment installations is not particularly successful in removing these micropollutants. The area of removal of trace contaminants by a whole range of technologies has been extensively reviewed (Asano et al., 2007). The present article summarises the various techniques available, and gives more details of the most recent advances, with a particular emphasis on membrane methods. Conventional Treatment A recent evaluation has been made of the removal of pharmaceuticals, including psychiatric drugs, angiotensin agents, antihistamines, -blockers, other cardiac drugs, and endocrine disrupting compounds (EDCs) after various treatments at a full-scale drinking water treatment plant (Huerta-Fontela et al., 2011). The plant used raw water from the Llobregat River in Spain, which had total organic carbon (TOC) levels ranging from 2.7 to 10 mg/L over the four months of the study. The treatment steps were break-point chlorination coagulant (Al2[SO4]3, AlCl3, Al2O3) and flocculant (poly-DADMAC) addition sand filtration 6-29% dilution with ground water ozone granular activated carbon (GAC) filtration post chlorination to a residual of 0.8-1.2 mg/L chlorine. Coagulation, flocculation and sand filtration gave poor removal, but the oxidation processes were better, taking out 20 of the 35 compounds studied. GAC was effective in adsorbing very hydrophobic material, so that overall 30 of the compounds could be removed to 99.7% efficiency from original amounts of above 4 mg/L. Nevertheless, five compounds were detected at trace levels in the finished water: phenytoin, atenolol, sotalol, hydrochlorothiazide and carbamazapine epoxide. Ozone and Ozone-Hydrogen Peroxide The use of ozone as a method for removing trace contaminants has been reviewed (Asano et al., 2007) and surveyed at pilot plant level, where 50-98% could be eliminated (Ternes et al., 2003). Advanced oxidation methods (O3/H2O2, O3/UV) slightly increased removals in some cases. Thus the removal of iopamidol was raised from 57% to 80% with the addition of peroxide when the ozone dose was 10 mg/L, and from 84% to 88% with UV when the ozone dose was15 mg/L. It was recommended that an economic assessment be carried out before full scale implementation, bearing in mind that an added advantage is the inactivation of microorganisms. The kinetics and product formation of the anti-epileptic drug carbamazepine during reactions with ozone have been investigated (McDowell et al., 2005). This compound passes through sewage treatment plants essentially unchanged and in Germany can be found in many rivers and waterways at levels averaging 250 ng/L. It was discovered that ozone reacts rapidly with the double bond in carbamazepine. Rhine River water when spiked with 500 ng/L carbamazepine was found to have the additive completely oxidised by ozone doses of 0.3 mg/L. Similar results were obtained for spiked Lake Zurich water. The ozone oxidation of EDCs and pharmaceuticals has been the topic of a recent study (Snyder et al., 2006). Ozone reacts with organic contaminants by either the direct action of O3 or free radicals that are formed, especially the hydroxyl radical .OH. Molecular O3 reacts quickly with amines, phenols and the double bonds in aliphatic compounds, but the .OH radical is faster and less selective. Surface water was spiked with the target compounds which included estrogenic and androgenic steroids, pharmaceuticals, pesticides, and industrial chemicals. Full-scale water treatment plants were sampled before and after ozonation to determine if bench- and pilot-scale results could accurately predict removal in real life systems. The majority of contaminants were more than 90% removed at O3 levels commonly used for 2

disinfection. Atrazine, iopromide, meprobamate, and tris-chloroethylphosphate were the most recalcitrant, with removals generally less than 50%. The addition of H2O2 for advanced oxidation was of little benefit compared to O3 alone. However, the combination provided a marginal increase in the removal of dilantin, diazepam, iopromide, and meprobamate, but decreased the removal efficacy of pentoxifylline, caffeine, testosterone, progesterone, and androstenedione. In wastewater experiments, O3 and O3/H2O2 were shown to remove estrogenicity to below the detection limit. It was concluded that, overall, ozone is a highly effective oxidant for removing the majority of trace organic contaminants. Challenging compounds will require greater doses of these or other oxidants. No single oxidant can remove all trace organic micropollutants to below the detection limit. Other methods of enhancing the production of .OH radicals include sonochemistry, which is effective as a way of increasing the generation of .OH radicals to degrade recalcitrant organic compounds (Pang et al., 2011). Ultraviolet Radiation-Hydrogen Peroxide An advanced oxidation process in the form of ultraviolet radiation-hydrogen peroxide (UV-HP) has been used as a way of treating source waters containing natural organic matter (NOM) contaminated with a typical synthetic organic chemical and EDC in the form of the pesticide alachlor (Song et al., 2008). Alachlor was chosen as a model compound because of its hydrophobicity and its tendency to complex with NOM. It was discovered that although its decomposition was significantly hindered by the presence of NOM, the process was effective in destroying alachlor in NOM-containing waters because of the high reactivity of the hydroxyl radical with alachlor. Specific Adsorbents A novel method of selective extraction of micropollutants uses an artificial molecular receptor specifically designed for the target molecules (Le Noir et al., 2007). The polymer particles were made from 4-vinylpyridine crosslinked with ethylene glycol dimethacrylate, which were polymerised around 17-estradiol as the template. Treatment of a 2 ng/L solution of 17estradiol gave 100 0.6% removal, versus 77.3 3.7% when the polymer particles were made in the absence of a template, which shows the potential of the method. Lower removals were achieved with GAC. The adsorbent could be regenerated with solvent, methanol in this case. The progress of research on various polymers such as porous particles and imprinted polymer particles as adsorbents for EDCs has been reviewed recently (Suna et al., 2011). Membrane Processes Nanofiltration Nanofiltration (NF) technology for trace contaminant removal has been reviewed comprehensively (Nghiem and Schfer, 2005; Hofman et al., 2006). While size exclusion can be responsible for separation, several studies have shown that even large molecules such as EDCs can pass through some membranes (Kiso et al., 2000, 2001). An adsorption diffusion mechanism can occur. This has been broken down into physical selectivity: charge repulsion, size exclusion or steric hindrance; and chemical selectivity: solvation energy, hydrophobic interaction or hydrogen bonding. Thus membranes may adsorb estrone, which will result in an initially high retention. Hydrogen bonding has been postulated as the mechanism of adsorption of this compound (Nghiem and Schfer, 2002). If the pore size of the membranes is larger than the estrone molecules, breakthrough may be observed when the adsorptive sites are saturated. In all, eight different NF membranes or the tighter reverse osmosis (RO) membranes were tested; the RO membranes had much smaller pore sizes that could not be penetrated by estrone. No

breakthrough was observed with these membranes. A review of mechanisms for estrogens has recently been published (Schfer et al., 2011). The rejection of trace organic compounds by a variety of NF membranes under a range of experimental conditions has been reviewed in an extensive report that covers 20 scientific papers (Hofman et al., 2006). Results for approximately 120 different compounds of various types, not just EDCs and pharmaceuticals, were assessed for 30 different membrane types, resulting in a data base containing more than 1,000 rejection measurements. A statistical analysis showed that 60% of the rejections were above 80%. It was concluded that hydrodynamics play an important role in the rejection performance of spiral wound membranes. Also, that convective transport through the NF membranes was important, and that on the full scale rejections can drop significantly at high recoveries of above 80%. A model was developed that can be used for the design of full-scale plants. Some particular findings were Rejections of small neutral organic compounds such as alcohols and sugars correlated well with molecular weight (MW), with the transition from low to high rejection occurring near the MW cut off of the membrane at ~200 Da Rejections of more polar or partly dissociated small organic acids also show a good correlation with MW, but with the transition occuring below the MW cut off Rejection is influenced by the adsorption of organic compounds, which may take 3 or 4 days to reach equilibrium before a stable rejection takes place. Some compounds such as alkyl phenols have a high affinity for the membrane so that they are completely adsorbed Phenyl urea herbicides were a special case in that with methoxy methyl urea groups there is a decrease in rejection when the MW is increased. While NF is successful in removing larger pollutants, it is claimed that it can also tackle smaller hydrophilic and/or charged organic micropollutants (Verliefde et al., 2007). Because of the polar nature of some micropollutants such as nitrosodimethylamine (NDMA) and methyl tert-butyl ether (MTBE), they are not completely removed by conventional water treatment, including GAC. Qualitative prediction of NF rejection has been achieved by using several solute parameters, such as log Kow (the octanol-water partition coefficient), pKa and molar mass (although it is molecular size that is the issue) to predict removal efficiency. Polarity is important, as solutes with a high dipole moment may align in the direction of the pores of the membrane because of electrostatic interactions with the membrane charge. This causes the solutes to permeate more easily, so they have a lower rejection. Polarity can also be a function of hydrophobicity, and hydrophobic interactions between solute and membrane can result in adsorption of the compounds on the membrane surface and in the pores. The higher the hydrophobicity of a compound, expressed by log Kow, the greater the likelihood of adsorption on the membrane. This can result in a higher initial rejection. Solution of the solute in the membrane and diffusion through the membrane are also contributors to transport mechanisms. The more a compound adsorbs onto the membrane, the easier it will dissolve in the membrane, to be transported to the other side. A higher log Kow should thus lead to a lower rejection. Rejection will obviously also be affected by the size of the solute molecule, and can be explained as a steric hindrance phenomenon. This applies to hydrophilic compounds as they are expected to be more hydrated, increasing the effective size of the molecule, so that there is better rejection than for non-hydrated species. The relative influence of hydrophobicity is hence more important than molecular mass in determining rejection in NF. Comparison with experimental data in the literature confirmed that the set of parameters used is adequate to give a rough estimate of the likely removal efficiency. Some 17 compounds were surveyed in thus way, the details being shown in Table 1. Membrane pore data are as provided by the membrane manufacturer. The many references for the experimental values, resulting in some cases several sets of results for the one compound, are given in the original paper (Verliefde et al., 2007). The predicted rejection of MTBE is moderate, and the found value is 89.6%. The prediction for NDMA is the same, but there are no experimental data to date. 4

Table 1. Predicted and experimental rejection of micropollutants in NF (Verliefde et al., 2007) Micropollutant Hormones 17-Estradiol Estrone Progesterone Testosterone Industrial chemicals Bisphenol A p-Dimethylphthalate p-Diethylphthalate Nonylphenol MTBE Pesticides Atrazine Simazine Chlorpyrifus Pharmaceuticals Primidone Carbamazepine Ibuprofen Log Kow 4.01 3.13 3.87 3.32 Molecular Weight 272 270 314 288 Predicted Rejection Found Rejection

High, then moderate/high High, then moderate/high High, then moderate/high High, then moderate/high

85-100 0-60, 85-100 90-100 80-100

3.3 2.2 3.2 4.4 0.9

228 194 222 220 88

High, then moderate/high High, then moderate/low High, then moderate High, then moderate Moderate

70-100, 45 65-80 65-80 70-90 89.6

2.6 2.1 5.2

216 202 350

High, then moderate High, then moderate High

26-99 17-96 >99

0.9 2.4 3.9

218 236 206

Moderate/high High, then moderate/high High

72-87 93 70-95

Information on pKa, the dissociation constant of micropollutants, is not always available; Table 2 summarises what is known. One investigation of the contribution of electrostatic interactions to the rejection of selected organic acids by NF membranes has shown that the rejection of negatively charged carboxylic acids was larger than expected from steric exclusion and was primarily driven by the surface charge of the membrane and the degree of ionisation of the acid (Bellona and Drewes, 2005). At pH levels below its pKa, ibuprofen was less than 50% charged and, being hydrophobic, is adsorbed on the membrane. At pH levels above the pKa, adsorption of ibuprofen was less with increasing pH because of the increase in the amount of the anionic form, which has a strong electrostatic repulsion influence. However, this effect does not appear to be as strong as hydrophobic interactions. Rejection was at a minimum at pH 5. It was found that ibuprofen was predominately removed by adsorption and above pH 5 it was removed by electrostatic repulsion. A tighter membrane gave higher rejections. The radius of the ibuprofen molecule is 0.38 nm and the pore radii of NF membranes used in another study were 0.8-1.4 nm (Park and Choi, 2005). Hence rejection of this pharmaceutical compound has to be by a mechanism other than one involving molecular size, confirming that the high negative charge on the molecule at pH 7 and the negative character of the membranes are responsible for repulsion of the micropollutant.

Table 2. Predicted and experimental rejection of micropollutants in NF as a function of pKa (Verliefde et al., 2007; Bellona and Drewes, 2005; Albert and Serjeant, 1962) Micropollutant Nonylphenol Atrazine pKa 10.25 1.7 Charge at pH 7 6% anionic 99.9% cationic Category 3-4 3-4 Predicted Rejection High, then moderate High, then moderate Found Rejection 70-90 68-98 95 81-99 >80 26-97 88-93 >75 17-96 70-95* 70-83

Simazine

1.62

99.9% cationic 99.2% anionic

3-4

High, then moderate

Ibuprofen

4.91

7-8

High because of hydrophobic interactions and charge repulsion

*pH <4.9; pH >4.9


Because conventional coagulation and adsorption on GAC are inefficient in the removal of estrogens, there are many examples that use further treatment involving membranes (Bodzek and Dudziak, 2006). NF has been applied in this way with polyamide and cellulosic commercial membranes. The membranes alone removed >63% of a mixture of six natural and synthetic estrogens, but after coagulation with poly(aluminium chloride) this was elevated to >81.5% removal. Depending on the type of estrogen, with the polyamide membrane the retention coefficients ranged from 81.5% for estriol to 100% for mestranol. Much higher values were obtained for synthetic estrogens, which have a lower solubility in water. Of course the role of the coagulant is to simply insolubilise the micropollutants, so a better coagulant is needed for this purpose. A specifically designed inorganic coagulant should have some hydrophobic surfaces too. A recent example is an organic-modified polysilicato iron made for treating dye wastewater (Fu et al., 2011). A recent study has been published of the interactions between trace organics, NF membranes, fouling layers, and NOM macromolecules present in raw water (Hajibabania et al., 2011). The adsorption of organics onto the membrane surface and their association with NOM were shown to be dependent on the physicochemical properties of the organics and the molecular weight of the NOM. Model compounds were used to simulate the effect of the nature of the foulant on adsorption behaviour and fouling. The characteristics of the NOM influenced the amount of adsorption of trace organic compounds. A decrease in rejection was observed with alginate- and humic acid-fouled membranes. Some groups bound more with alginate rather than humic acid, and this enhanced adsorption could be closely correlated with the number of functional groups and the molecular size of the NOM. It was found that two counteractive mechanisms were commonly involved in the rejection of hydrophilic non-ionic solutes: adsorption onto the NOM resulting in an increased rejection, and the presence of cake-enhanced concentration polarization leading to a decrease in rejection. The negative effect of fouling on the removal of trace organics was more dominant than the influence of adsorption onto the NOM. The effect of fouling on the rejection of hydrophobic trace organics was more limited. Their rejection was strongly influenced by the surface charge densities of the fouled membranes, since a decrease in rejection was observed with alginate- and humic acid-fouled membranes. For hydrophobic ionic compounds, two mechanisms were again involved: electrostatic repulsion of the compounds with the membrane surface, and to a lesser extent, adsorption onto NOM in the bulk and on the 6

fouling layer. It was hoped that once the properties of the main trace organics were properly defined, along with the characteristics of the feed water, long term rejection performances of NF membrane could be better predicted. In practical applications of NF/RO membranes to municipal wastewater treatment, the feed water always contains organic macromolecules at TOC levels of up to 10 mg/L (Kimura et al., 2009). These are mainly soluble microbial products produced during biological treatment processes such as an activated sludge process. A study of the influence of these organic macromolecules on the removal of six pharmaceuticals by NF/RO membranes was undertaken. Two types of biological treatment, conventional activated sludge followed by media filtration (tertiary treatment) and a membrane bioreactor (MBR), were examined as sources of feed waters to NF/RO membranes. Removal of the pharmaceuticals was higher from these feed waters than from deionised pure water spiked with the pharmaceuticals, with the increase being significant in the case of the NF membrane. Alteration of the membrane surface because of membrane fouling and association of the pharmaceuticals with organic macromolecules were postulated to account for this increase in removal. It was proposed that the organic macromolecules present in tertiary effluent enhanced removal of the pharmaceuticals by NF because of a modification of the membrane surface. However, the organic macromolecules present in the MBR effluent seemed to enhance removal of the pharmaceuticals by NF because of association with them. The different mechanisms highlight the different properties of the organic macromolecules present in the two types of effluent. The prediction of membrane rejections based on a detailed understanding of organic compound rejection levels as a function of the properties of the compounds and the membrane is a formidable task. An alternative way is to develop high-science models of quantitative structureproperty relations (QSPR) that take into account the simultaneous correlation of organic compound rejection with multiple molecular parameters for the membrane. Artificial neural networks have this capability for building multi-parameter QSPRs with wide applicability (Libotean et al., 2008). Such models have been developed using the results of RO performance covering 50 different organic compounds and five different commercial RO membranes at the Orange County Water District of Southern California. Organic adsorption and passage experiments demonstrated that they can be qualitatively and quantitatively related to chemical structure. The most significant molecular descriptors characterising the adsorbed organics included the size of the smallest ring, the dipole moment, the dipole hybridization and the heat of formation. For the passage of organics through a membrane, the most relevant molecular descriptors were the size of the smallest ring, the molecular weight, the shape index and the energy gained when an electron is added to the lowest unoccupied molecular orbital. Predictions were consistent with the experimental observation that higher organics rejection and lowest passage occur with polyamide membranes compared to cellulose acetate membranes. It was felt that with expanded data bases this quite fundamental approach could provide a suitable forecasting system for organics/RO membrane behaviour. Summary of NF Studies The estimated outcomes for micropollutant removal by NF are shown in Table 3, which lists eight categories of behaviour (Verliefde et al., 2007). Charge effects are the most dominant, with hydrophobic and size characteristics having moderate effects.

Table 3. Predicted rejection of organic micropollutants in NF (Verliefde et al., 2007) Category 1 2 3 4 5 6 7 8 Hydrophobicity (logKow) <2* <2 >2 >2 <2 <2 >2 >2 Molecular Size < pore size > pore size < pore size > pore size < pore size > pore size < pore size > pore size Charge Status Uncharged Rejection Mechanism Steric hindrance Uncharged Steric hindrance Uncharged Hydrophobic interaction Uncharged Hydrophobic interaction Charged Charge repulsion Charged Charge repulsion Charged Charge repulsion Charged Charge repulsion Rejection Prediction Moderate Moderate to high Low Moderate High Very high High Very high

* Hydrophilic; Hydrophobic

Photocatalytic Membrane Reactors There is a question as to whether further advanced techniques should be devised. Photolysis is an option (Asano et al., 2007). Photocatalytic NF reactors have been suggested as worthwhile (Molinari et al., 2002a, 2005). The possibility of using sunlight makes the approach a promising one. Results obtained by using various configurations of photocatalytic membrane reactors have been reported. The configurations studied were: irradiation of the cell containing the membrane, with either the catalyst deposited on the membrane catalyst in suspension, confined by the membrane or entrapment of the photocatalyst on a membrane irradiation of the re-circulation tank and catalyst in suspension, confined by the membrane. The catalysts employed include TiO2, ZnO, WO3, CdS, MoS2, CdSe and Fe3O4. The most widely used is TiO2, in both the anatase and rutile forms, with anatase being the most photoactive. In work on 4-nitrophenol with a sulphonated polyether sulphone NF membrane and TiO2, three factors were found to be important in obtaining a very low value of <2.5 mg/L in the permeate, using a 40 mg/L feed (Molinari et al., 2002b). They were rejection, photocatalytic degradation and adsorption. Rejection was significantly lower than that obtained in the absence of photodegradation, most likely because of the small molecular size of the products of the photodegradation process. The pressure in the membrane cell, 600 kPa in this case, the pH of the polluted water and the molecular size of the pollutants can influence the permeate flux of the membrane, which ranged from 20 to 40 L/m2h. Membrane Bioreactors The MBR technique has been outlined in the major reuse review (Asano et al., 2007). The removal of 40 trace organic contaminants from a synthetic sewage by a laboratory MBR has 8

aimed at determining the relationship between molecular structure and ease of biological removal. (Tadkaew et al., 2011). The reactor was seeded with activated sludge from a local sewage works. There was a better than 85% removal of 14 very hydrophobic compounds. All hydrophilic and moderately hydrophobic compounds having strong electron withdrawing functional groups such as carboxylic acid, amides or halogen groups resulted in less than 20% removal efficiencies. On the other hand, most compounds having electron donating groups such as hydroxyl, primary amino, ether, alkyl or ester groups were removed to better than 70% efficiency, although there were several exceptions which were inexplicable. An even more recent reuse study has examined the on-site evaluation of the removal of 100 micropollutants, including some pharmaceuticals, using seven advanced wastewater treatment processes, one an MBR (Ruel et al., 2011). For 18 of the compounds the removal efficiencies were significantly higher than those obtained with activated sludge plants. Conclusions The main message from this review is that no single treatment can remove all trace organic micropollutants. Optimal treatment will very much depend on the specific micropollutants in the original wastewater, so knowledge of what is present initially is required. A plausible combination could be a membrane bioreactor, followed by an oxidative procedure, then NF. There is scope for the development of new approaches, especially oxidation methods, and for better predicting rejection behaviour in NF. References Albert, A. and Serjeant, E. P. (1962). Ionisation Constants of Acids and Bases. Methuen, London, pp. 173-175. Asano, T., Burton, F. L., Leverenz, H. L., Tsuchihashi, R. and Tchobanoglous, B. (2007). Water Reuse: Issues, Technologies and Applications. McGraw Hill, New York. Bellona, C. and Drewes, J. E. (2005). The role of membrane surface charge and solute physicochemical properties in the rejection of organic acids by NF membranes. J. Membrane Sci. 249, 227-234. Bodzek, M. and Dudziak, M. (2006). Elimination of steroidal sex hormones by conventional water treatment and membrane processes. Desalination 198, 24-32. Eertmans, F., Dhooge, W., Stuyvaert, S. and Comhaire, F. (2003). Endocrine disruptors: effects on male fertility and screening tools for their assessment. Toxicology in Vitro 17, 515524. Fu, Y., Gao, B., Zhang, Y., Zhang, X. and Shi, N. (2011). Organic modifier of poly-silicic-ferric coagulant: Characterization, treatment of dyeing wastewater and floc change during coagulation. Desalination 277, 6773. Hajibabania, S., Verliefde, A., McDonald, J. A., Khand, S. J. and Le-Clecha, P. (2011). Fate of trace organic compounds during treatment by nanofiltration. J. Membrane Sci. 373, 130-139. Hofman, J. A. M. H., Gijsbertsen, A. J. and Cornelissen, E. (2006). Nanofiltration retention models for organic contaminants. Report for Project 2945, Water Research Foundation, Denver. http://waterrf.org/ProjectsReports/PublicReportLibrary/2945.pdf Huerta-Fontela, M., Galceran, M. T. and Vetura, F. (2011). Occurrence and removal of pharmaceuticals and hormones through drinking water treatment. Water Research 45, 14321442. Kasprzyk-Hordern, B., Dinsdale, R. M. and Guwy, A. J. (2009). The removal of pharmaceuticals, personal care products, endocrine disruptors and illicit drugs during wastewater treatment and its impact on the quality of receiving waters. Water Research 43, 363-380. Kaempf, J. (2011). The case against desalination. Water Journal 38(3), 39-40. Khan, S. J. (2011). The case for direct potable reuse in Australia. Water Journal 38(4), 92-96. 9

Kimura, K., Iwase, T., Kita, S. and Watanabe, Y. (2009). Influence of residual organic macromolecules produced in biological wastewater treatment processes on removal of pharmaceuticals by NF/RO membranes. Water Research 43, 3751-3758. Kiso, Y., Nishimura, Y., Kitao, T. and Nishimura, K. (2000). Rejection properties of nonphenylic pesticides with nanofiltration membranes. J. Membrane Sci. 171, 229-237. Kiso, Y., Sugiura, Y., Kitao, T. and Nishimura, K. (2001). Effects of hydrophobicity and molecular size on rejection of aromatic pesticides with nanofiltration membranes. J. Membrane Sci. 192, 1-10. Kolpin, D. W., Furlong, E. T., Meyer, M. T., Thurman, E. M., Zaugg, S. D., Barber, L. B. and Buxton, H. T. (2002). Pharmaceuticals, hormones, and other organic wastewater contaminants in US streams. Environ. Sci. Technol. 36, 1202-1211. Le Noir, M., Lepeule, A.-S., Guieysse, B. and Mattiasson, B. (2007). Selective removal of 17estradiol at trace concentration using a molecularly imprinted polymer. Water Research 41, 2825-2831. Libotean, D., Giralt, J., Rallo, R., Cohen, Y., Giralt, F., Ridgeway, H. F., Rodriguez, G. and Phipps, D. (2011). Organic compounds passage through RO membranes. J. Membrane Sci. 313, 23-43. Malawaraarachchi, M. G., OToole, J., Sinclair, M. and Leder, K. (2011). Review of greywater use in Australia. Water Journal 38(4), 97-102. McDowell, D., Huber, M., Wagner, M., von Gunten, U. and Ternes, T. (2005). Ozonation of carbamazepine in drinking water: identification and kinetic study of major oxidation products. Environ. Sci. Technol. 39, 8014-8022. Molinari, R., Palmisano, L., Drioli, E., and Schiavello, M. (2002a). Studies on various reactor configurations for coupling photocatalysis and membrane processes in water purification. J. Membrane Sci. 206, 399415. Molinari, R., Borgese, M., Drioli, E., Palmisano, L. and Schiavello, M. (2002b). Hybrid processes coupling photocatalysis and membranes for degradation of organic pollutants in water. Catalysis Today 75, 7785. Molinari, R., Giorno, L., Drioli, E., Palmisano, L. and Schiavello, M. (2005). Photocatalytic nanofiltration reactors. In: Nanofiltration Principles and Applications, (Ed. A. I. Schfer, A. G. Fane and T. D. Waite), Elsevier, Oxford, pp.435-458. Nghiem, L. D. and Schfer, A. I. (2002). Adsorption and transport of trace contaminant estrone in NF/RO Membranes. Environ. Eng. Sci. 19, 441-451. Nghiem, L. D. and Schfer, A. I. (2005). Trace contaminant removal with nanofiltration. In: Nanofiltration Principles and Applications, (Ed. A. I. Schfer, A. G. Fane and T. D. Waite), Elsevier, Oxford, pp.479-520. Pang, Y. L., Abdullah, A. Z. and Bhatia, S. (2011). Review on sonochemical methods in the presence of catalysts and chemical additives for treatment of organic pollutants in wastewater. Desalination 277, 1-14. Park, G.-Y. and Choi, J. (2005). Transport of pharmaceutical and NOM in NF and tight UF membranes. Proc Membrane Technol. Conf., Phoenix, Amer. Water Works Assoc. Ruel, S. M., Choubert, J. M., Esperanza, M., Mige, C., Madrigal, P. N., Budzinski, H., Le Mnach, K., Lazarova, V. and Coquery, M. (2011). On-site evaluation of the removal of 100 micro-pollutants through advanced wastewater treatment processes for reuse applications. Water Sci. Technol. 63, 2486-2497. Schfer, A. I., Akanyeti, I. and Semio, A. J. C. (2011). Micropollutant sorption to membrane polymers: A review of mechanisms for estrogens. Advances in Colloid and Interface Science 164, 100117. Snyder, S. A., Wert, E. C., Rexing, D. J., Zegers, R. E. and Drury, D. D. (2006). Ozone oxidation of endocrine disruptors and pharmaceuticals in surface water and wastewaters. Ozone Sci. Eng. 28, 445-460.

10

Song, W., Ravindran, V. and Pirbazari, M. (2008). Process optimisation using a kinetic model for the ultraviolet radiation-hydrogen peroxide decomposition of natural and synthetic compounds in groundwater. Chem. Eng. Sci. 63, 3249-3270. Suna, S., Hunaga, J. and Zhaoa, C. (2011). Polymeric particles for the removal of endocrine disruptors. Sep. & Purif. Rev. 40, 312-337. Tadkaew, N., Hai, F. I., McDonald, J. A., Khan, S. J. and Nghiem, L. D. (2011). Removal of trace organics by MBR treatment: The role of molecular properties. Water Research 45, 24392451. Ternes, T. A., Stber, J., Herrmann, N., McDowell, D., Ried, A., Kampmann, M. and Teiser, B. (2003). Ozonation: a tool for removal of pharmaceuticals, contrast media and musk fragrances from wastewater? Water Research 37, 1976-1982. Verliefde, A., Cornelissen, E., Gary Amy, G., Van der Bruggen, B. and van Dijk, H. (2007). Priority organic micropollutants in water sources in Flanders and the Netherlands and assessment of removal possibilities with nanofiltration. Environ. Pollution 146, 281-289. Williams, M., Woods, M., Kumar, A.,Ying, G. G., Shareef, A., Karkkainen, M. and Kookana, R. (2007). Endocrine disrupting chemicals in the Australian riverine environment: A pilot study on estrogenic compounds. Report PR071403, Land & Water Australia, Canberra. Yangali-Quintanilla, V., Sadmani, A., McConville, M., Kennedy, M. and Amy, G. (2009). Rejection of pharmaceutically active compounds and endocrine disrupting compounds by clean and fouled nanofiltration membranes. Water Research 43, 2349- 2362. Zuccato, E., Castiglioni, S., Bagnati, R., Chiabrando, C., Grassi, P. and Fanelli, R. (2008). Illicit drugs, a novel group of environmental contaminants. Water Research 42, 961-968.

Dr Brian Bolto (email: brian.bolto@csiro.au) is a Visiting Scientist and Dr Manh Hoang (email: manh.hoang@csiro.au) is Leader of Advanced Water Treatment, both at CSIRO Materials Science and Engineering, located at Clayton, Victoria (postal address: Private Bag 33, Clayton South, Vic.

11

Вам также может понравиться