Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Endocrine disrupting compounds removal from wastewater, a new challenge

Process Biochemistry, 2006
...Read more
Endocrine disrupting compounds removal from wastewater, a new challenge Muriel Auriol a,c , Youssef Filali-Meknassi b,c , Rajeshwar D. Tyagi a, * , Craig D. Adams c , Rao Y. Surampalli b a University of Quebec, INRS-ETE, 490 de la Couronne, QC, Canada G1K 9A9 b U.S. EPA, P.O. Box 17-2141, Kansas City, KS 66117, USA c University of Missouri-Rolla, Civil Engineering Department, 1870 Miner Circle Rolla, MO 65409-1060, USA Received 27 July 2005; received in revised form 16 September 2005; accepted 29 September 2005 Abstract Various natural chemicals and some contaminants of industrial source present an endocrine activity. Nowadays, many questions related to these compounds are not resolved and the persistent character of these compounds makes it a major problem for future generations. This study concentrated on some specific groups of endocrine disrupting chemicals (estrogens and alkylphenols). In this review, a number of treatment processes will be discussed with regard to their potential on endocrine disrupting chemicals removal. # 2005 Elsevier Ltd. All rights reserved. Keywords: Endocrine disrupting chemicals; Estrogens; Alkylphenols; EDC; Wastewater; Hormone 1. Introduction The human growth, development coordination and matura- tion imply a complex interaction of hormonal signals whose chronology and dose can have permanent consequences on the future form and function of many tissues [1,2]. Human exposure to very low doses during critical periods, for example at the cellular differentiation period, can alter the development course of these tissues and this may result in permanent character changes in the mature living beings [1,2]. Considering the complexity of endocrine systems, it is not surprising that a wide range and varied substances cause endocrine disruption and these include both natural and synthetic chemicals [3,4]. Indeed, according to an European Union study, 118 substances were classified as potential endocrine disrupters (EDCs); and a peculiar pri- ority was assigned to the carbon disulfide, o-phenylphenol, tetrabrominated diphenyl ether, 4-chloro-3-methylphenol, 2,4- dichlorophenol, resorcinol, 4-nitrotoluene, 2,2 0 -bis(4-(2,3- epoxypropoxy)phenyl)propane, 4-octylphenol, estrone (E1), 17a-ethinylestradiol (EE2), and 17b-estradiol (bE2) [5]. EDCs are often dominant and can disperse quickly in the environment. EDCs are released to the atmosphere as a result of combustion and incineration activities (polycyclic aromatic hydrocarbons (PAHs), dioxins) [6], but the principal sinks for EDCs are groundwater, river, and lakes [7]. The four main classes of EDCs (natural steroidal estrogens, synthetic estrogens, phytoestrogens, and various industrial chemicals) are generally represented with respect to their estrogenic potency [8]. The natural and synthetic estrogens generally display much stronger estrogenic effects than the phyto- and xenoestrogens. However, the concentrations of phyto- and xenoestrogens in the aquatic environment are usually higher [9]. The list of trace contaminants or EDCs, resulting from human activities and found in wastewater, is long [10–12]. However, in general natural (E1, bE2, estriol [E3]) and synthetic (EE2, mestranol) hormones are the major contributors to the estrogenic activity observed in sewage effluents [13–15] and the receiving water. Recent research showed that several sewage treatment plant (STP) effluents and rivers in the United Kingdom [14,16–20] and in the United States [21,22] contain sufficient amount of estrogenic compounds to induce harmful effects on fish (Tables 1–3). Field studies using caged trout (Oncorhynchus mykiss), wild cyprinid roach (Rutilus rutilus) [40], and estuarine flounder (Platichthys flessus) [41,42] showed that the estrogenicity persists in receiving water and www.elsevier.com/locate/procbio Process Biochemistry 41 (2006) 525–539 * Corresponding author. Tel.: +1 418 654 2617; fax: +1 418 654 2600. E-mail address: tyagi@inrs-ete.uquebec.ca (R.D. Tyagi). 1359-5113/$ – see front matter # 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.procbio.2005.09.017
that the concentration of these compounds present in the rivers and the estuaries are high enough to induce deleterious reproductive consequences. The incidence of hermaphroditic wild fish near STPs initiated an investigation on STPs effluent estrogenicity. Caged fish held downstream of some STPs produced vitellogenin (VTG), indicating the presence of estrogenic substances [17,18,43]. In 1990, British scientists showed that male rainbow trout produced the yolk precursor protein VTG when they were exposed to sewage effluents or contaminated surface water [44]. Other studies have also shown that birds, reptiles, and mammals in polluted areas undergo alterations of the endocrine reproductive system [45]. Natural and synthetic estrogen hormones (such as bE2, E3, E1, and EE2) seem to be responsible for endocrine disruption in fish [13,28,46]. Indeed, several studies showed that even low concentrations (ng/L) of bE2 can induce VTG in male species and rainbow trout (O. mykiss) experimentally exposed to these chemicals [46,47]. Purdom et al. [16] and Hansen et al. [48] noticed that concentrations of bE2 as low as 1 ng/L induces VTG M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 526 Table 1 Estrogens concentrations in STP influent Sampling site Influent concentrations (ng/L) Analysis method Reference Estrone 17b-Estradiol Estriol 17a-Ethinylestradiol Paris, France 9.6–17.6 11.1–17.4 11.4–15.2 4.9–7.1 SPE/GC-MS [23] England 1.8–4.1 <0.3 <LOD a SPE/GC-MS-MS [24] Germany 66 22.7 SPE/LC-ESI-MS-MS [25] Italy 52 12 80 3 SPE/LC-MS-MS [26] Roma, Italy 31 9.7 57 4.8 SPE/LC-MS-MS [25] Roma, Italy 44 11 72 SPE/LC-ESI-MS-MS [27] Barcelona, Spain <2.5–115 <5–30.4 <0.25–70.7 <5 SPE/LC-MS [28] Japan 5 SPE/ELISA [29] ELISA: enzyme-linked immunosorbent assay; ESI: interface electrospray; GC-MS: gas chromatography-mass spectrometry; GC-MS-MS: gas chromatography- tandem mass spectrometry; LC-MS: liquid chromatography-mass spectrometry; LC-MS-MS: liquid chromatography-tandem mass spectrometry; LOD: limit of detection; SPE: solid phase extraction. a 0.3 ng/L. Table 2 Estrogens concentrations in STP effluent Samplig site Effluent concentration (ng/L) Analysis method Reference Estrone 17b-Estradiol 17a-Estradiol Estriol 17a-Ethinylestradiol Mestranol Paris, France 6.2–7.2 4.5–8.6 5.0–7.3 2.7–4.5 SPE/GC-MS [23] Denmark <2–11 <1–4.5 <1–5.2 [30] Netherlands <0.4–47 <0.6–12 <0.1–5 <0.2–7.5 SPE/GC-MS-MS [31] Sweden 5.8 1.1 4.5 SPE/GC-MS [32] England 1.4–76 2.7–48 <LOD a –4.3 SPE/GC-MS [13] England <LOD b <LOD b <LOD b SPE/GC-MS-MS [24] England 6.4–29 1.6–7.4 2–4 <LOD c SPE/GC-NCI-MS [33] Germany 9 <LOD d 1 <LOD d SPE/GC-MS-MS [34] Germany 14.6 4.6 SPE/LC-MS-MS [25] Germany 7 6 3 3 4 SPE/HRGC-MS [35] Germany (SW) <LOD e –18 <LOD f –15 <LOD f –12 <LOD g –2.7 SPE/GC-MS [9] Italy 3 1.4 20.4 0.6 SPE/LC-ESI-MS-MS [26] Roma, Italy 24 4 11.7 1.4 SPE/LC-MS-MS [25] Roma, Italy 17 1.6 2.3 SPE/LC-ESI-MS-MS [27] Centre Italy 5–30 3–8 n.d.–1 n.d. SPE/LC-ESI-MS-MS [36] Barcelona, Spain <2.5–8.1 <5–14.5 <0.25–21.5 <5 SPE/LC-MS [28] Japan <LOD d SPE/ELISA [29] Japan 2.5–34 0.3–2.5 SPE/LC-MS-MS [37] Canada 3 6 9 <LOD d SPE/GC-MS-MS [34] California, USA 0.2–4.1 0.2–2.4 SPE/ELISA [38] HRGC-MS: high resolution gas chromatography-mass spectrometry; NCI: negative chemical ionization. a 0.2 ng/L. b 0.3 ng/L. c 0.05 ng/L. d 1 ng/L. e 0.7 ng/L. f 0.4 ng/L. g 0.6 ng/L.
Process Biochemistry 41 (2006) 525–539 www.elsevier.com/locate/procbio Endocrine disrupting compounds removal from wastewater, a new challenge Muriel Auriol a,c, Youssef Filali-Meknassi b,c, Rajeshwar D. Tyagi a,*, Craig D. Adams c, Rao Y. Surampalli b a University of Quebec, INRS-ETE, 490 de la Couronne, QC, Canada G1K 9A9 b U.S. EPA, P.O. Box 17-2141, Kansas City, KS 66117, USA c University of Missouri-Rolla, Civil Engineering Department, 1870 Miner Circle Rolla, MO 65409-1060, USA Received 27 July 2005; received in revised form 16 September 2005; accepted 29 September 2005 Abstract Various natural chemicals and some contaminants of industrial source present an endocrine activity. Nowadays, many questions related to these compounds are not resolved and the persistent character of these compounds makes it a major problem for future generations. This study concentrated on some specific groups of endocrine disrupting chemicals (estrogens and alkylphenols). In this review, a number of treatment processes will be discussed with regard to their potential on endocrine disrupting chemicals removal. # 2005 Elsevier Ltd. All rights reserved. Keywords: Endocrine disrupting chemicals; Estrogens; Alkylphenols; EDC; Wastewater; Hormone 1. Introduction The human growth, development coordination and maturation imply a complex interaction of hormonal signals whose chronology and dose can have permanent consequences on the future form and function of many tissues [1,2]. Human exposure to very low doses during critical periods, for example at the cellular differentiation period, can alter the development course of these tissues and this may result in permanent character changes in the mature living beings [1,2]. Considering the complexity of endocrine systems, it is not surprising that a wide range and varied substances cause endocrine disruption and these include both natural and synthetic chemicals [3,4]. Indeed, according to an European Union study, 118 substances were classified as potential endocrine disrupters (EDCs); and a peculiar priority was assigned to the carbon disulfide, o-phenylphenol, tetrabrominated diphenyl ether, 4-chloro-3-methylphenol, 2,4dichlorophenol, resorcinol, 4-nitrotoluene, 2,20 -bis(4-(2,3epoxypropoxy)phenyl)propane, 4-octylphenol, estrone (E1), 17a-ethinylestradiol (EE2), and 17b-estradiol (bE2) [5]. * Corresponding author. Tel.: +1 418 654 2617; fax: +1 418 654 2600. E-mail address: tyagi@inrs-ete.uquebec.ca (R.D. Tyagi). 1359-5113/$ – see front matter # 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.procbio.2005.09.017 EDCs are often dominant and can disperse quickly in the environment. EDCs are released to the atmosphere as a result of combustion and incineration activities (polycyclic aromatic hydrocarbons (PAHs), dioxins) [6], but the principal sinks for EDCs are groundwater, river, and lakes [7]. The four main classes of EDCs (natural steroidal estrogens, synthetic estrogens, phytoestrogens, and various industrial chemicals) are generally represented with respect to their estrogenic potency [8]. The natural and synthetic estrogens generally display much stronger estrogenic effects than the phyto- and xenoestrogens. However, the concentrations of phyto- and xenoestrogens in the aquatic environment are usually higher [9]. The list of trace contaminants or EDCs, resulting from human activities and found in wastewater, is long [10–12]. However, in general natural (E1, bE2, estriol [E3]) and synthetic (EE2, mestranol) hormones are the major contributors to the estrogenic activity observed in sewage effluents [13–15] and the receiving water. Recent research showed that several sewage treatment plant (STP) effluents and rivers in the United Kingdom [14,16–20] and in the United States [21,22] contain sufficient amount of estrogenic compounds to induce harmful effects on fish (Tables 1–3). Field studies using caged trout (Oncorhynchus mykiss), wild cyprinid roach (Rutilus rutilus) [40], and estuarine flounder (Platichthys flessus) [41,42] showed that the estrogenicity persists in receiving water and 526 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 Table 1 Estrogens concentrations in STP influent Sampling site Paris, France England Germany Italy Roma, Italy Roma, Italy Barcelona, Spain Japan Influent concentrations (ng/L) Estrone 17b-Estradiol Estriol 17a-Ethinylestradiol 9.6–17.6 1.8–4.1 66 52 31 44 <2.5–115 – 11.1–17.4 <0.3 22.7 12 9.7 11 <5–30.4 5 11.4–15.2 – – 80 57 72 <0.25–70.7 – 4.9–7.1 <LODa – 3 4.8 – <5 – Analysis method Reference SPE/GC-MS SPE/GC-MS-MS SPE/LC-ESI-MS-MS SPE/LC-MS-MS SPE/LC-MS-MS SPE/LC-ESI-MS-MS SPE/LC-MS SPE/ELISA [23] [24] [25] [26] [25] [27] [28] [29] ELISA: enzyme-linked immunosorbent assay; ESI: interface electrospray; GC-MS: gas chromatography-mass spectrometry; GC-MS-MS: gas chromatographytandem mass spectrometry; LC-MS: liquid chromatography-mass spectrometry; LC-MS-MS: liquid chromatography-tandem mass spectrometry; LOD: limit of detection; SPE: solid phase extraction. a 0.3 ng/L. that the concentration of these compounds present in the rivers and the estuaries are high enough to induce deleterious reproductive consequences. The incidence of hermaphroditic wild fish near STPs initiated an investigation on STPs effluent estrogenicity. Caged fish held downstream of some STPs produced vitellogenin (VTG), indicating the presence of estrogenic substances [17,18,43]. In 1990, British scientists showed that male rainbow trout produced the yolk precursor protein VTG when they were exposed to sewage effluents or contaminated surface water [44]. Other studies have also shown that birds, reptiles, and mammals in polluted areas undergo alterations of the endocrine reproductive system [45]. Natural and synthetic estrogen hormones (such as bE2, E3, E1, and EE2) seem to be responsible for endocrine disruption in fish [13,28,46]. Indeed, several studies showed that even low concentrations (ng/L) of bE2 can induce VTG in male species and rainbow trout (O. mykiss) experimentally exposed to these chemicals [46,47]. Purdom et al. [16] and Hansen et al. [48] noticed that concentrations of bE2 as low as 1 ng/L induces VTG Table 2 Estrogens concentrations in STP effluent Samplig site Paris, France Denmark Netherlands Sweden England England England Germany Germany Germany Germany (SW) Italy Roma, Italy Roma, Italy Centre Italy Barcelona, Spain Japan Japan Canada California, USA Effluent concentration (ng/L) Estrone 17b-Estradiol 17a-Estradiol Estriol 17a-Ethinylestradiol Mestranol 6.2–7.2 <2–11 <0.4–47 5.8 1.4–76 <LODb 6.4–29 9 14.6 7 <LODe–18 3 24 17 5–30 <2.5–8.1 – 2.5–34 3 – 4.5–8.6 <1–4.5 <0.6–12 1.1 2.7–48 <LODb 1.6–7.4 <LODd 4.6 6 <LODf–15 1.4 4 1.6 3–8 <5–14.5 <LODd 0.3–2.5 6 0.2–4.1 – – <0.1–5 – – – – – – – – – – – – – – – – – 5.0–7.3 – – – – – 2–4 – – 3 – 20.4 11.7 2.3 n.d.–1 <0.25–21.5 – – – – 2.7–4.5 <1–5.2 <0.2–7.5 4.5 <LODa–4.3 <LODb <LODc 1 – 3 <LODf–12 0.6 1.4 – n.d. <5 – – 9 0.2–2.4 – – – – – – – <LODd – 4 <LODg–2.7 – – – – – – – <LODd – HRGC-MS: high resolution gas chromatography-mass spectrometry; NCI: negative chemical ionization. a 0.2 ng/L. b 0.3 ng/L. c 0.05 ng/L. d 1 ng/L. e 0.7 ng/L. f 0.4 ng/L. g 0.6 ng/L. Analysis method Reference SPE/GC-MS – SPE/GC-MS-MS SPE/GC-MS SPE/GC-MS SPE/GC-MS-MS SPE/GC-NCI-MS SPE/GC-MS-MS SPE/LC-MS-MS SPE/HRGC-MS SPE/GC-MS SPE/LC-ESI-MS-MS SPE/LC-MS-MS SPE/LC-ESI-MS-MS SPE/LC-ESI-MS-MS SPE/LC-MS SPE/ELISA SPE/LC-MS-MS SPE/GC-MS-MS SPE/ELISA [23] [30] [31] [32] [13] [24] [33] [34] [25] [35] [9] [26] [25] [27] [36] [28] [29] [37] [34] [38] 527 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 Table 3 Concentrations of estrogens present in river water Sampling site France Netherlands England Germany Italy Spain (NE) Japan Tokyo, Japan Japan California, USA Concentration (ng/L) Estrone 17b-Estradiol 17a-Estradiol Estriol 17a-Ethinylestradiol Mestranol 1.1–3.0 <0.1–3.4 0.2–10 <LODd 1.5 4.3 – – 0.2–6.6 – 1.4–3.2 <0.3–5.5 <LODa–7.1 <LODd 0.11 6.3 <LODe 32 0.6–1.0 0.05–0.8 – <0.1–3 – – – – – – – – 1.0–2.5 – <LODb–3.1 – 0.33 8 – 5.5 – – 1.1–2.9 <0.1–4.3 <LODc <LODd 0.04 – – – – <0.05–0.07 – – – <LODd – – – – – – Analysis method Reference SPE/GC-MS SPE/GC-MS-MS SPE/GC-NCI-MS SPE/GC-MS-MS SPE/LC-ESI-MS-MS SPE/LC-MS SPE/ELISA SPE/TR-FIA SPE/LC-MS-MS SPE/ELISA [23] [31] [33] [34] [26] [28] [29] [39] [37] [38] TR-FIA: time-resolved fluoroimmunoassay. a 0.03 ng/L. b 0.06 ng/L. c 0.05 ng/L. d 0.5 ng/L. e 1 ng/L. in male trout. In addition, Routledge et al. [46] and Larsson et al. [32] noted that EE2 can be a potential danger to fish and other aquatic organisms, even present at concentrations of 0.1–10 ng/ L. In the study carried out by Purdom et al. [16], EE2 could induce VTG in male fish for a concentration as low as 0.1 ng/L. The alkylphenol polyethoxylates (APEOs) group and their breakdown products, alkylphenols (APs) and alkylphenol carboxylates (APECs), have been shown to be estrogenic as well [46,49]. However, NP and OP are known to be more toxic than their EO precursors [50]. Its frequent use and its stability have as a consequence increased rivers contamination and bioaccumulation risk in the trophic chain [51]. Moreover, NP is present in large amount in STPs sludge and would have as a consequence a diminution of fish reproduction in subsequent receiving water [52]. Several studies proved that NP causes production of vitellogenin in male fish [8,28,53]. Indeed, alkylphenols can have estrogenic effects in fish at concentrations from 1 to 10 mg/L [46,54]. Although nonylphenol polyethoxylates (NPnEO) have been removed from household detergents since 1986, river water quality measurements indicate that there is still NP, nonylphenol ethoxylate (NP1EO) and nonylphenol diethoxylate (NP2EO) concentrations that are as high as 0.571, 0.710, and 0.106 mg/L, respectively [55]. Moreover, Ahel et al. [56] found in Swiss rivers, concentrations in NP2EO above 2.550 mg/L. Several studies have also confirmed the presence of NPnEOs and octylphenol polyethoxylates (OPnEO) in raw sewage, final effluents, sediments, fish, mussels, and even in surface and drinking water, at concentrations ranging from ng/L to mg/L (Tables 4 and 5). Although these values were below acute and chronic toxicity levels, some studies have shown that they individually could be sufficient to produce estrogenic effects [8,46]. Some studies confirmed also the presence of alkylphenol polyethoxylates (APnEOs) in Canadian surface water, sediments, sludge, and sewage treatment plants [60,63,66–68] and St. Lawrence River downstream of the Montreal region. Sabik et al. [69] evaluated the types and levels of APnEO and their metabolites in the municipal effluent of Montreal treatment plant, in surface water, and sediments downstream from the STP. They further studied whether APnEOs were bioconcentrated by mussels (Elliptio complanata) caged and introduced into the St Lawrence River downstream of the major urban zone of Montreal. The analyses were performed on 4-tertoctylphenol (4-t-OP), 4-n-nonylphenol (4-n-NP), nonylphenol polyethoxylates (NP1–16EO), nonylphenoxyacetic acid and Table 4 Concentrations of alkylpehnols and their ethoxylates in STP effluent Sampling site Germany Spain Japan Switzerland Canada USA Concentration (mg/L) NP NP1EO NP2EO OP 0.025–0.77 6–289 0.08–1.24 8 0.8–15 LODa–37 – – 0.21–2.96 49 – – – – – 44 – – 0.002–0.673 – 0.02–0.48 – 0.17–1.7 LODb–0.673 HPLC: high-performance liquid chromatography. a 11 ng/L. b 2 ng/L. Analysis method Reference SPE/HRGC-MS SPE/LC-MS SPE/GC-MS LLE/HPLC GC-MS SPE/HPLC [35] [57] [58] [59] [60] [61] 528 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 Table 5 Concentrations of alkylpehnols and their ethoxylates in river water Sampling site Concentration (mg/L) Analysis method Reference NP NP1EO NP2EO OP Germany Spain Japan Taiwan 0.0067–0.134 LODa–644 0.05–1.08 1.8–10 – – 0.04–0.81 – – – – – 0.0008–0.054 – 0.01–0.18 – SPE/HRGC-MS SPE/LC-MS SPE/GC-MS SPE/GC-MS [35] [57] [58] [62] Canada <0.01–0.92 – – <0.02–7.8 – <0.02–10 <0.005–0.084 – GC-MS SPE/HPLC [63] USA LODb–1.19 12–95 0.077–0.416 – – 0.056–0.326 – – 0.038–0.398 LODc–0.081 – 0.00156–0.007 SPE/HPLC SPE/GCMS SPE/LC-MS [61] [64] [65] a b c 0.15 mg/L. 11 ng/L. 2 ng/L. nonylphenoxyethoxyacetic acid (NP1EC and NP2EC), and octylphenol-mono and di-ethoxycarboxylic acids (OP1EC and OP2EC). The results showed that many of the target chemicals were present in all the studied matrices (in water from ng/L to mg/L reaching ppm levels in sediments and mussels). 2. Endocrine disrupting compounds removal from wastewater The EDCs presence in the environment is likely to disturb the ecosystems and to affect human health. Thus, the need for developing reliable detection methods, analysis tools, and adapted wastewater treatment processes is now the subject of a quasi-consensus between the scientific communities. 2.1. Conventional treatment processes Municipal and industrial wastewaters contain a multitude of persistent organic compounds derived from domestic and industrial applications. These compounds pass through wastewater treatment systems without being totally intercepted (Table 6) and are continuously discharged into the environment and mainly into surface water and/or groundwater. Although APEOs are highly treatable in conventional biological treatment facilities, effluent from wastewater removed plants is still one of the major sources of APs and APEOs due to incomplete removal and degradation of these surfactants. The concentrations of these APEO metabolites varied among different treatment plants depending on the plant design and efficiency [6]. Many communities in worldwide, such as Europe, use surface or groundwater resources for drinking water production, which contain a significant portion of this wastewater effluent [4]. Svenson et al. [76] detected low but significant levels of estrogenicity in the Swedish rivers estuary, downstream of the STPs. Several studies showed that male fish feminization is linked to the estrogenic compounds occurrence in the STP effluents [9,31,32,34,40]. Current wastewater treatment plants were normally, and in the best cases, designed for carbon, nitrogen, and phosphorus (CNP) removal but a partial EDCs removal is often achieved simultaneously. However, a very few data on the EDCs, and in particular on estrogens, fate in STPs processes are available in the literature [71,77,78]. Indeed, although transformation or degradation processes may eliminate some EDCs from wastewater at variable levels, a large ambiguity persists on the occurred EDCs removal processes mechanism (Table 7). For example, removal pathways for organic pollutants during secondary biological treatment include adsorption onto microbial flocs and removal through the waste sludge, biological or chemical degradation, transformation, and volatilization during aeration [83]. However, Mastrup et al. [82] estimated that less than 10% of natural and synthetic estrogens are removed via biodegradation process, and although a considerable amount is adsorbed to the sludge, the majority of the compounds remain soluble in the effluent. Whereas Johnson et al. [25] could not determine whether biodegradation or sorption is the most important removal mechanisms of these compounds. Thus, it is necessary to look further on the removal mechanisms to improve the existing treatment systems effectiveness and to develop new treatment strategies to remove EDCs from wastewater and sludge. 2.1.1. Physical treatments The nonpolar and hydrophobic nature of many EDCs makes them sorb onto particulates. This suggests that the general effect of wastewater treatment processes would be to concentrate organic pollutants, including EDCs, in the sewage sludge. Mechanical separation techniques, such as sedimentation, would result in significant removal from the aqueous phase to primary and secondary sludges [83]. In conventional treatment system, most of compounds remain in aqueous phase in the effluent, whereas a considerable amount is adsorbed onto sludge during the treatment [81,82]. Concerning estrogens, the log Kow values of estrogens (Table 8) indicate that these compounds should appreciably adsorb onto sediment and sludge [88]. This assumption is emphasized by the detection of high concentrations of estrogens in water released by dewatering sewage sludge [89] and in digested sewage sludge (49 ng/g of bE2 and 37 ng/g 529 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 Table 6 EDCs removal during various STPs treatment processes Compound Concentration Removal efficacy (%) Treatment process Matrice type Reference >80 86 59 96 100 1 2 2 2 2 Municipal waste landfill Municipal STP Domestic STP Domestic STP Municipal STP [29] [27] [25] [70] [71] Influent Effluent 17b-Estradiol 5 ng/L 11 ng/L 9.69 ng/L 28.1 ng/L – <1 ng/L 1.6 ng/L 4 ng/L 1.2 ng/L – Estrone 44 ng/L 31 ng/L 43.1 ng/L – 17 ng/L 24 ng/L 12.3 ng/L – 61 23 69 83 2 2 2 2 Municipal STP Domestic STP Domestic STP Municipal STP [27] [25] [70] [71] Estriol 72 ng/L 57.29 ng/L 381.5 ng/L 2.3 ng/L 11.71 ng/L 5.6 ng/L 97 80 99 2 2 2 Municipal STP Domestic STP Domestic STP [27] [25] [70] 17a-Ethinylestradiol 4.84 ng/L – 1.40 ng/L – 71 78 2 2 Domestic STP Municipal STP [25] [71] Phenol Nitrophenol 2,4-Dichlorophenol NP1EO NP2EO 6 mg/L 11 mg/L 83 mg/L 140.03 mg/L 140.03 mg/L No detected No detected 16 mg/L 1.99 mg/L 1.99 mg/L – – 81 99 99 3 3 3 4 4 Municipal + tannery industry STP Municipal + tannery industry STP Municipal + tannery industry STP Industrial + domestic STP Industrial + domestic STP [72] [72] [72] [73] [73] NP 2.8 mg/L 1.5 mg/L 57.64 mg/L 10 mg/L 73 mg/L <0.05 mg/L 6.6 mg/L 0.65 mg/L 1 mg/L 47.5 mg/L >98 – 99 90 35 1 3 4 2 5 Municipal waste landfill Municipal + tannery industry STP Industrial + domestic STP Domestic STP Industrial STP [29] [72] [73] [70] [74] 4-NP 4-t-OP PCBs 2.37 mg/L 0.88 mg/L 46 ng/L 0.95 mg/L 0.32 mg/L 1.2 ng/L 60 64 97 6 6 1 Municipal STP Municipal STP Municipal waste landfill [75] [75] [29] BPA 0.13 mg/L 7.1 mg/L 2.5 mg/L 1.776 mg/L 0.55 mg/L <0.005 mg/L No detected No detected 0.210 mg/L 0.14 mg/L >96 – – 88 75 1 3 3 6 2 Municipal waste landfill Municipal + tannery industry STP Municipal STP Municipal STP Domestic STP [29] [72] [72] [75] [70] PCDD PCDF 21 pg/L 8.7 pg/L 5.2 pg/L 3.3 pg/L 75 62 1 1 Municipal waste landfill Municipal waste landfill [29] [29] (1) Biodegradation/sedimentation + additional treatment with charcoal; (2) activated sludge; (3) physicochemical treatment + biological processes; (4) pretreatment + primary clarifier + aeration tanks + secondary clarifier; (5) pretreatment + primary settling + biofilters; (6) primary clarifier + activated sludge + biological nitrogen removal + biological phosphorus removal + settle tank. of E1) [88]. In the same way, log Kow were between 4.00 and 6.19 for the APE metabolites (Table 8), suggesting that these substances are hydrophobic substances and may become associated with organic matter [6]. Other researches also studied the estrogens interactions with natural particles at expected environmental levels, or those of the activated sludge treatment. Most results proved that the adsorbed contaminants amount depends on particulate size and roughness, hence depends on the available particle surface as well as material characteristics. Whereas Schäfer and Waite [12] results showed that the adsorbed amount of a chemical is a function of particle mass. This was reflected in the results with activated sludge, where large particles of about 100 mm showed the lowest adsorption. When the particle surface area was considered, the estrogens adsorption on activated sludge was the highest of all the compounds studied [12]. However, if contaminants are adsorbed on activated sludge particles, they accumulate in the wastewater treatment plants sludge. In this case, the application of digested sludge, as fertilizer, on agricultural fields may cause a potential contamination of soil and ground water [34]. If contaminants are dissolved or associated with dissolved natural organics or even stable and unstable colloids, then they get transported easily through wastewater treatment plant [90]. Domestic sewage generally contains fats, mineral oils, greases, and surfactants [83] and so, in addition to sorption onto suspended solids as a removal mechanism, it is possible that compounds may partition onto the nonpolar fat and lipid material in raw sewage. Trace organic compounds, such as natural hormones [91], a wide range of pesticides [92], alkyl phthalates [93], and personal care and pharmaceutically active products (PPCPs) 530 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 Table 7 Ambiguity on the occurred estrogens removal processes mechanism Coumpond Sorption (%) Biodegradation Soluble Reference (%) Estriol – – 80–95 95 – – [79] [26] Estrone – a – [80] 17b-Estradiol a – 28 – 90 – – – – [80] [34] [81] 68 – – 20 – – [81] [78] 17a-Ethinylestradiol 87 b 17b-Estradiol equivalent 3b Estrogen Great amount 10 – [75] Majority [82] a Johnson and Sumpter [80] do not give a precise percentage but support that 17b-estradiol is adsorbed whereas estrone is biodegradable. b On the basis of estrogenic activity. [90] can be removed using nanofiltration (NF) or reverse osmosis (RO) and subsequently accumulate in the concentrate [90]. Schäfer et al. [91] observed that some NF membranes remove E1 by size exclusion and others by adsorption. These adsorptive effects may be driven by hydrogen bonding between E1 and the membrane [91]. Schäfer et al. [90] showed also that the presence of natural or chemical particulates, which adsorbs such contaminants, could significantly increase the potential of MF, UF, and NF to remove trace contaminants. Although MF and UF were not expected to remove such small and polar compounds, Schäfer and Waite [12] observed that trace contaminants removal using submerged MF (Memcor) and UF (Zenon) membranes was as high as during powdered activated carbon (PAC) treatment. This removal was high at low and neutral pH, while it decreased substantially at a pH higher than 10.5. Schäfer and Waite [12] attributed this to adsorption effects, comparable to hydrogen bonding and hydrophobic sorption. Indeed, the contaminants adsorption on hydrophobic membranes is expected to be higher than on hydrophilic materials. Chang et al. [94] studies on microfiltration confirmed that significant concentrations of natural hormones, such as E1, could accumulate on hydrophobic hollow fibre membranes as a result of sorption processes. Table 8 Log octanol/water partition coefficients of estrogens and xenoestrogens Compound log Kow Reference 17a-Ethinylestradiol 17b-Estradiol Estrone Estriol BPA Phenol 4-n-NP 4-t-OP NP1EO NP2EO OP1EO OP2EO 3.67–4.15 3,94–4.01 2.45–3.43 2.55–2.81 3.32, 3.43 1.48 4.48, 6.19 4.12, 5.66 4.17 4.21 4.10 4.00 [71,84,85] [84,85] [84,85] [84,85] [86] [86] [86,87] [86,87] [87] [87] [6] [6] However, they also noticed that the membrane retention decreases with increase in the amount of E1 accumulated on the membrane surface. According to Schäfer et al. [90], an appropriate wastewater pre-treatment followed by a hybrid process: MF or UF, combined with, for example, PAC, coagulation or magnetic ion exchange (MIEX), could remove a considerable amount of small-sized contaminants. These contaminants could be pharmaceuticals, EDCs, including hormones, some agrochemicals, viruses, etc. It is important to understand such retention and adsorption effects prior to membrane selection if the membrane is expected to act as a reliable barrier to contaminants. Such adsorption effects are also very important for the understanding of the pollutants fate in treatment systems and possible contaminants desorption during feed quality changes or cleaning operations [12]. Thus, investigations of both fundamental and applied aspects of membrane operation and performance must be carried out to optimize its effectiveness and to contribute to improve treatments strategies. 2.1.2. Chemical treatments Preliminary results indicate that activated carbon is effective for removing some EDCs and PPCPs. In addition, coagulants, such as aluminium and ferric salts, have been used to remove organic matter, although their use is often deemed impractical due to the high costs [95]. However, studies have done a comparative investigation of common adsorbents used in the water and wastewater treatment industry, including PAC, ferric chloride coagulant (FeCl3), and MIEX, that generally allow to remove small-sized contaminants (such as EDCs, including hormones and some agrochemicals) [12]. Results showed that both FeCl3 and MIEX1 are not very suitable to remove the majority of trace contaminants (EDCs and PPCPs). In contrast, Schäfer and Waite [12] showed that PAC is more adequate and appears to be the preferential choice for E1 removal, when PAC is added in a sufficiently high dosage. The EDCs and PPCPs removal is minimal during coagulation since the previous process tends to favour the removal of large and hydrophobic compounds. Indeed, the latter are generally responsible for subsequent adsorption and decantation processes of small-sized contaminants, such as EDCs [12]. 2.1.3. Biological treatments Biological degradation and transformation occur aerobically by biological oxidation in activated sludge, trickling filters, or anaerobically in the sewage system or anaerobic sludge digesters. However, in a study on the distribution of natural estrogens (E1 and bE2) in 18 municipal treatment plants across Canada, Servos et al. [96] noticed that the trickling filter could not reach any removal of bE2. Moreover, Svenson et al. [76] reported that trickling filters were less effective than activated sludge systems (Table 9) to eliminate estrogenic activity, and the highest removal rates were obtained at plants with comprehensive treatment technologies, i.e. combined biological and chemical removal of organic matter, nitrogen, and phosphorus. 531 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 Table 9 Estrogenic activity evaluation during various municipal STPs treatment processes Treatment process Concentration (ng estradiol equivalent/L) Biological treatment Precipitation Influent Effluent – – – – AS AS AS AS AS AS AS biosorption AS + nitrogen removal AS + nitrogen removal Trickling filter/AS Trickling filter Trickling filter Trickling filter Biorotor Direct, Al Direct, Al Direct, Fe(III) Direct, lime Pre, Fe(III) Simultaneous, Al Simultaneous, Al Simultaneous, Fe(III) Simultaneous + post, Fe(III) Pre + post, Al Post, Al Post, Al Pre, Fe(II) Pre + post, Al Pre, Al Post, Al Post, Al Post, Fe(III) 11.9 10.8 5.45 4.15 29.8 5 10.2 5 12.5 8 6.05 3.85 19.5 6.95 6.75 22.35 3.05 1.6 12.4 12.7 5.9 1.1 12.3 0.3 4.3 1.6 1.45 2.55 1.2 <0.1 <0.1 <0.1 1.7 14.85 10.75 5.25 Removal efficacy (%) – – – 73.5 58.7 94 57.8 68 88.4 68.1 80.2 >97.4 >99.5 >98.5 74.8 33.6 – – AS: activated sludge. The activated sludge process is commonly used to treat wastewater in large cities and mainly to remove organic compounds present in STP influent [80]. However, not all compounds are completely broken down or converted to biomass. Indeed, estrogenic alkylphenols and steroid estrogens, for example, found in STP effluent are the breakdown products of incomplete biodegradation of their respective parent compounds [80]. Batch studies realized by Johnson and Sumpter [80] have indicated that E1, EE2, and alkylphenols will not be completely eliminated in activated sludge, in the current configurations of the process. Field data suggested that the activated sludge process can remove over 85% of bE2, E3, and EE2, while the removal performance for E1 appears to be less and more variable [80]. Indeed, in a review on steroid estrogens removal effectiveness, Johnson et al. [25] reported that the activated sludge process could remove 88% of bE2 and 74% of E1. Moreover, Baronti et al. [26] listed six STPs using activated sludge system close to Rome. They reported average removals of 87% of bE2, 61% of E1, 85% of EE2, and 95% of E3 [26]. Ternes et al. [34] studied a number of natural and synthetic estrogens in sewage at a municipal STP near Frankfurt/Main and found that about 2/3 of the incoming bE2 and 16a-hydroxyestrone was eliminated in the STP whereas the elimination efficiencies for E1 and EE2 were low (<10%). In subsequent laboratory experiments with activated sludge from the same plants, Ternes et al. [71] confirmed the persistence of EE2 under aerobic conditions while both E1 and bE2 were degraded fairly rapidly under these conditions (bE2 via E1). In the same way, Esperanza et al. [7] reported that removal efficiencies for E1 and EE2 were around 60% and 65%, respectively, in two pilot-scale municipal wastewater treatment plants, although more than 94% of bE2 entering in the aeration tank was eliminated. Whereas high removals of E3, bE2 [26,70], and EE2 [26] were achieved, no more than 69% of E1 were removed by activated sludge treatment [26,70], and in 4 out of 30 events, E1 outlet levels were even larger than inlet levels [26]. Onda et al. [70] and Johnson and Sumpter [80] concluded that it is necessary to consider bE2 conversion to E1, in E1 effluent concentration. Indeed, batch results obtained by Onda et al. [70] and Esperanza et al. [7] indicated that bE2 was transformed to E1, such as intermediate product. Lee and Liu [97] examined the fate of bE2 in aerobic and anaerobic reactors with activated sludge and observed the rapid degradation of bE2 to E1 but did not observe any other stable major metabolites. Furthermore, Estrogens are either excreted in urine as glucuronide or sulfated conjugates in both humans and animals [98,99]. Indeed, Andreolini et al. [100] observed that E1 is excreted in latepregnancy urine preferentially in conjugated form, estrone-3sulfate (E1-3S). Adler et al. (2001, quoted by Servos et al. [96]) reported that 50% of bE2 and 58% of E1 were conjugated in raw sewage. On this basis, Johnson and Sumpter [80] supposed that the anomalous behaviour of free E1 observed in their study, in those of Shore et al. [101] and Baronti et al. [26], was the result of the microbial deconjugation of E1-3S in the sewer system during the activated sludge STP treatment. Indeed, several studies suggested that the deconjugation could occur during STPs process through microbial processes in the sewage treatment plants [13,41,71,98,99,102], and in rivers [41]. Ternes et al. [71] reported during batch reactor studies that the glucuronides of bE2 (17b-estradiol-(17 or 3)-b-D-glucuronide) were rapidly cleaved in contact with diluted activated sludge resulting in the release of bE2. After less than 15 min, the 17bestradiol-glucuronide was cleaved and both bE2 and E1 could be detected. Carballa et al. [103] investigated the behavior of natural estrogens (E1 and bE2) along the different units of a 532 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 municipal STP located in Galicia (Spain). During the secondary treatment (conventional activated sludge), the increase of E1 concentration in the effluent could be explained by the oxidation of bE2 in the aeration tank and by the cleavage of the conjugates. Furthermore, D’Ascenzo et al. [27] investigated the fate of the conjugated forms of the three most common natural estrogens occurring in the municipal aqueous environment. Levels of conjugated and free E3, bE2, and E1 were studied considering three scenarios: (1) female urine, (2) a septic tank collecting domestic wastewater, and (3) influents and effluents of six activated sludge sewage treatment plants. They confirmed through laboratory biodegradation tests that glucuronated estrogens are readily deconjugated in domestic wastewater, presumably due to the large amount of the bglucuronidase enzyme [104] produced by fecal bacteria (Escherichia coli). Since most of estrogens and androgens are mainly excreted in conjugated form, the occurrence of these free hormones in the aquatic environment (e.g., STP effluents and rivers) is probably due to their deconjugation by bacteria in situ [13,26,31,34,53,79,80,101,105]. According to D’Ascenzo et al. [27] study, the sewage treatment completely removed residues of estrogen glucuronates and with good efficiency (84– 97%) the other analytes, but not E1 (61%) and E1-3S (64%). Therefore, D’Ascenzo et al. [27] concluded following this study, that E1 appears to be the most important natural EDC, considering that (1) the amount of the E1 species discharged from STPs into the receiving water was more than ten times larger than bE2 species, (2) E1 has half the estrogenic potency of bE2, and (3) some E1-3S fraction could be converted to E1 in the aquatic environment. Moreover, the estrogens form greatly influences their estrogenic potency. Matsui et al. [89] compared the estrogenic activity of various substances using the EC50 of the YES response. For instance, the conjugated form 17b-estradiol 3sulfate was 5.3  105 and 17b-estradiol 17-b-D-glucuronide and 17b-estradiol 3-b-D-glucuronide were only 5.9  107 and 3.1  105, respectively, relative to the activity of bE2 [89]. The estrogenic potentials of the conjugated forms of estrogens are clearly much lower. The cleavage of glucuronide during treatment or in the collection system may therefore greatly increase the estrogenicity of the effluent [96]. Like natural hormones (such as bE2), synthetic hormone, EE2, used as estrogenic compound in contraceptives, is metabolized in human body before its excretion. It is found then especially in conjugated forms [99,106]. This conjugation, which inactivate hormonal action of these compounds (e.g., glucuronated and sulphates), increases its water solubility, and thus these compounds become more mobile in environment than free hormones [81]. Indeed, Carr and Griffin [106] noticed that after 24 h, only 3% of EE2 amount (20–50 mg/day) contained in contraceptives remain in plasma, whereas over 60% are excreted in urine. In activated sludge, Turan [107] reported no change in EE2 concentration over 120 h of treatment. However, when Vader et al. [108] added hydrazine as an external electron donor to provide unlimited reducing energy, EE2 degradation increases slightly. This demonstrated that EE2 degradation is mediated by monooxygenase activity. Moreover, Vader et al. [108] found that under non-nitrifying conditions, there was no degradation of EE2, while nitrifying sludge oxidized EE2 to more hydrophobic compounds. Layton et al. [98] also found in laboratory experiments that sludges, which failed to nitrify, also failed to degrade EE2. Vader et al. [108] suggested that the seasonal and temperature effects on nitrification may therefore result in changes in the ability of treatment systems to remove EE2 and related compounds. In addition, Lee and Liu [97] showed in batch experiments that bE2 was more persistent under anaerobic conditions than under aerobic conditions but was still biodegradable by the culture. In addition, Shi et al. [109] investigated the biodegradability of natural and synthetic estrogens using nitrifying activated sludge (NAS) and ammonia-oxidizing bacterium Nitrosomonas europaea. The results confirmed that NAS significantly degrades both natural and synthetic estrogens. Among the four estrogens, bE2 was most easily degraded. NAS degraded 98% of bE2 at 1 mg/L within 2 h, which indicates that NAS also has excellent bE2-degradation ability. Regarding EE2, Shi et al. [109] found a similar trend to Vader et al. [108]. Shi et al. [109] showed also that ammonia-oxidizing bacteria such as N. europaea can contribute to the estrogen degradation by NAS. However, NAS degrades estrogens and their degradation intermediates, while N. europaea only degrades estrogens with no further degradation of their intermediates. Thus, other microorganisms could exist in NAS which are not ammonia-oxidizing bacteria, and are responsible for intermediates degradation. Indeed, E1 was generated when NAS degraded bE2, whereas E1 was not generated when N. europaea degraded bE2. Obviously, bE2 degradation via E1 by NAS is considered to be caused by other heterotrophic bacteria and not by nitrifying bacterium such as N. europaea. On the other hand, since estrogens are hydrophobic organic compounds with low volatility, sorption to sludge could play an important role in removal of these compounds during the waste treatment process. Johnson and Sumpter [80] suggested that the principal mechanisms for steroid estrogens removal in activated sludge processes could be sorption and biodegradation. In general, for more hydrophobic compounds, such as EE2, sorption to sludge is likely to play a significant role in removal of these compounds from solution, while for relatively weakly hydrophobic compounds, such as E3, biodegradation would be a privileged factor [80]. In a recent Danish study on removal processes in activated sludge [30], the results indicated that at common sludge densities in Danish STPs about 35–45% of E1, 55–65% of bE2 and EE2 can be expected to be sorbed to the sludge. The degradation of these compounds was studied under aerobic and anaerobic conditions in a simulated activated sludge system. It is concluded that under anaerobic conditions, the degradation rates for E1 and EE2 were considerably (10–20 times) lower than under aerobic conditions while the degradation of bE2 was not significantly diferent [30]. Moreover, steroids removal can be influenced by hydraulic retention time (HRT) and high sludge retention time (SRT) used by STPs [83]. Indeed, in another Danish literature review on substances causing feminization of fish [77], it was concluded M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 that a high HRT and SRT in the activated sludge treatment process have a positive influence on the ability of an STP to remove estrogen. However, a study of mass balance of estrogens in STP in Germany [75] demonstrated that most of the estrogenic activity in the wastewater was biodegraded during treatment rather than adsorbed onto suspended solids. There was a 90% reduction in estrogenic load, and less than 3% of the estrogenic activity was found in the sludge (Table 7). Moreover, radiolabelled bE2 was used in a study of estrogen fate in STP [110]. Fuerhacker et al. [110] concluded that at low concentration, the majority of radiolabelled bE2 remained in the liquid phase, and thus the physical-chemical properties, such as the octanol/water partition coefficient, did not reflect the situation at neon gram range [110]. In another study (Johnson, 1999, quoted by Birkett and Lester [83]), suspended solids content was an important factor. A higher suspended solids content resulted in a higher removal of estrogens, while an increase in influent estrogen concentration caused a decrease in removal, probably due to the EDCs sorption on the suspended solids. In the case of the surfactants group, the oxidative shortening of the polyethoxylate chain occurs easily and rapidly in aerobic conditions. However, complete mineralization is poor due to the presence of the highly branched alkyl group on the phenolic ring. The hydrophilic group in ethoxylated compounds contains more abundant carbon than the hydrophobic alkyl group. These moieties (available by the successive removal of ethoxy groups) are therefore potential sources of bacterial nutrients. This chain shortening results in the formation of recalcitrant intermediates such as nonylpehnol (NP), octylphenol (OP), and mono to triethoxylate alkylphenols (NP1EO, NP2EO, and NP3EO) [6]. Ultimate biodegradation of these metabolites occurs more slowly, due to the presence of the benzene ring and their limited water solubility [83]. Moreover, since APs are high lipophilic, in particular 4NP, sorb onto the solid phase making them more resistant to biodegradation [7,58,111,112], whereas APECs are more water-soluble and have a very limited tendency to be found in the solid phase. However, Ying et al. [6] reported that aerobic conditions facilitate further biotransformation of APE metabolites than anaerobic conditions. In STP, bisphenol A is easily removed by biodegradation mechanisms (Matsui et al., 1988, quoted by Birkett and Lester [83,113]). Polychlorinated biphenyls (PCBs) are stable molecules with low aqueous solubility and biological, chemical, and physical recalcitrance. As a result, they exhibit minimal degradation in the STP [114], and according to McIntyre and Lester (1981, quoted by Birkett and Lester [83]), the major removal mechanism of PCB, such as organotins [115,116], is via adsorption to suspended matter and sludge flocs. Air stripping has been also noted as an important factor for compounds with HC greater than 100 Pa m3 h1 [83]. For polyaromatic hydrocarbon compounds, degradation times could be as long as 80– 600 h in a conventional STP, since the experiments were run in ideal conditions with a temperature of 20 8C and pre-adapted bacteria. During volatilization, significant removal was seen, and during photodegradation some compounds demonstrated significant losses in settled sewage. In accordance with Melcer et al. 533 [117], Hegeman et al. [118], and Chiou et al. [119], PAH’s removal during primary sedimentation is a function of molecular weight and suspended solids removal efficiency, since they tend to partition onto the solid phase. In conclusion, activated sludge processes allow a relatively high EDCs removal [26,27,70,76,80], however it does not permit to reach lower estrogens, alkylphenols, and BPA effluent concentrations, than the maximum limit levels reported as producing estrogenic effects in fish and other aquatic organisms (Tables 6 and 9). For example, Servos et al. [96] reported that the degradation of estrogens in aerobic batch reactors with a sewage sludge was very rapid, with bE2 and E1 being reduced by >95% in less than 24 h. However, even after 120 h, traces of E1 and estrogenicity could be detected [96]. In addition, it is necessary to notice that the concentration and the removal rates obtained in different studies are not easily comparable, since the treatment conditions at the studied wastewater treatment plants are different or sometimes not clearly described. Moreover, the sampling strategy and the analytical methods vary from a study to another [80]. Membrane bioreactors can be defined as systems integrating biological degradation of waste products with membrane filtration [120]. These treatment systems proved a quite effective removal of organic and inorganic contaminants as well as biological entities from wastewater [121]. Indeed, since estrogens bind readily to organic matter, membrane bioreactor could provide a suitable environment for EDCs removal due to the high organic content in the mixed liquor, and the retention of all particular and colloidal matter before the draw phase. In addition, the possibility of maintaining high SRT in the membrane bioreactor leads to a diverse microbial culture, including slow growing organisms, capable of breaking down complex organic compounds [122]. Thus, compared to other biological treatment, Buenrostro-Zagal et al. (2000, quoted by Cicek [121]) found a better removal effectiveness of 2,4dichlorophenoxyacetic acid using a selective extractive membrane bioreactor. Furthermore, Wintgens et al. [123] investigated membrane bioreactors application and nanofiltration with the aim of evaluating the potential of EDC removal. It was obvious throughout the results, that most of the load was reduced in the membrane bioreactor, while granulated activated carbon treatment, applied downstream, was only a further polishing stage. Inded, some membrane bioreactors configurations allow the retention, and consequently break down of many EDCs without requiring sophisticated tertiary treatment processes [121]. 2.2. Advanced treatment processes 2.2.1. Chlorination process Several studies Hu et al. [124] and Moriyama et al. [125] showed that bE2 and EE2, respectively, reacted rapidly with HOCl and are completely removed (Table 10). However, several chlorinated by-products formed. Moreover, it has been reported that some of chlorinated products have carcinogenicity and/or mutagenicity [125]. Thus, it is important to identify the products from the reaction of EDCs with available chlorine and their estrogenic activities associated [124,125]. Indeed, Hu 534 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 Table 10 Removal of estrogens by advanced treatment processes Compound Concentration Removal (%) Reaction time Added dose Reference 0.015 mg/Lc 9.7–28 ng/Ld 3.0–21 ng/Ld >80 95 18 min 10 min 5 mg O3/L 5 mg O3/L [126] [127] Chlorination 17b-Estradiol 17b-Estradiol 17a-Ethinylestradiol 50 mg/Le 107 M e 0.2 mmol/Le 100 100a 100 10 min 36 h 5 min 1.46 mg/L of sodium hypochlorite 1.5 mg/L of chlorine 1 mmol/L of chlorine [124] [128] [125] MnO2 17a-Ethinylestradiol 15 mg/Le 81.7 1.12 h – [129] TiO2 17b-Estradiol 0.05–3 mmol/Le 98 3.5 h – [130] TiO2 + UV 17b-Estradiol 106 M e 30 min 3h 1.0 g/L of TiO2 in suspension [131] Ozonation Estrone Estrone, 17b-estradiol a b c d e 99 100b Complete removal of estrogenic activity. Decomposed completely into CO2. Municipal STP effluent. Wastewater from secondary treatment. Synthetic water. et al. [124] could determine mainly the formation of 4-chloroE2, 2,4-dichloro-E2, and 2,4-dichloro-E1, and others compounds non-identified. Hu et al. [124] concluded that the products in aqueous chlorinated bE2 solution elicited estrogenic activity. Moreover, Moriyama et al. [125] confirmed the formation of two products in highly chlorinated solutions after 60 min (4-chloro-EE2, 1–6 mol%; 2,4-dichloro-EE2, 3– 25 mol%). The estrogenic activities of 4-chloro-EE2 were similar to those of the parent EE2. 2.2.2. Ozonation and advanced oxidation processes Ternes et al. [126], Nakagawa et al. [127], and Kosaka et al. [132] could remove considerably various estrogens during ozonation treatment (Table 10). Huber et al. [133] determined, in bench-scale experiments, the rate constants of EE2 for ozonation (kO3 ¼ 7  109 M1 s1 ) and AOP (kOH = 9.8  109 M1 s1). However, EDCs co-exist with other organic and inorganic compounds, whose concentrations are relatively high in environmental water. The reaction of HO is less selective, and thus the generated HO is ineffectively consumed by the coexisting compounds. It is assumed that EDCs removal efficiencies are dependent on the initial concentrations of EDCs, co-existing compounds and their reactivities toward ozone and HO. Furthermore, the ozonation products formed are currently unknown [126]. However, hydroxylated estrogens should lose their affinity for the estrogen receptor to greatly reduce the known estrogenic activities of wastewater, but this assumption has not been proved [126]. Moreover, Huber et al. [133] concluded that modifications caused by ozonation or AOPs should be sufficient to eliminate the estrogenic effects of EE2. However, the reactions with ozone and OH radicals during an ozonation process will not result in the complete mineralization of EE2. 2.2.3. Treatment with manganese oxide Rudder et al. [129] obtained an EE2 removal of 81.7% using manganese oxide (MnO2) (Table 10). Moreover, since the MnO2 reactor was not yet saturated after 40 days of treatment, they concluded that EE2 was not only adsorbed to the MnO2 granules, but most probably also degraded into others compounds. Thus, the self-regenerating cycle of MnO2 seems possible. This can make this treatment cost-effective, because the matrix does not have to be replaced [129]. However, Rudder et al. [129] did not identify the EE2 metabolites and neither their estrogenic activity. 2.2.4. Photolysis reactions Photolysis reactions have been extensively studied for estrogens removal from aqueous environment [130,131,134,135] (Table 10). Liu and Liu [135] examined the UV-light and UV–vis-light (high-pressure mercury lamp) direct photolysis of two estrogens, bE2 and E1, in aqueous solution at high concentrations. They could show that the photolysis of both the estrogens causes the breakage and oxidation of benzene rings to produce compounds containing carbonyl groups. Moreover, Ohko et al. [131] concluded in his study on the bE2 degradation by titanium dioxide (TiO2) photocatalysis, that the phenol moiety of the bE2 molecule should be the starting point of the photocatalytic oxidation. In addition, since the intermediate products do not have a phenol ring, Ohko et al. [131] presumed that their estrogenic activities are negligible. 3. Discussion and conclusion It has generally been observed that primary treatment alone results in no or only limited removal of estrogens from sewage, while secondary treatment involving activated sludge reduces M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 significantly all estrogens concentrations. Moreover, a long SRT appears to have a positive influence on the activated sludge system ability to eliminate estrogens. It appears also that bE2 and E3 are very efficiently removed in the latter systems while the removal rate of E1 and EE2 is somewhat lower. It seems also that EE2 only undergoes significant removal degradation when nitrification steps are present. On the other hand, according to the literature, the main estrogens removal mechanism in the activated sludge system seems to be sorption to sludge particles and/or microbiological degradation. However, due essentially to issues met during estrogens sludge analysis; there is no publication to date which could prove it. Indeed, almost all the studies only analyzed the estrogens in the STP influent and effluent, and so assumed that the difference was adsorbed in STP sludge. Moreover, the highest EDCs removal achieved with the different above exposed treatment processes does not allow generally to reach effluent concentrations, which respect the maximum limit levels determined as producing estrogenic effects in fish and other aquatic organisms (Tables 6 and 9). So, it would be interesting to look further into investigation on treatment processes to achieve concentrations in effluent below estrogenic limits. Indeed, Donova et al. [136] reported that a wide variety of microorganisms of different taxonomy could have the ability of steroids biotransformation. The use of these microorganisms in STP would be interesting to evaluate. The decomposition processes (such as ozonation and chlorination processes) display a high potential for removing recalcitrant compounds (e.g., estrogens). However, little data exists, and most of researchers used synthetic water with estrogens concentrations over environmental relevant concentrations (Table 10). Therefore, it should be investigated whether these techniques are also feasible for estrogens removal at ng/L levels and from water containing others particles, such as wastewater. Moreover, the majority of advanced treatments produce by-products whose estrogenicity is unknown or in some cases higher or similar to their precursors [124,125]. Treatment of wastewater and sludge contaminated with phenols and other aromatic compounds (e.g., BPA, bE2, and EE2) with enzymes such as peroxidases [137–140] or polyphenol oxidases [139,141,142] is a new and interesting strategy. Since current researches develop the production of enzymes by using municipal and industrial wastewater and sludge, as basic substrate, the overall costs of enzyme production would be reduced [143]. Therefore, the enzymatic treatment process would be a cost-effective alternative for removing EDCs from municipal and/or industrial wastewater. Finally, regarding to treatment processes advantages and disadvantages with respect to the EDCs removal, we could observe through this review that:  Coagulation processes using iron or aluminium salts does not allow any EDCs removal and it is an expensive treatment process.  PAC coagulation could remove a considerable amount of small-sized contaminants such as EDCs including hormones. 535  Filtration processes (UF, MF, NF), used as hybrid process or not, can allow relatively high EDCs removal, however they are also expensive, require a significant maintenance to avoid membranes clogging.  Membrane bioreactor combines the adsorption and biodegradation processes, and thus would be a good compromise for simultaneous CNP and EDCs removal.  Advanced processes allow a high removal of recalcitrant compounds, however many by-products are released and could have an estrogenic activity higher than their precursors. In conclusion, EDCs are of a general concern and are significant research subject. The epidemiological data gives evidence of a possible relationship between chemical exposure and harmful observed effects of endocrine disruption in the living beings. Recent studies on the conventional wastewater treatment processes effectiveness show that the STPs are a significant EDCs point source, particularly for surface water and under ground water. Therefore, future research priorities should include wastewater treatment plant optimization to increase EDCs removal. Acknowledgments The authors are sincerely thankful the ‘‘Natural Sciences and Engineering Council of Canada (Grant A4984)’’ and ‘‘Fonds Québécois de la Recherche sur la Nature et les Technologies’’ (Que., Canada), and to the ‘‘Generalitat of Catalunya’’ (Spain) for financial assistance. References [1] Colborn T, Vom Saal FS, Soto AM. Developmental effects of endocrinedisrupting chemicals in wildlife and humans. Environ Health Perspect 1993;101:378–84. [2] Health Canada. Human health and exposure to chemicals which disrupt estrogen, androgen and thyroid hormone physiology. Environmental and Occupational Toxicology Division, Environmental Health Directorate, HPB. Tunney’s Pasture, P.L., Canada; 1999. [3] Environment Canada. Endocrine disrupting substances in the environment.; 1999, http://www.ec.gc.ca/eds/fact/eds_e.pdf [July 19, 2003]. [4] Filali-Meknassi Y, Tyagi RD, Surampalli RY, Barata C, Riva MC. Endocrine disrupting compounds in wastewater, sludge treatment processes and receiving waters: overview. Pract Period Hazard Tox Radioact Waste Manag 2004;8:1–18. [5] Commission of the European Communities. The implementation of the Community strategy for endocrine disrupters: A range of substances suspected of interfering with the hormone systems of humans and wildlife. COM [1999], vol. 706.; 2001. p. 45. [6] Ying G-G, Williams B, Kookana R. Environmental fate of alkylphenols and alkylphenol ethoxylates—a review. Environ Int 2002;28:215–26. [7] Esperanza M, Suidan MT, Nishimura F, Wang Z-M, Sorial GA, Zaffiro A, et al. Determination of sex hormones and nonylphenol ethoxylates in the aqueous matrixes of two pilot-scale municipal wastewater treatment plants. Environ Sci Technol 2004;38:3028–35. [8] Servos MR. Review of the aquatic toxicity, estrogenic responses and bioaccumulation of alkylphenols and alkylphenol polyethoxylates. Water Qual Res J Can 1999;34:123–77. [9] Spengler P, Körner W, Metzger JW. Substances with estrogenic activity in effluents of sewage treatment plants in southwestern Germany. 1. Chemical analysis. Environ Toxicol Chem 2001;20:2133–41. 536 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 [10] USEPA, Report on environmental endocrine disruption: an effects assessment and analysis. Report No. EPA/630/R-96/012, Washington, DC, USA, 1997. [11] Council NR. Hormonally active agents in the environment. Committee on hormonally active agents in the environment. Washington, DC: National Research Council, 1999. [12] Schäfer AI, Waite TD. Removal of endocrine disrupters in advanced treatment—the Australian approach. In: Proceedings of the IWA World Water Congress, Workshop Endocrine Disruptors, IWA Specialist Group on assessment and control of hazardous substances in water (ACHSW); 2002. p. 37–51. [13] Desbrow C, Routledge EJ, Brighty GC, Sumpter JP, Waldock M. Identification of estrogenic chemicals in STW effluent. 1. Chemical fractionation and in vitro biological screening. Environ Sci Technol 1998;32:1549–58. [14] Rodgers-Gray TP, Jobling S, Morris S, Kelly C, Kirby S, Janbakhsh A, et al. Long-term temporal changes in the estrogenic composition of treated sewage effluent and its biological effects on fish. Environ Sci Technol 2000;34:1521–8. [15] Aerni H-R, Kobler B, Rutishauser BV, Wettstein FE, Fischer R, Giger W, et al. Combined biological and chemical assessment of estrogenic activities in wastewater treatment plant effluents. Anal Bioanal Chem 2004;378:688–96. [16] Purdom CE, Hardiman PA, Bye VJ, Eno NC, Tyler CR, Sumpter J. Estrogenic effects of effluents from sewage treatment works. Chem Ecol 1994;8:275–85. [17] Harries JE, Sheahan DA, Jobling S, Matthiessen P, Neall P, Routledge EJ, et al. A survey of estrogenic activity in United Kingdom inland waters. Environ Toxicol Chem 1996;15:1993–2002. [18] Harries JE, Sheahan DA, Jobling S, Matthiessen P, Neall P, Sumpter JP, et al. Estrogenic activity in five United Kingdom rivers detected by vitellogenesis in caged male trout. Environ Toxicol Chem 1997;16:534– 42. [19] Nichols KM, Miles-Richardson SR, Snyder EM, Giesy JP. Effects of exposure to minicipal wastewater in situ on the reproductive physiology of the fathead minnow (Pimephales promelas). Environ Toxicol Chem 1999;18:2001–12. [20] Harries JE, Janbakhsh A, Jobling S, Matthiessen P, Sumpter JP, Tyler CR. Estrogenic potency of effluent from two sewage treatment works in the United Kingdom. Environ Toxicol Chem 1999;18:932–7. [21] Folmar LC, Denslow ND, Rao V, Chow M, Crain DA, Enblom J, et al. Vitellogenin induction and reduced serum testosterone concentrations in feral male carp (Cyprinus carpio) captured near a major metropolitan sewage treatment plant. Environ Health Perspect 1996;104:1096–101. [22] USEPA, Draft detailed review paper on a fish two-generation toxicity test. EPA contract number 68-W-01-023, Washington, DC, USA, 2002. [23] Cargouët M, Perdiz D, Mouatassim-Souali A, Tamisier-Karolak S, Levi Y. Assessment of river contamination by estrogenic compounds in Paris area (France). Sci Total Environ 2004;324:55–66. [24] Fawell JK, Sheahan D, James HA, Hurst M, Scott S. Oestrogens and oestrogenic activity in raw and treated water in severn trent water. Water Res 2001;35:1240–4. [25] Johnson AC, Belfroid A, Di Corcia A. Estimating steroid estrogen inputs into activated sludge treatment works and observations on their removal from the effluent. Sci Total Environ 2000;256:163–73. [26] Baronti C, Curini R, D’Ascenzo G, Di Corcia A, Gentili A, Samperi R. Monitoring natural and synthetic estrogens at activated sludge sewage treatment plants and in receiving river water. Environ Sci Technol 2000;34:5059–66. [27] D’Ascenzo G, Di Corcia A, Gentili A, Mancini R, Mastropasqua R, Nazzari M, et al. Fate of natural estrogen conjugates in municipal sewage transport and treatment facilities. Sci Total Environ 2003; 302:199–209. [28] Petrovic M, Solé M, López de Alda MJ, Barceló D. Endocrine disruptors in sewage treatment plants, receiving river waters, and sediments: integration of chemical analysis and biological effects on feral carp. Environ Toxicol Chem 2002;21:2146–56. [29] Behnish PA, Fujii K, Shiozaki K, Kawakami I, Sakai S. Estrogenic and dioxin-like potency in each step of a controlled landfill leachate treatment plant in Japan. Chemosphere 2001;43:977–84. [30] Danish Environmental Protection Agency (DEPA). Degradation of estrogens in sewage treatment processes. Environmental Project No. 899. Danish Environmental Protection Agency, Danish Ministry of the Environment; 2004. [31] Belfroid AC, Van der Horst A, Vethaak AD, Schafer AJ, Rijs GBJ, Wegener J, et al. Analysis and occurrence of estrogenic hormones and their glucuronides in surface water and wastewater in Netherlands. Sci Total Environ 1999;225:101–8. [32] Larsson DGJ, Adolfsson-Erici M, Parkkonen J, Petterson M, Berg AH, Olsson P-E, et al. Ethinyloestradiol—an undesired fish contraseptive? Aquat Toxicol 1999;42:91–7. [33] Xiao X-Y, McCalley DV, Mcevoy J. Analysis of estrogens in river water and effluents using solid-phase extraction and gas chromatographynegative chemical ionization mass spectrometry of the pentafluorobenzoyl derivates. J Chromatogr A 2001;923:195–204. [34] Ternes TA, Stumpf M, Mueller J, Haberer K, Wilken RD, Servos M. Behavior and occurrence of estrogens in municipal sewage treatment plants. I. Investigations in Germany, Canada and Brazil. Sci Total Environ 1999;225:81–90. [35] Kuch HM, Ballschmiter K. Determination of endocrine-disrupting phenolic compounds and estrogens in surface and drinking water by HRGC(NCI)-MS in the pictogram per liter range. Environ Sci Technol 2001;35:3201–6. [36] Laganà A, Bacaloni A, De Leva I, Faberi A, Fago G, Marinob A. Analytical methodologies for determining the occurrence of endocrine disrupting chemicals in sewage treatment plants and natural waters. Anal Chim Acta 2004;501:79–88. [37] Isobe T, Shiraishi H, Yasuda M, Shinoda A, Suzuki H, Morita M. Determination of estrogens and their conjugates in water using solidphase extraction followed by liquid chromatography-tandem mass spectrometry. J Chromatogr A 2003;984:195–202. [38] Huang C-H, Sedlak DL. Analysis of estrogenic hormones in municipal wastewater effluent and surface water using enzyme-linked immunosorbent assay and gas chromatography/tandem mass spectrometry. Environ Toxicol Chem 2001;20:133–9. [39] Majima K, Fukui T, Yuan J, Wang G, Matsumoto K. Quantitative measurement of 17b-estradiol and estriol in river water by time-resolved fluoroimmunoassay. Anal Sci 2002;18:869–74. [40] Jobling S, Nolan M, Tyler CR, Brighty G, Sumpter JP. Widespread sexual disruption in wild fish. Environ Sci Technol 1998;32:2498–506. [41] Allen Y, Matthiesen P, Scott AP, Haworth S, Feist S, Thain JE. The extent of estrogenic contamination in the UK estuarine and marine environments-further survey of flounder. Sci Total Environ 1999;233: 5–20. [42] Lye CM, Frid CLJ, Gill ME, McCormick D. Abnormalities in the reproductive health of flounder Platichthys flesus exposed to effluent from a sewage treatment works. Mar Pollut Bull 1997;34:34–41. [43] White R, Jobling S, Hoare SA, Sumpter JP, Parker MG. Environmentally persistent alkylphenolic compounds are estrogenic. Endocrinology 1994;135:175–82. [44] Fent K. Ökotoxikologie: Umweltchemie – Toxikologie – Ökologie, vol. 21. Stuttgart, Germany: Georg Thieme Verlag, 1998. [45] Preziosi P. Endocrine disrupters as environmental signallers: an introduction. Pure Appl Chem 1998;70:1617–31. [46] Routledge EJ, Sheahan D, Desbrow C, Brighty GC, Waldock M, Sumpter JP. Identification of estrogenic chemicals in STW effluent. 2. In vivo responses in trout and roach. Environ Sci Technol 1998;32:1559–65. [47] Sumpter JP, Jobling S, Tyler CR. Oestrogenic substances in the aquatic environment and their potential impact on animals, particularly fish. In: Taylor EW, editor. Toxicology of aquatic pollution: physiological, molecular and cellular approaches. Cambridge University Press; 1996. p. 205–24. [48] Hansen P-D, Dizer H, Hock B, Marx A, Sherry J, McMaster M, et al. Vitellogenin—a biomarker for endocrine disruptors. Trends Anal Chem 1998;17:448–51. M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 [49] Jobling S, Reynolds T, White R, Parker MG, Sumpter JP. A variety of environmentally persistent chemicals, including some phthalate plasticizers, are weakly estrogenic. Environ Health Perspect 1995;103: 582–7. [50] Renner R. European bans on surfactant trigger transatlantic debate. Environ Sci Technol 1997;31:316A–20A. [51] Cravedi J-P, Les xéno-estrogènes, perturbateurs endocriniens potentiels. Laboratoire des Xénobiotiques, Centre INRA de Toulouse, 1999. http://www.inra.fr/PRESSE/COMMUNIQUES/nhsa99/dp7.htm [Sept. 15, 2003]. [52] Sea-River Magazine, Le nonylphénol: un ‘‘perturbateur endocrinien’’ pour les poissons. 2001. http://www.sea-river-news.com/06_1.htm [Sept. 15, 2003]. [53] Kirk LA, Tyler CR, Lye CM, Sumpter JP. Changes in estrogenic and androgenic activities at different stages of treatment in wastewater treatment works. Environ Toxicol Chem 2002;21:972–9. [54] Jobling S, Sumpter JP. Detergent components in sewage effluent are weakly oestrogenic to fish: an in vitro study using rainbow trout (Oncorhynchus mykiss) hepatocytes. Aquat Toxicol 1993;27:361–72. [55] Fuerhacker M, Scharf S, Pichler W, Ertl T, Haberl R. Sources and behaviour of bismuth active substances (BiAS) in a municipal sewage treatment plant. Sci Total Environ 2001;277:95–100. [56] Ahel M, Molnar E, Ibric S, Giger W. Estrogenic metabolites of alkylphenol polyethoxylates in secondary sewage effluents and rivers. In: Proceedings of the third IWA specialized conference, hazard assessment and control of environmental contaminants—ECOHAZARD’99; 1999. p. 25–32. [57] Solé M, de Alda LMJ, Castillo M, Porte C, Ladegaard-Pedersen K, Barcelo D. Estrogenicity determination in sewage treatment plants and surface waters from the Catalonian area (NE Spain). Environ Sci Technol 2000;34:5076–83. [58] Isobe T, Nishiyama H, Nakashima A, Takada H. Distribution and behavior of nonylphenol, octylphenol, and nonylphenol monoethoxylate in Tokyo metropolitan area: their association with aquatic particles and sedimentary distributions. Environ Sci Technol 2001;35:1041–9. [59] Ahel M, Giger W. Determination of alkylphenols and alkylphenol monoand diethoxylates in environmental samples by high-performance liquid chromatography. Anal Chem 1985;57:1577–83. [60] Lee HB, Peart TE. Determination of 4-nonylphenol in effluent and sludge from sewage treatment plants. Anal Chem 1995;67:1976–80. [61] Snyder SA, Keith TL, Verbrugge DA, Snyder EM, Gross TS, Kannan K, et al. Analytical methods for detection of selected estrogenic compounds in aqueous mixtures. Environ Sci Technol 1999;33:2814–20. [62] Ding W-H, Tzing S-H, Lo J-H. Occurrence and concentrations of aromatic surfactants and their degradation products in river waters of Taiwan. Chemosphere 1999;38:2597–606. [63] Bennie DT, Sullivan CA, Lee HB, Peart TE, Maguire RJ. Occurrence of alkylphenols and alkylphenol mono- and diethoxylates in natural water of the Laurentian Great Lakes Basin and the upper St. Lawrence River. Sci Total Environ 1997;193:263–75. [64] Dachs J, van Ry DA, Eisenreich SJ. Occurrence of estrogenic nonylphenols in the urban and coastal atmosphere of the lower Hudson river estuary. Environ Sci Technol 1999;33:266–2679. [65] Ferguson PL, Iden CR, Brownawell BJ. Distribution and fate of neutral alkylphenol ethoxylate metabolites in a sewage-impacted urban estuary. Environ Sci Technol 2001;35:2428–35. [66] Metcalfe CD, Hoover L, Sang S. Nonylphenol ethoxylates and their use in Canada, vol. 3. World Wildlife Fund Canada Report, 1996. [67] Bennie DT, Sullivan CA, Lee HB. Alkylphenol polyethoxylate metabolites in Canadian sewage treatment plant waste streams. In: Proceedings of the SETAC 18th Annual Meeting, Society of Environmental Toxicology and Chemistry; 1997. [68] Lee HB, Peart TE, Bennie DT, Maguire RJ. Determination of nonylphenol polyethoxylates and their carboxylic acid metabolites in sewage treatment plant sludge by supercritical carbon dioxide extraction. J Chromatogr A 1997;785:385–94. [69] Sabik H, Gagné F, Blaise C, Marcogliese DJ, Jeannot R. Occurrence of alkylphenol polyethoxylates in the St. Lawrence river and their biocon- [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] 537 centration by mussels (Elliptio complanata). Chemosphere 2003;51: 349–56. Onda K, Nakamura Y, Miya A, Katsu Y. The behavior of estrogenic substances in the biological treatment process of sewage. Water Sci Technol 2003;47:109–16. Ternes TA, Kreckel P, Mueller J. Behaviour and occurrence of estrogens in municipal sewage treatment plants. II. Aerobic batch experiments with activated sludge. Sci Total Environ 1999;225:91–9. Farré M, Klöter G, Petrovic M, Alonso MC, de Alda LMJ, Barceló D. Identification of toxic compounds in wastewater treatment plants during a field experiment. Anal Chim Acta 2002;456:19–30. Planas C, Guadayol JM, Droguet M, Escalas A, Rivera J, Caixach J. Degradation of polyethoxylated nonylphenols in a sewage treatment plant. Quantitative analysis by isotopic dilution-HRGC/MS. Water Res 2002;36:982–8. Sheahan DA, Brighty GC, Daniel M, Kirby SJ, Hurst MR, Kennedy J, et al. Estrogenic activity measured in a sewage treatment works treating industrial inputs containing high concentrations of alkylphenolic compounds—a case study. Environ Toxicol Chem 2002;21:507–14. Körner W, Bolz U, Sübmuth W, Hiller G, Schuller W, Hanf V, et al. Input/ output balance of estrogenic active compounds in a major municipal sewage plant in Germany. Chemosphere 2000;40:1131–42. Svenson A, Allard A-S, Ek M. Removal of estrogenicity in Swedish municipal sewage treatment plants. Water Res 2003;37:4433–43. Danish Environmental Protection Agency (DEPA). Feminisation of fish— the effect of estrogenic compounds and their fate in sewage treatment plants and nature. Environmental Project No. 729. Danish Environmental Protection Agency, Danish Ministry of the Environment; 2002 . Danish Environmental Protection Agency (DEPA). Evaluation of analytical chemical methods for detection of estrogens in the environment. Working Report No. 44. Danish Environmental Protection Agency, Danish Ministry of the Environment; 2003. Lee H, Peart T. Determination of 17b-estradiol and its metabolites in sewage effluent by solid-phase extraction and gas chromatography mass spectrometry. J AOAC Int 1998;81:1209–16. Johnson AC, Sumpter JP. Removal of endocrine-disrupting chemicals in activated sludge treatment works. Environ Sci Technol 2001;35:4697–703. Kozak RG, D’Haese I, Verstraete W. Pharmaceuticals in the environment: focus on 17a-ethinyloestradiol. In: Kümmerer K, editor. Pharmaceuticals in the environment. Sources, fate, effects and risk. Heidelberg, Germany: Springer Verlag; 2001. p. 49–65. Mastrup M, Jensen RL, Schäfer AI, Khan S. Fate modeling—an important tool for water recycling. In: Schäfer AI, Sherman P, Waite TD, editors. Recent advances in water recycling technologies. Australia: Brisbane; 2001. p. 103–12. Birkett JW, Lester JN. Endocrine disrupters in wastewater and sludge treatment processes. London, UK: IWA Publishing, 2003. Lai KM, Johnson KK, Scrimshaw MD, Lester JN. Binding of waterborne steroid estrogens to solid phases in river and estuarine systems. Environ Sci Technol 2000;34:3890–4. Sayles G. Biological fate of estrogenic compounds associated with sewage treatment: a review. Cincinnati, OH: Effective Risk Management of Endocrine Disrupting Chemicals Workshop, 2001. Hu J-Y, Aizawa T. Quantitative structure–activity relationships for estrogen receptor binding affinity of phenolic chemicals. Water Res 2003;37:1213–22. Ahel M, Giger W. Partitioning of alkylphenols and alkylphenol polyethoxylates between water and organic solvents. Chemosphere 1993; 26:1471–8. Ternes TA, Andersen H, Gilberg D, Bonerz M. Determination of estrogens in sludge and sediments by liquid extraction and GC/MS/MS. Anal Chem 2002;74:3498–504. Matsui S, Takigami H, Matsuda T, Taniguchi N, Adachi J, Kawami H, et al. Estrogen and estrogen mimics contamination in water and the role of sewage treatment. Water Sci Technol 2000;42:173–9. Schäfer AI, Mastrup M, Jensen RL. Enhancing particle interactions and removal of trace contaminants from water and wastewaters. Desalination 2002;147:243–50. 538 M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 [91] Schäfer AI, Nghiem DL, Waite TD. Removal of natural hormone estrone from water and wastewater using nanofiltration and reverse osmosis. Environ Sci Technol 2004;38:1888–96. [92] Kiso Y, Nishimura Y, Kitao T, Nishimura K. Rejection properties of nonphenylic pesticides with nanofiltration membranes. J Membr Sci 2000; 171:229–37. [93] Kiso Y, Kon T, Kitao T, Nishimura K. Rejection properties of alkyl phthalates with nanofiltration membranes. J Membr Sci 2001;182:205–14. [94] Chang S, Waite TD, Schäfer AI, Faneb AG. Adsorption of trace steroid estrogens to hollow fibre membranes hydrophobic. Desalination 2002; 146:381–6. [95] European Commission. Pollutants in urban wastewater and sewage sludge. London: European Commission, 2001. [96] Servos MR, Bennie DT, Burnison BK, Jurkovic A, McInnis R, Neheli T, et al. Distribution of estrogens, 17b-estradiol and estrone, in Canadian municipal wastewater treatment plants. Sci Total Environ 2005;336:155– 70. [97] Lee HB, Liu D. Degradation of 17b-estradiol and its metabolites by sewage bacteria. Water Air Soil Pollut 2002;134:353–68. [98] Layton AC, Gregory BW, Seward JR, Schultz TW, Sayler GS. Mineralization of steroidal hormones by biosolids in wastewater treatment systems in Tennessee, USA. Environ Sci Technol 2000;34:3925–31. [99] Jürgens MD, Holthaus KIE, Johnson AC, Smith JJL. The potential for estradiol and ethynilestradiol degradation in English rivers. Environ Toxicol Chem 2002;21:480–8. [100] Andreolini F, Borra C, Caccamo F, Di Corcia A, Samperi R. Estrogen conjugates in late-pregnancy fluids—extraction and group separation by a graphitized carbon-black cartridge and quantification by high-performance liquid-chromatography. Anal Chem 1987;59:1720–5. [101] Shore LS, Gurevitz M, Shemesh M. Estrogen as an environmental pollutant. Bull Environ Contam Toxicol 1993;51:361–6. [102] Tyler CR, Routledge EJ. Oestrogenic effects in fish in English rivers with evidence of their causation. Pure Appl Chem 1998;70:1795–804. [103] Carballa M, Omil F, Lema JM, Llompart M, Garcia-Jares C, Rodriguez I, et al. Behavior of pharmaceuticals, cosmetics and hormones in a sewage treatment plant. Water Res 2004;38:2918–26. [104] Dray J, Dray TF, Ullmann A. Hydrolysis of urinary metabolites of different steroid hormones by glucuronidase from Escherichia coli. Ann Inst Pasteur 1972;123:853–7. [105] Tyler C, Jobling RS, Sumpter JP. Endocrine disrupting in wildlife: a critical review of the evidence. Crit Rev Toxicol 1998;28:319–61. [106] Carr BR, Griffin JE. Fertility controls and its complications. In: Wilson JD, Foester DW, Kronenberg HM, Reed LP, editors. Williams textbook of endocrinology. Philadelphia: W.B. Saunders Company; 1998. p. 901–25. [107] Turan A. Excretion of natural and synthetic estrogens and their metabolites: occurrence and behaviour. Endocrinically active chemicals in the environment. Berlin: German Federal Environment Agency, 1995. [108] Vader JS, van Ginkel CG, Sperling FMGM, de Jong J, de Boer W, de Graaf JS, et al. Degradation of ethinyl estradiol by nitrifying activated sludge. Chemosphere 2000;41:1239–43. [109] Shi J, Fujisawa S, Nakai S, Hosomi M. Biodegradation of natural and synthetic estrogens by nitrifying activated sludge and ammonia-oxidizing bacterium Nitrosomonas europaea. Water Res 2004;38:2323–30. [110] Fuerhacker M, Breithofer A, Jungbauer A. 17b-Estradiol: behaviour during waste water treatment. Chemosphere 1999;39:1903–9. [111] Tanghe T, Devriese G, Verstraete W. Nonylphenol degradation in lab scale activated sludge units is temperature dependent. Water Res 1998; 32:2889–96. [112] John DM, House WA, White GF. Environmental fate of nonylphenol ethoxylates: differential adsorption of homologs to components of river sediment. Environ Toxicol Chem 2000;19:293–300. [113] Staples CA, Dorn PB, Klecka GM, O’Block ST, Harris LR. A review of the environmental fate, effects, and exposures of bisphenol A. Chemosphere 1998;36:2149–73. [114] Morris SL, Lester JN. Behaviour and fate of polychlorinated biphenyls in a pilot wastewater treatment plant. Water Res 1994;28:1553–61. [115] Chau YK, Zhang S, Maguire RJ. Occurrence of butyltin species in sewage and sludge in Canada. Sci Total Environ 1992;121:271–81. [116] Fent K. Organotin compounds in municipal wastewater and sewage sludge: contamination, fate in treatment processes and ecotoxicological consequences. Sci Total Environ 1996;185:151–9. [117] Melcer H, Steel P, Bedford WK. Removal of polycyclic aromatic hydrocarbons and heterocyclic nitrogen compounds in a municipal treatment plant. Water Environ Res 1995;67:926–34. [118] Hegeman MJM, van der Weijden C, Loch FG. Sorption of benzo(a)pyrene and phenanthrene on suspended harbour sediment as a function of sediment concentration and salinity, a laboratory study using solvent partition coefficient. Environ Sci Technol 1995;29:363–71. [119] Chiou CT, McGroddy SE, Kile DE. Partition characteristics of polycyclic aromatic hydrocarbons on soils and sediments. Environ Sci Technol 1998;32:264–9. [120] Cicek N, Winnen H, Suidan MT, Wrenn BE, Urbain V, Manem J. Effectiveness of the membrane bioreactor in the biodegradation of high molecular weight compounds. Water Res 1998;32:1553–63. [121] Cicek N. Membrane bioreactors in the treatment of wastewater generated from agricultural industries and activities. In: Proceedings of the AIC 2002 Meeting CSAE/SCGR Program Saskatoon; 2002. [122] Cicek N, Franco JP, Suidan MT, Urbain V, Manem J. Characterization and comparison of a membrane bioreactor and a conventional activatedsludge system in the treatment of wastewater containing high-molecularweight compounds. Water Environ Res 1999;71:64–70. [123] Wintgens T, Gallenkemper M, Melin T. Endocrine disrupter removal from wastewater using membrane bioreactor and nanofiltration technology. Desalination 2002;146:387–91. [124] Hu J-Y, Cheng S, Aizawa T, Terao Y, Kunikane S. Products of aqueous chlorination of 17b-estradiol and their estrogenic activities. Environ Sci Technol 2003;37:5665–70. [125] Moriyama K, Matsufuji H, Chino M, Takeda M. Identification and behavior of reaction products formed by chlorination of ethynylestradiol. Chemosphere 2004;55:839–47. [126] Ternes TA, Stuber J, Herrmann N, McDowell D, Ried A, Kampmann M, et al. Ozonation: a tool for removal of pharmaceuticals, contrast media and musk fragrances from wastewater? Water Res 2003;37: 1976–82. [127] Nakagawa S, Kenmochi Y, Tutumi K, Tanaka T, Hirasawa I. A study on the degradation of endocrine disruptors and dioxins by ozonation and advanced oxidation processes. J Chem Eng Jpn 2002;35:840–7. [128] Lee B-C, Kamata M, Akatsuka Y, Takeda M, Ohno K, Kamei T, et al. Effects of chlorine on the decrease of estrogenic chemicals. Water Res 2004;38:733–9. [129] Rudder JD, Wiele TVD, Dhooge W, Comhaire F, Verstraete W. Advanced water treatment with manganese oxide for the removal of 17a-ethynylestradiol (EE2). Water Res 2004;38:184–92. [130] Coleman HM, Eggins BR, Byrne JA, Palmer FL, King E. Photocatalytic degradation of 17-[beta]-oestradiol on immobilised TiO2. Appl Catal B Environ 2000;24:L1–5. [131] Ohko Y, Iuchi K-I, Niwa C, Tatsuma T, Nakashima T, Iguchi T, et al. 17Beta-estradiol degradation by TiO2 photocatalysis as a means of reducing estrogenic activity. Environ Sci Technol 2002;36:4175–81. [132] Kosaka K, Yamada H, Matsui S, Shishida K. The effects of the coexisting compounds on the decomposition of micropollutants using the ozone/hydrogen peroxide process. Water Sci Technol 2000;42: 353–61. [133] Huber MM, Canonica S, Park G-Y, von Gunten U. Oxidation of pharmaceuticals during ozonation and advanced oxidation processes. Environ Sci Technol 2003;37:1016–24. [134] Segmuller BE, Armstrong BL, Dunphy R, Oyler AR. Identification of autoxidation and photodegradation products of ethynylestradiol by online HPLC–NMR and HPLC–MS. J Pharm Biomed 2000;23:927–37. [135] Liu B, Liu X. Direct photolysis of estrogens in aqueous solutions. Sci Total Environ 2004;320:269–74. [136] Donova MV, Egorova OV, Nikolayeva VM. Steroid 17b-reduction by microorganisms—a review. Process Biochem 2005;40:2253–62. [137] Wagner M, Nicell JA. Detoxification of phenolic solutions with horseradish peroxidase and hydrogen peroxide. Water Res 2002;36: 4041–52. M. Auriol et al. / Process Biochemistry 41 (2006) 525–539 [138] Sakuyama H, Endo Y, Fujimoto K, Hatano Y. Oxidative degradation of alkylphenols by horseradish peroxidase. J Biosci Bioeng 2003;96:227–31. [139] Suzuki K, Hirai H, Murata H, Nishida T. Removal of estrogenic activities of 17b-estradiol and ethinylestradiol by ligninolytic enzymes from white rot fungi. Water Res 2003;37:1972–5. [140] Filali-Meknassi Y, Auriol M, Adams C, Tyagi RD. Natural and synthetic hormones removal by enzymatic degradation. In: Proceedings of 20th Canadian Association on water quality (CAWQ); 2004. 539 [141] Ikehata K, Nicell JA. Characterization of tyrosinase for the treatment of aqueous phenols. Bioresour Technol 2000;74:191–9. [142] Bevilaqua JV, Cammarota MC, Freire DMG, Sant’Anna Jr GL. Phenol removal through combined biological and enzymatic treatment. Bras J Chem Eng 2002;19:151–8. [143] Ikehata K, Buchanan ID, Smith DW. Recent developments in the production of extracellular fangal peroxidases and laccases for waste treatment. J Environ Eng Sci 2004;3:1–19.
Keep reading this paper — and 50 million others — with a free Academia account
Used by leading Academics
James Bashkin
University of Missouri - St. Louis
Sudip Shyam
University of Waterloo, Canada
Sabina Begic
University of Tuzla
Yucel Kadioglu
Ataturk University