Next Article in Journal
Particle-Based Workflow for Modeling Uncertainty of Reactive Transport in 3D Discrete Fracture Networks
Next Article in Special Issue
Meta-Evaluation of Water Quality Indices. Application into Groundwater Resources
Previous Article in Journal
Permeability Coefficient of Low Permeable Soils as a Single-Variable Function of Soil Parameter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Insight into Ingredients of Toxicological Interest in Personal Care Products and A Small–Scale Sampling Survey of the Greek Market: Delineating a Potential Contamination Source for Water Resources

1
School of Spatial Planning and Development, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
2
School of Science and Technology, Hellenic Open University, Aristotelous 18, 263 35 Patra, Greece
3
Department of Electrical and Electronics Engineering, University of West Attica, 250 Thivon Av., 122 44 Athens, Greece
4
Department of Civil Engineering, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
*
Authors to whom correspondence should be addressed.
Water 2019, 11(12), 2501; https://doi.org/10.3390/w11122501
Submission received: 10 October 2019 / Revised: 20 November 2019 / Accepted: 26 November 2019 / Published: 27 November 2019

Abstract

:
Wastewater is not a waste but a valuable resource that should be reused. Nevertheless, it should be devoid of physical, chemical, and microbiological parameters that can harm the consumer. Along with the multitude of possible pollutants found in wastewater treatment plants (WWTPs), emerging pollutants, such as Personal Care Products (PCPs), have arisen. The present research examines some of the main ingredients commonly found in PCPs, focusing on their toxicological profile on their occurrence in WWTPs influents and effluents worldwide and on their persistence and biodegradability. A small-scale market sampling of PCPs was performed in Athens, Greece, in June 2019, and their individual ingredients were recorded, coded according to their main activity, scanned for the presence of ingredients of important toxicological profile, and finally analyzed for the presence of other candidates of toxicological interest. Results show that some ingredients of concern (i.e., parabens and triclosan) are a decreasing trend. On the other hand, information on the presence of synthetic musks and perfume synthesis is scarce and encumbered by brand protection. Finally, UV filters are numerous, and they are used in various combinations, while other ingredients of toxicological interest are also present. Since the reclaimed water may well be used to cover irrigation needs in Greek areas with water deficiency or to enrich bodies of surface water, it is important to know what PCP ingredients are on the rise in the market, to monitor their presence in WWTPs influents and effluents and to extend research on their environmental fate and behavior.

1. Introduction

Wastewater is not a waste. It is an essential part of the biosphere and a critical parameter of life. In the context of a circular economy and recharging of bodies of surface- or groundwater, wastewater is considered an essential resource after proper treatment. This task is accomplished in wastewater treatment plants (WWTPs) through a series of specialized treatments (primary, secondary, tertiary, etc.) [1]. The necessity of wastewater treatment is substantiated in [2], which is enforced in Greek law through [3] and its updates. Nowadays the majority of the population in Greece (9,643,700 people) has access to sanitation through 260 wastewater treatment plants (data of 2011).
The Directive [2] has been in force for 25 years; during these years, many changes have occurred: Depletion of natural resources, palpable manifestations of climate change, as well as novel environmental threats and pressures [4]. As such, the directive is now in the process of re-evaluation [4]. Throughout these 25 years, emerging pollutants such as personal care products (PCPs) have been added to the bulk of contaminants that should be treated in a WWTP. The short- and long-term effects of these pollutants on humans and other organisms are still mostly unknown. Their maximum concentration below effect level is in most cases still not defined. As such, their continuous emission and their presence in influents of WWTP, even in small quantities, cause concern [5]. Global water shortage and water quality deterioration are expected to increase in the following years, aided by climate change, thus amending policies and management plans of freshwater resources [6]. In this context, the present research focuses on PCPs, which comprise a vast number of market commodities that are used by millions of people for their personal hygiene needs and their beautification [7]. Most PCPs meet the definitions of cosmetics, while some are defined as both cosmetics and drugs (e.g., anti-dandruff shampoos). PCPs also comprise wet or dry hand tissues, sanitary towels, and baby-wipes [7]. Most PCPs are consumed through the dermal route, thus some part is washed off, they are used and misused extensively, and they may be disposed of improperly, contributing to pollution of water bodies [8]. Some studies have incriminated certain PCPs ingredients for negative effects on the health of humans and of other living organisms. Furthermore, due to their lipophilic nature, some of these ingredients adsorb particles and sediment and bioaccumulate along the trophic chain [5]. Much less is known for their fate and behavior in the environment and the degree of their catabolism in WWTPs. The aim of the present research is to show a comprehensive human hazard assessment of the most important ingredients of toxicological interest found in PCPs, focusing on their occurrence in WWTPs, since the recovered wastewater is commonly used to recharge water bodies. The identification of such substances in wastewater is very critical since typical WWTP processes usually do not remove them, and thus other solutions should be foreseen if they are present in order to reclaim and circulate the recovered water safely. An investigation of fundamentally distinct PCPs that are found in the Greek market (2019) was also performed in order to reveal present trends of these ingredients. Some useful conclusions are drawn regarding these polluting sources of WWTP influents, which may render effluents problematic for reuse.

2. Materials and Methods

2.1. Ingredients of Toxicological Interest Commonly Found in PCPs

2.1.1. Parabens

Parabens (PBs) are esters of parahydroxybenzoic acid (Figure 1) and are widely used as preservatives due to their antimicrobial and antifungal activities. Despite their long and extensive use, concerns have arisen due to indications of endocrine disruption [9] in a number of in vitro and in vivo studies. More analytically, PBs tested positive in vitro for estrogenic activity [10], they induced the development of human MCF-7 breast cancer cells, and they altered the expression of estrogen-dependent genes [11,12,13,14]. Furthermore, in vivo studies in rodents showed an increase in uterine weight of immature female mice after exposure to isobutylparaben [12] and a testosterone decrease and/or other reproductive system alterations in male rats [15,16,17]. Epidemiological studies in humans revealed a statistically significant correlation between BuPB levels in urine and sperm DNA damage [18]. Furthermore, there was a statistically significant inverse correlation of PB levels in urine and thyroid hormones’ levels in serum, which was more prominent in females [19].
The environmental threats arising from the presence of PBs in WWTPs were noticed quite early (1996) [20]. Nevertheless, they are easily biodegradable in aerobic conditions and as such, can be effortlessly degraded in secondary biological treatment [21], reaching removal rates of up to 96.1% [22] at conventional WWTPs (Table 1). When advanced oxidation processes (AOPs) are applied to the effluents, removal rate may reach up to 99.9% [23].

2.1.2. Triclosan

Triclosan is a broad-spectrum antimicrobial agent (Figure 2) that can be found in numerous PCPs such as soaps, shampoos, deodorants, cosmetic creams, toothpaste, and mouthwash [31,32]. Its mode of action involves inhibition of lipid acids’ biosynthesis and subsequent cellular membrane destabilization [33]. In 1998, triclosan worldwide annual production was approximately 1500 tn, out of which 350 tn were produced in the EU. Triclosan quantities used in the EU were approximately 450 tn in 2006 [34]. In the same countries, it has been estimated that 85% of triclosan quantities were incorporated in PCPs, while only 5% were used in fabrics and 10% in plastics [35].
Triclosan is highly toxic to organisms such as algae and invertebrates [36], a characteristic which is not relevant for human toxicology, nevertheless it is highly relevant from an ecotoxicological point of view. Furthermore, in long-term exposure, algae have been proven to be the most susceptible trophic level, and their development was negatively affected even at concentrations as low as 1 μg L−1 [37].
Regarding mammalian toxicology, triclosan showed both anti-androgenic and anti-estrogenic action in MCF-7 cells in vitro [38]. In vivo experiments in higher vertebrates showed anti-androgenic effects on male Wistar rats [39], quicker sexual maturation in female rats [40], or increased uterine weight in immature rats [41]. Triclosan also reduced T4 (a thyroid hormone) levels in rat serum [42], and there was a concomitant negative effect both on thyroid hormones in dams and on sexual maturation of their progeny [43]. According to a recent epidemiological study [44], maternal urinary levels of triclosan concentrations during pregnancy were inversely correlated with infants’ birth weight, length, head circumference, and gestational age. In addition, triclosan (and parabens) levels in urine of children aged 6–12 years old were positively associated with aeroallergen sensitization and manifestation of atopic asthma [45].
Triclosan is less readily biodegradable than PBs in WWTPs since it is hydrophobic, not particularly volatile, and hydrolytically stable [46]. There are high discrepancies between the removal success rates of triclosan in WWTPs, which are between approximately 17% and 98% (Table 2). According to [46] and [31], triclosan is firstly sorbed onto the suspended solids of activated sludge and afterward, it is directly biodegraded. An in situ experiment in Psitalleia WWTP in Athens showed that 45% of influent triclosan was adsorbed on sewage sludge, 46% was degraded, and the remaining 5% was detected in the effluent [47].

2.1.3. Synthetic Musks

Ingredients used for perfume synthesis in PCPs are brand-protected, and they do not have to be explicitly written on the PCP container. The ingredients are simply stated as “parfum” and they usually contain synthetic musks. These musks add pleasant odor to perfumes, soaps, shampoos, and cosmetic creams and they are also used in other house commodities [55]. They belong to the categories of nitro-, polycyclic-, macrocyclic- or alicyclic musks (Table 3). Most nitro musks are outphased while alicyclic musks are supposed to be more environmentally friendly [56].
Musks are lipophilic, and they may accumulate in adipose tissue and breast milk in mammals [57]. A number of toxicological and ecotoxicological findings have been attributed to nitro-musks; according to the concise review of [57], tumorigenesis and cancer propagation may be facilitated via non-genotoxic mechanisms of nitro-musks. These compounds have also been found to be endocrine disruptors in in vitro models, however, there is still the question of whether the concentrations tested were environmentally relevant.
The polycyclic musk tonalide (AHTN) has shown estrogenic activity in MCF-7 cells [58], anti-estrogenic activity in 293HEK cells [59], and both tonalide and galaxolide have inhibited progesterone and cortisol production in H295R cells [60]. On the other hand, galaxolide did not show estrogenic activity in MCF-7 cells [58] but it did show anti-estrogenic activity in 293HEK cells [59]. According to [61], some compounds may act both as agonists and as antagonists and this depends on the cell line and the type of estrogen receptor tested; these compounds are named “selective estrogen receptor modulators” and tonalide may well be one of them. The “fourth generation of musks” [62] (cyclomusk, helvetolide, romandolide) are not as well studied but may be a promising alternative in the fragrance industry [56].
Due to their lipophilic nature, synthetic musks are not easily biodegradable, and they tend to accumulate [61,63]. These characteristics enhance their persistence in WWTP (Table 4), which were not able to eliminate musks to a high extent [64]. To make matters worse, it has been estimated that 77% of skin applied HHCB [65], and 69.6% of skin applied AHTN [66] end up in sewage.

2.1.4. UV Filters

UV filters are extremely popular ingredients in PCPs since they protect from photo-induced skin damage and aging (Table 5). They belong to various chemical categories, and they can roughly be divided into organic and inorganic (ZnO and TiO2) filters. In the EU, 25 organic and 2 inorganic filters are available in the market [73].
A few studies have highlighted the significant ecotoxicological effect of UV filters on numerous aquatic organisms from invertebrates to fish [74,75]. Of high ecotoxicological importance is their ability (benzophenones, parabens, cinammates, camphor derivatives) to cause coral bleaching [74] with disastrous ecological effects. Regarding human toxicology, [76] concluded that most of them show multiple hormonal activities (estrogenic, anti-estrogenic, androgenic, anti-androgenic) in in vitro receptor-based assays. For example, non-cytotoxic concentrations of various UV-filters in recombinant yeast cells exhibiting human estrogen (hERα) or androgen (hAR) receptor produced multiple agonistic or antagonistic activities [77]. Furthermore, based on in vitro data on MCF-7 cell lines and on in vivo data on female rats, 4MBC and OMC were partial agonists for estrogenic receptors [78]. Finally, some UV filters (benzophenone-3 (BP3), 4-methylbenzylidene camphor (4-MBC), octinoxate (OMC)) have activated inflammatory cytokines in macrophages, and this is a mediator of allergic reactions [79], while epidemiological data have linked UV filters (benzophenones) with allergic dermatitis [80].
WWTPs are the main culprits for water pollution by UV filters since the latter were not successfully biodegraded [81]. According to [82], a conventional Modified Ludzack-Ettinger system of activated sludge showed the following removal efficiencies for UV filters: 99% for octisalate (EHS) and octocrylene (OC), >70% for benzophenone-1 (BP1), benzophenone-3 (BP3) and homosalate (HMS), 38%–77% for 4-methylbenzylidene camphor (4-MBC), 30%–55% for octinoxate (EHMC) and 10%–50% for avobenzone (BMDBM) and ethylhexyl-dimethyl PABA (OD-PABA). Different efficiencies were noted for a sequencing batch reactor secondary treatment, and important differences were also found for disinfection methods (chlorination vs UV irradiation) (Table 6).

2.2. Small Scale PCPs Market Sampling

A small-scale market sampling of PCPs was performed in a large chain supermarket in Athens, Greece, in April 2019. The PCPs collected are shown in the following Table 7.
The ingredients that were not supposed to be the active agent (e.g., retinol in face cream or chlorhexidine in a mouthwash) of the 52 individual items were further coded according to their action in the final product (e.g., whether it acted as an emulsifier, an antioxidant, etc.). According to this classification the categories “preservatives”, “antimicrobial agents”, “parfum” and “UV filters” were investigated further. Each ingredient was sought on the open-access database https://www.ewg.org/skindeep [86]. The database gives a rudimentary hazard assessment of PCPs ingredients, mainly on the toxicological topics of cancer, developmental and reproductive toxicity, sensitization potential and immunotoxicity, and less on ecotoxicity, endocrine disruption, bioaccumulation, etc.

3. Results and Discussion

The 52 individual items contained in total 420 ingredients. Most PCPs contained 20–30 ingredients with a number of exceptions; a face cream contained 59 ingredients besides the active one, while a nail enamel contained 7 ingredients. These ingredients were grouped in the following categories: Softeners, hydrating agents, hygroscopic agents, surfactants, emulsifiers and emulsion stabilizers, cleansing agents, antioxidants, opacity regulators, absorbents, pH buffers, UV filters, parfum agents, solvents, preservatives, antimicrobial agents, antifoam agents, chelates, bulking agents, color additives, viscosity stabilizers, antistatic agents. It must be noted that some ingredients may act in more than one way; the substance benzyl salicylate, which was commonly found here, may act as a preservative as well as a parfum. The results for the categories in question (“preservatives”, “antimicrobial agents”, “parfum” and “UV filters”) are shown in Figure 3.
Some useful qualitative and quantitative deductions can be made from this analysis; regarding preservatives in products commonly found in the Greek market, only 7.7% of the examined products contained PBs. This is somehow expected because parabens are slowly being phased out for marketing reasons, after the ban of [87] on isopropyl, isobutyl and benzyl-paraben, and the opinion of the European Scientific Committee on Consumer Safety on certain uses of propyl and butyl paraben [88]. As such, a lot of products now have a disclaimer “paraben-free” on their container. Thus, it can be safely assumed that their limited presence in WWTP effluents (they are easily biodegradable in most cases) will be further minimized. Nevertheless, another substance found was benzyl salicylate, which is also a parfum, UV filter, and it is supposed to be of moderate hazard (possible ecotoxicity, endocrine disruption) according to [86]. This substance was also present in the influents of two WWTPs in Spain, and it was somewhat successfully removed after secondary treatment [89]. It looks as if the trends in preservatives for PCPs are changing since consumers are clearly not fond of parabens (i.e., the safety of propyl and butyl paraben could not be unequivocally proven for certain uses, [88]). As such, the research should be targeted to the environmental fate and the biodegradability of these ingredients that will replace parabens. Triclosan was also conspicuously absent from our data set. Triclosan has also been phased out of a variety of products in EU, USA, Japan, and Canada; nowadays, in the EU it is not allowed in hygiene products and in detergent soaps [90]. Even though a recent evaluation [91] concluded that unless new data were presented, the current uses of triclosan in cosmetics did not pose a significant risk to humans, it can be assumed that triclosan was not popular and it will continue to decrease in WWTPs effluents. Furthermore, since different limitations exist in EU, USA, and Asian countries [91], multinational companies may not be willing to launch so many different products.
In total, in our products, only 16 out of 52 contained an antimicrobial agent. Out of these, DMDM hydantoin and 2-bromo-2-nitropropane-1,3-diol (bronopol) released formaldehyde locally. There were very little data (bronopol) on their presence in influents, which makes deductions on their removal efficiency problematic [92].
Much less can be deduced from the “Parfum ingredients” category; in 46 out of the 52 products the terms “parfum” or “aroma” were included in the ingredients, referring to aromatic compositions. Since they were industrial formulas not readily available to the public, information may be obtained by contacting the responsible physical or legal person and only on possible adverse effects, according to [73]. As such, it was almost impossible to quantify the hazard posed by the parfum ingredients and to backtrack the source of pollution in WWTPs. As stated before, these substances are lipophilic, not readily biodegradable, and of significant toxicological profile. They are also bioaccumulative, toxic to aquatic wildlife (even the polycyclic musks that are less toxic than the nitro-musks), and they have been detected in high concentrations in surface waters receiving WWTP effluents [36].
For UV filters, the variability noted in the products tested was quite high. Furthermore, these ingredients may be used in numerous combinations (one particular product contained 8 agents). Benzyl salicylate was again the most common ingredient, followed by TiO2. Avobenzone (BM-DBM), which was insufficiently removed in WWTP [82] was also present to a modest extent (7.7%) followed by octinoxate (OMC, EHMC), octocrylene (OC), and others. Octinoxate is a partial agonist for estrogenic receptors [78] and linked to cytokine activation [79]. According to the data presented in Table 6, this substance was moderately removed from conventional WWTPs. The case of avobenzone was further complicated since it forms chloro-metabolites and chlorophenols under disinfection (chlorination, UV-irradiation) [93,94]. These products are generally toxic and little research has been carried out on their environmental fate, mainly on swimming pools and not on WWTP effluents. Furthermore, in the “rest” category, two products contained nanoforms of inorganic UV filters. Nanomaterials showed a completely different environmental fate and behavior than their bulk counterparts; as such, special attention should be paid to their increasing use.
Within this context, it is important not to forget that around 93% of the effluents produced by Greek WWTPs end up in lakes, rivers, and coastal water [95], while simultaneously, many parts of Greece suffer from water shortage, especially in the summer months, primarily in the agricultural sector. Reclamation and reuse of wastewater is a necessity for Greek society, and this situation may be aggravated due to climate change. Therefore, emphasis should be given on priority substances as laid out in [96], however, the possibility of PCPs ingredients in Greek WWTP effluents should be taken into account when reclamation and reuse are considered.

4. Conclusions

PCPs are an important category of emerging water pollutants due to their extensive use and the multitude of ingredients they contain. The ingredients examined in the present manuscript were parabens (preservatives), triclosan (antimicrobial agents), synthetic musks (parfums), and UV filters, which are of significant toxicological (and ecotoxicological) interest. These ingredients are degraded in WWTPs at various removal rates; some of them at unsatisfactory removal rates. A small-scale investigation for PCPs marketed in Greece revealed a decrease in the occurrence of parabens and triclosan. A few reasons are given for this apparent decrease. There was also a high diversity in UV filters, and it is known that some of them (i.e., avobenzone) are rendered more reactive after water disinfection. The occurrence of (possibly persistent) synthetic musks could not be evaluated due to non-disclosure of information. Other ingredients with unknown behaviors in WWTPs were also present. Based on this preliminary research, the following deductions can be made; (a) the multitude and the variability of ingredients found in PCPs (preservatives, emulsifiers, antimicrobial agents, bulking agents, parfums etc.) pose a threat to wastewater treatment and reclamation; (b) some classic ingredients of PCPs may be phased out and may be replaced by new ones, thus the monitoring of PCPs in WWTPs should be updated; (c) non-disclosure of information of the final PCP product makes it difficult to backtrack the pollution source. It is, therefore, important to extend and update research on fate and behavior of PCPs in the environment and especially on how well these substances can be degraded in WWTPs. It is also imperative to choose more environmental-friendly ingredients for PCPs thus that water reuse becomes a viable option that does not undermine public health and environmental quality.

Author Contributions

Data curation, writing—review and editing: C.E., writing—original draft preparation, review and editing, investigation, formal analysis: M.B., conceptualization, methodology, supervision: C.P., data validation, supervision: A.K.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tchobanoglous, G.; Burton, F.L.; Stensel, H.D. Wastewater Engineering: An Overview. In Metcalf & Eddy, Inc. Wastewater Engineering Treatment and Reuse, 4th ed.; McGraw-Hill: New York, NY, USA, 1991; pp. 1–25. [Google Scholar]
  2. Council Directive 91/271/EEC of 21 May 1991 Concerning Urban Waste-Water Treatment. Available online: https://eur-lex.europa.eu/eli/dir/1991/271/oj (accessed on 14 November 2019).
  3. Joint Ministerial Decree 5673/400/1997: Measures and Terms for Municipal Wastewater Treatment. Available online: http://www.elinyae.gr/el/item_details.jsp?cat_id=928&item_id=3776 (accessed on 26 July 2019).
  4. European Commission DG Environment. Evaluation of the Urban Waste Treatment Directive 91/271/EC. Ref. Ares (2017)4989291-12/10/2017. 2017. Available online: https://ec.europa.eu/info/law/better-regulation/initiative/1143/publication/121146 /attachment/090166e5b5b3f8ab_en (accessed on 25 July 2019).
  5. Molins-Delgado, D.; Díaz-Cruz, M.S.; Barcelo, D. Introduction: Personal Care Products in the Aquatic Environment. In Personal Care Products in the Aquatic Environment; Silvia Díaz-Cruz, M.S., Barcelo, D., Eds.; Springer International Publishing: Cham, Switzerland, 2015; pp. 1–34. [Google Scholar] [CrossRef]
  6. Cassardo, C.; Jones, A.A. Managing water in a changing world. Water 2011, 3, 618–628. [Google Scholar] [CrossRef]
  7. Bekyrou, M. Toxic substances in Personal Care Products: The situation internationally. Examples from the Greek market. Master’s Thesis, Hellenic Open University, Patra, Greece, 2019. (In Greek). [Google Scholar]
  8. Montes-Grajales, D.; Fennix-Agudelo, M.; Miranda-Castro, W. Occurrence of personal care products as emerging chemicals of concern in water resources: A review. Sci. Total Environ. 2017, 595, 601–614. [Google Scholar] [CrossRef] [PubMed]
  9. Boberg, J.; Taxvig, C.; Christiansen, S.; Hass, U. Possible endocrine disrupting effects of parabens and their metabolites. Reprod. Toxicol. 2010, 30, 301–312. [Google Scholar] [CrossRef] [PubMed]
  10. Routledge, E.J.; Parker, J.; Odum, J.; Ashby, J.; Sumpter, J.P. Some alkyl hydroxy benzoate preservatives (parabens) are estrogenic. Toxicol. Appl. Pharmacol. 1998, 153, 12–19. [Google Scholar] [CrossRef]
  11. Byford, J.R.; Shaw, L.E.; Drew, M.; Pope, G.S.; Sauer, M.J.; Darbre, P.D. Oestrogenic activity of parabens in MCF7 human breast cancer cells. J. Steroid Biochem. Mol. Biol. 2002, 80, 49–60. [Google Scholar] [CrossRef]
  12. Darbre, P.D.; Byford, J.R.; Shaw, L.E.; Horton, R.A.; Pope, G.S.; Sauer, M.J. Oestrogenic activity of isobutylparaben in vitro and in vivo. J. Appl. Toxicol. 2002, 22, 219–226. [Google Scholar] [CrossRef]
  13. Darbre, P.D.; Byford, J.R.; Shaw, L.E.; Hall, S.; Coldham, N.G.; Pope, G.S.; Sauer, M.J. Oestrogenic activity of benzylparaben. J. Appl. Toxicol. 2003, 23, 43–51. [Google Scholar] [CrossRef]
  14. Okubo, T.; Yokoyama, Y.; Kano, K.; Kano, I. ER-dependent estrogenic activity of parabens assessed by proliferation of human breast cancer MCF-7 cells and expression of ERα and PR. Food Chem. Toxicol. 2001, 39, 1225–1232. [Google Scholar] [CrossRef]
  15. Oishi, S. Effects of butylparaben on the male reproductive system in rats. Toxicol. Ind. Health 2001, 17, 31–39. [Google Scholar] [CrossRef]
  16. Oishi, S. Effects of propyl paraben on the male reproductive system. Food Chem. Toxicol. 2002, 40, 1807–1813. [Google Scholar] [CrossRef]
  17. Oishi, S. Effects of butyl paraben on the male reproductive system in mice. Arch. Toxicol. 2002, 76, 423–429. [Google Scholar] [CrossRef] [PubMed]
  18. Meeker, J.D.; Yang, T.C.; Ye, X.; Calafat, A.M.; Hauser, R.B. Urinary concentrations of parabens and serum hormone levels, semen quality parameters, and sperm DNA damage. Environ. Health Perspect. 2011, 119, 252–257. [Google Scholar] [CrossRef] [PubMed]
  19. Koeppe, E.S.; Ferguson, K.K.; Colacino, J.A.; Meeker, J.D. Relationship between urinary triclosan and paraben concentrations and serum thyroid measures in NHANES 2007-2008. Sci. Total Environ. 2013, 445–446, 299–305. [Google Scholar] [CrossRef] [PubMed]
  20. Paxéus, N. Organic pollutants in the effluents of large wastewater treatment plants in Sweden. Water Res. 1996, 30, 1115–1122. [Google Scholar] [CrossRef]
  21. González-Mariño, I.; Quintana, J.B.; Rodríguez, I.; Cela, R. Evaluation of the occurrence and biodegradation of parabens and halogenated by-products in wastewater by accurate-mass liquid chromatography-quadrupole-time-of-flight-mass spectrometry (LC-QTOF-MS). Water Res. 2011, 45, 6770–6780. [Google Scholar] [CrossRef]
  22. Jonkers, N.; Kohler, H.-P.E.; Dammshäuser, A.; Giger, W. Mass flows of endocrine disruptors in the Glatt River during varying weather conditions. Environ. Pollut. 2009, 157, 714–723. [Google Scholar] [CrossRef]
  23. Li, W.; Shi, Y.; Gao, L.; Liu, J.; Cai, Y. Occurrence, fate and risk assessment of parabens and their chlorinated derivatives in an advanced wastewater treatment plant. J. Hazard. Mater. 2015, 300, 29–38. [Google Scholar] [CrossRef]
  24. Lee, H.-B.; Peart, T.E.; Svoboda, M.L. Determination of endocrine-disrupting phenols, acidic pharmaceuticals, and personal-care products in sewage by solid-phase extraction and gas chromatography-mass spectrometry. J. Chromatogr. A 2005, 1094, 122–129. [Google Scholar] [CrossRef]
  25. Kasprzyk-Hordern, B.; Dinsdale, R.M.; Guwy, A.J. 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 Res. 2009, 43, 363–380. [Google Scholar] [CrossRef]
  26. González-Mariño, I.; Quintana, J.B.; Rodríguez, I.; Cela, R. Simultaneous determination of parabens, triclosan and triclocarban in water by liquid chromatography/electrospray ionisation tandem mass spectrometry. Rapid Commun. Mass Spectrom. 2009, 23, 1756–1766. [Google Scholar] [CrossRef]
  27. Carmona, E.; Andreu, V.; Picó, Y. Occurrence of acidic pharmaceuticals and personal care products in Turia River Basin: From waste to drinking water. Sci. Total Environ. 2014, 484, 53–63. [Google Scholar] [CrossRef] [PubMed]
  28. Yu, Y.; Huang, Q.; Wang, Z.; Zhang, K.; Tang, C.; Cui, J.; Feng, J.; Peng, X. Occurrence and behavior of pharmaceuticals, steroid hormones, and endocrine-disrupting personal care products in wastewater and the recipient river water of the Pearl River Delta, South China. J. Environ. Monit. 2011, 13, 871–878. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, W.; Kannan, K. Fate of parabens and their metabolites in two wastewater treatment plants in New York State, United States. Environ. Sci. Technol. 2016, 50, 1174–1181. [Google Scholar] [CrossRef]
  30. Karthikraj, R.; Vasu, A.K.; Balakrishna, K.; Sinha, R.K.; Kannan, K. Occurrence and fate of parabens and their metabolites in five sewage treatment plants in India. Sci. Total Environ. 2017, 593–594, 592–598. [Google Scholar] [CrossRef] [PubMed]
  31. Bester, K. Triclosan in a sewage treatment process—Balances and monitoring data. Water Res. 2003, 37, 3891–3896. [Google Scholar] [CrossRef]
  32. Huang, C.-L.; Abass, O.K.; Yu, C.-P. Triclosan: A review on systematic risk assessment and control from the perspective of substance flow analysis. Sci. Total Environ. 2016, 566–567, 771–785. [Google Scholar] [CrossRef]
  33. Dann, A.B.; Hontela, A. Triclosan: Environmental exposure, toxicity and mechanisms of action. J. Appl. Toxicol. 2011, 31, 285–311. [Google Scholar] [CrossRef]
  34. Singer, H.; Müller, S.; Tixier, C.; Pillonel, L. Triclosan: Occurrence and fate of a widely used biocide in the aquatic environment: Field measurements in wastewater treatment plants, surface waters, and lake sediments. Environ. Sci. Technol. 2002, 36, 4998–5004. [Google Scholar] [CrossRef]
  35. SCCP (Scientific Committee on Consumer Products). Opinion on Triclosan, COLIPA n° P32, 2009, (SCCP/1192/08). Available online: https://ec.europa.eu/health/ph_risk/committees/04_sccp/docs/sccp_o_166.pdf (accessed on 14 November 2019).
  36. Brausch, J.M.; Rand, G.M. A review of personal care products in the aquatic environment: Environmental concentrations and toxicity. Chemosphere 2011, 82, 1518–1532. [Google Scholar] [CrossRef]
  37. Orvos, D.R.; Versteeg, D.J.; Inauen, J.; Capdevielle, M.; Rothenstein, A.; Cunningham, V. Aquatic toxicity of triclosan. Environ. Toxicol. Chem. 2002, 21, 1338–1349. [Google Scholar] [CrossRef]
  38. Gee, R.H.; Charles, A.; Taylor, N.; Darbre, P.D. Oestrogenic and androgenic activity of triclosan in breast cancer cells. J. Appl. Toxicol. 2008, 28, 78–91. [Google Scholar] [CrossRef] [PubMed]
  39. Kumar, V.; Chakraborty, A.; Kural, M.R.; Roy, P. Alteration of testicular steroidogenesis and histopathology of reproductive system in male rats treated with triclosan. Reprod. Toxicol. 2009, 27, 177–185. [Google Scholar] [CrossRef] [PubMed]
  40. Stoker, T.E.; Gibson, E.K.; Zorrilla, L.M. Triclosan exposure modulates estrogen-dependent responses in the female Wistar rat. Toxicol. Sci. 2010, 117, 45–53. [Google Scholar] [CrossRef] [PubMed]
  41. Jung, E.-M.; An, B.-S.; Choi, K.-C.; Jeung, E.-B. Potential estrogenic activity of triclosan in the uterus of immature rats and rat pituitary GH3 cells. Toxicol. Lett. 2012, 208, 142–148. [Google Scholar] [CrossRef] [PubMed]
  42. Crofton, K.M.; Paul, K.B.; Devito, M.J.; Hedge, J.M. Short-term in vivo exposure to the water contaminant triclosan: Evidence for disruption of thyroxine. Environ. Toxicol. Pharmacol. 2007, 24, 194–197. [Google Scholar] [CrossRef]
  43. Rodríguez, P.E.A.; Sanchez, M.S. Maternal exposure to triclosan impairs thyroid homeostasis and female pubertal development in Wistar rat offspring. J. Toxicol. Environ. Health 2010, 73, 1678–1688. [Google Scholar] [CrossRef]
  44. Etzel, T.M.; Calafat, A.M.; Ye, X.; Chen, A.; Lanphear, B.P.; Savitz, D.A.; Braun, J. Urinary triclosan concentrations during pregnancy and birth outcomes. Environ. Res. 2017, 156, 505–511. [Google Scholar] [CrossRef]
  45. Spanier, A.J.; Fausnight, T.; Camacho, T.F.; Braun, J. The associations of triclosan and paraben exposure with allergen sensitization and wheeze in children. Allergy Asthma Proc. 2014, 35, 475–481. [Google Scholar] [CrossRef] [Green Version]
  46. Stasinakis, A.; Petalas, A.; Mamais, D.; Thomaidis, N.S.; Gatidou, G.; Lekkas, T.D. Investigation of triclosan fate and toxicity in continuous-flow activated sludge systems. Chemosphere 2007, 68, 375–381. [Google Scholar] [CrossRef]
  47. Stasinakis, A.; Gatidou, G.; Mamais, D.; Thomaidis, N.S.; Lekkas, T.D. Occurrence and fate of endocrine disrupters in Greek sewage treatment plants. Water Res. 2008, 42, 1796–1804. [Google Scholar] [CrossRef]
  48. Stamatis, N.; Konstantinou, I. Occurrence and removal of emerging pharmaceutical, personal care compounds and caffeine tracer in municipal sewage treatment plant in Western Greece. J. Environ. Sci. Health B 2013, 48, 800–813. [Google Scholar] [CrossRef] [PubMed]
  49. Kosma, C.I.; Lambropoulou, D.A.; Albanis, T.A. Investigation of PPCPs in wastewater treatment plants in Greece: Occurrence, removal and environmental risk assessment. Sci. Total Environ. 2014, 466–467, 421–438. [Google Scholar] [CrossRef] [PubMed]
  50. Stasinakis, A.; Thomaidis, N.S.; Arvaniti, O.S.; Asimakopoulos, A.G.; Samaras, V.G.; Ajibola, A.; Mamais, D.; Lekkas, T.D. Contribution of primary and secondary treatment on the removal of benzothiazoles, benzotriazoles, endocrine disruptors, pharmaceuticals and perfluorinated compounds in a sewage treatment plant. Sci. Total Environ. 2013, 463–464, 1067–1075. [Google Scholar] [CrossRef] [PubMed]
  51. McAvoy, D.C.; Schatowitz, B.; Jacob, M.; Hauk, A.; Eckhoff, W.S. Measurement of triclosan in wastewater treatment systems. Environ. Toxicol. Chem. 2002, 21, 1323–1329. [Google Scholar] [CrossRef]
  52. Nakada, N.; Tanishima, T.; Shinohara, H.; Kiri, K.; Takada, H. Pharmaceutical chemicals and endocrine disrupters in municipal wastewater in Tokyo and their removal during activated sludge treatment. Water Res. 2006, 40, 3297–3303. [Google Scholar] [CrossRef]
  53. Yu, Y.; Wu, L.; Chang, A.C. Seasonal variation of endocrine disrupting compounds, pharmaceuticals and personal care products in wastewater treatment plants. Sci. Total Environ. 2013, 442, 310–331. [Google Scholar] [CrossRef]
  54. Kim, J.-W.; Ishibashi, H.; Hirano, M.; Jang, H.-S.; Kim, J.-G.; Takao, Y.; Ichikawa, N.; Shinohara, R.; Arizono, K. Contamination of pharmaceutical and personal care products in sewage treatment plants and surface waters in South Korea and their removal during activated sludge treatment. J. Environ. Chem. 2010, 20, 127–135. [Google Scholar] [CrossRef] [Green Version]
  55. Homem, V.; Silva, J.S.; Ratola, N.; Santos, L.; Alves, A. Long lasting perfume A review of synthetic musks in WWTPs. J. Environ. Manag. 2015, 149, 168–192. [Google Scholar] [CrossRef] [Green Version]
  56. Li, X.; Chu, Z.; Yang, J.; Li, M.; Du, M.; Zhao, X.; Zhu, Z.; Li, Y. A Class of Commercial Fragrance Additives in Personal Care Products (PCPs) Causing Concern as Emerging Contaminants. In Advances in Marine Biology; Chen, B., Zhang, B., Zhu, Z., Lee, K., Eds.; Academic Press: Cambridge, MA, USA, 2018; Volume 81, pp. 213–280. [Google Scholar] [CrossRef]
  57. Taylor, K.M.; Weisskopf, M.; Shine, J. Human exposure to nitro musks and the evaluation of their potential toxicity: An overview. Environ. Health. 2014, 13, 14. [Google Scholar] [CrossRef] [Green Version]
  58. Bitsch, N.; Dudas, C.; Körner, W.; Failing, K.; Biselli, S.; Rimkus, G.; Brunn, H. Estrogenic activity of musk fragrances detected by the E-screen assay using human mcf-7 cells. Arch. Environ. Contam. Toxicol. 2002, 43, 257–264. [Google Scholar] [CrossRef]
  59. Schreurs, R.H.M.M.; Sonneveld, E.; Jansen, J.H.J.; Seinen, W.; van der Burg, B. Interaction of Polycyclic Musks and UV Filters with the Estrogen Receptor (ER), Androgen Receptor (AR), and Progesterone Receptor (PR) in Reporter Gene Bioassays. Toxicol. Sci. 2005, 83, 264–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Li, Z.; Yin, N.; Liu, Q.; Wang, C.; Wang, T.; Wang, Y.; Qu, G.; Liu, J.; Cai, Y.; Zhou, Q.; et al. Effects of polycyclic musks HHCB and AHTN on steroidogenesis in H295R cells. Chemosphere 2013, 90, 1227–1235. [Google Scholar] [CrossRef] [PubMed]
  61. Schreurs, R.H.M.M.; Quaedackers, M.E.; Seinen, W.; van der Burg, B. Transcriptional activation of estrogen receptor ERα and ERβ by polycyclic musks is cell type dependent. Toxicol. Appl. Pharmacol. 2002, 183, 1–9. [Google Scholar] [CrossRef] [PubMed]
  62. Eh, M. New alicyclic musks: The fourth generation of musk odorants. Chem. Biodivers. 2004, 1, 1975–1984. [Google Scholar] [CrossRef]
  63. Simmons, D.B.D.; Marlatt, V.L.; Trudeau, V.L.; Sherry, J.P.; Metcalfe, C.D. Interaction of Galaxolide® with the human and trout estrogen receptor-α. Sci. Total Environ. 2010, 408, 6158–6164. [Google Scholar] [CrossRef]
  64. Vallecillos, L.; Borrull, F.; Pocurull, E. On-line coupling of solid-phase extraction to gas chromatography-mass spectrometry to determine musk fragrances in wastewater. J. Chromatogr. A 2014, 1364, 1–11. [Google Scholar] [CrossRef]
  65. HERA. Risk Assessment of HHCB (1,3,4,6,7,8-hexahydro-4,6,6,7,8,8-hexamethylcyclopenta- γ-2-benzopyran and related isomers). HERA: Human and Environmental Risk Assessment on Ingredients of Household Cleaning Products, 1–62. 2004. Available online: https://www.heraproject.com/files/29-HH-04-pcm%20HHCB%20HERA %20Human%20Health%20DISCL%20ed2.pdf (accessed on 25 July 2019).
  66. HERA. Risk Assessment of AHTN (6-Acetyl-1,1,2,4,4,7-hexamethyltetraline). Human and Environmental Risk Assessment on Ingredients of Household Cleaning Products. 2004. Available online: https://www.heraproject.com/files/28-HH-04-pcm%20AHTN%20HERA%20Human%20Health%20DISCL%20ed2.pdf (accessed on 25 July 2019).
  67. Chase, D.A.; Karnjanapiboonwong, A.; Fang, Y.; Cobb, G.P.; Morse, A.N.; Anderson, T.A. Occurrence of synthetic musk fragrances in effluent and non-effluent impacted environments. Sci. Total Environ. 2012, 416, 253–260. [Google Scholar] [CrossRef]
  68. Hu, Z.; Yali, S.; Zhang, S.; Niu, H.; Cai, Y. Assessment of synthetic musk fragrances in seven wastewater treatment plants of Beijing, China. Bull. Environ. Contam. Toxicol. 2011, 86, 302–306. [Google Scholar] [CrossRef]
  69. Lv, Y.; Yuan, T.; Hu, J.; Wang, W. Seasonal occurrence and behavior of synthetic musks (SMs) during wastewater treatment process in Shanghai, China. Sci. Total Environ. 2010, 408, 4170–4176. [Google Scholar] [CrossRef]
  70. Kupper, T.; Plagellat, C.; Brändli, R.; de Alencastro, L.F.; Grandjean, D.; Tarradellas, J. Fate and removal of polycyclic musks, UV filters and biocides during wastewater treatment. Water Res. 2006, 40, 2603–2612. [Google Scholar] [CrossRef]
  71. Yang, J.-J.; Metcalfe, C.D. Fate of synthetic musks in a domestic wastewater treatment plant and in an agricultural field amended with biosolids. Sci. Total Environ. 2006, 363, 149–165. [Google Scholar] [CrossRef] [PubMed]
  72. Artola-Garicano, E.; Borkent, I.; Hermens, J.L.M.; Vaes, W.H.J. Removal of two polycyclic musks in sewage treatment plants: Freely dissolved and total concentrations. Environ. Sci. Technol. 2003, 37, 3111–3116. [Google Scholar] [CrossRef] [PubMed]
  73. Regulation (EC) No 1223/2009 of the European Parliament and of the Council of 30 November 2009 on Cosmetic Products. Available online: https://eur-lex.europa.eu/legal-content/EN/ALL/?uri=CELEX%3A32009R1223 (accessed on 25 July 2019).
  74. Danovaro, R.; Bongiorni, L.; Corinaldesi, C.; Giovannelli, D.; Damiani, E.; Astolfi, P.; Pusceddu, A. Sunscreens Cause Coral Bleaching by Promoting Viral Infections. Environ. Health Perspect. 2008, 116, 441–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Kim, S.; Jung, D.; Kho, Y.; Choi, K. Effects of benzophenone-3 exposure on endocrine disruption and reproduction of Japanese medaka (Oryzias latipes)—A two generation exposure study. Aquat. Toxicol. 2014, 155, 244–252. [Google Scholar] [CrossRef]
  76. Díaz-Cruz, M.S.; Barceló, D. Chemical analysis and ecotoxicological effects of organic UV-absorbing compounds in aquatic ecosystems. Trends Anal. Chem. 2009, 28, 194–197. [Google Scholar] [CrossRef]
  77. Kunz, P.Y.; Fent, K. Multiple hormonal activities of UV filters and comparison of in vivo and in vitro estrogenic activity of ethyl-4-aminobenzoate in fish. Aquat. Toxicol. 2006, 79, 305–324. [Google Scholar] [CrossRef]
  78. Schlumpf, M.; Cotton, B.; Conscience, M.; Haller, V.; Steinmann, B.; Lichtensteiger, W. In vitro and in vivo estrogenicity of UV screens. Environ. Health Perspect. 2001, 109, 239–244. [Google Scholar] [CrossRef]
  79. Ao, J.; Yuan, T.; Gao, L.; Yu, X.; Zhao, X.; Tian, Y.; Ding, W. Organic UV filters exposure induces the production of inflammatory cytokines in human macrophages. Sci. Total Environ. 2018, 635, 926–935. [Google Scholar] [CrossRef]
  80. Goossens, A. Contact-Allergic Reactions to Cosmetics. J. Allergy 2011, 1–6. [Google Scholar] [CrossRef] [Green Version]
  81. Gago-Ferrero, P.; Mastroianni, N.; Díaz-Cruz, M.S.; Barceló, D. Fully automated determination of nine ultraviolet filters and transformation products in natural waters and wastewaters by on-line solid phase extraction–liquid chromatography–tandem mass spectrometry. J. Chromatogr. A 2013, 1294, 106–116. [Google Scholar] [CrossRef]
  82. Tsui, M.M.P.; Leung, H.W.; Lam, P.K.S.; Murphy, M.B. Seasonal occurrence, removal efficiencies and preliminary risk assessment of multiple classes of organic UV filters in wastewater treatment plants. Water Res. 2014, 53, 58–67. [Google Scholar] [CrossRef] [PubMed]
  83. Cunha, S.; Pena, A.; Fernandes, J.O. Dispersive liquid-liquid microextraction followed by microwave-assisted silylation and gas chromatography-mass spectrometry analysis for simultaneous trace quantification of bisphenol A and 13 ultraviolet filters in wastewaters. J. Chromatogr. A 2015, 1414, 10–21. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, Y.-S.; Ying, G.G.; Shareef, A.; Kookana, R.S. Occurrence and removal of benzotriazoles and ultraviolet filters in a municipal wastewater treatment plant. Environ. Pollut. 2012, 165, 225–232. [Google Scholar] [CrossRef] [PubMed]
  85. Balmer, M.E.; Buser, H.-R.; Müller, M.D.; Poiger, T. Occurrence of some organic UV filters in wastewater, in surface waters, and in fish from Swiss lakes. Environ. Sci. Technol. 2005, 39, 953–962. [Google Scholar] [CrossRef]
  86. Environmental Working Group. 2019. Available online: https://www.ewg.org/skindeep (accessed on 14 November 2019).
  87. Commission Regulation (EU) No 358/2014 of 9 April 2014 amending Annexes II and V to Regulation (EC) No 1223/2009 of the European Parliament and of the Council on cosmetic products. Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=celex%3A32014R0358 (accessed on 14 November 2019).
  88. SCCS (Scientific Committee on Consumer Safety). Opinion on Parabens, COLIPA No 582, 2013, SCCS/1514/13. Available online: https://ec.europa.eu/health/scientific_committees/consumer_safety/docs/sccs_o_132.pdf (accessed on 14 November 2019).
  89. Godayol, A.; Besalú, E.; Anticó, E.; Sanchez, J.M. Monitoring of sixteen fragrance allergens and two polycyclic musks in wastewater treatment plants by solid phase microextraction coupled to gas chromatography. Chemosphere 2015, 119, 363–370. [Google Scholar] [CrossRef]
  90. Commission Implementing Decision (EU) 2016/110 of 27 January 2016 Not Approving Triclosan as an Existing Active Substance for Use in Biocidal Products for Product Type 1. Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=celex%3A32014R0358 (accessed on 14 November 2019).
  91. Lee, J.D.; Lee, J.Y.; Kwack, S.J.; Shin, C.Y.; Jang, H.J.; Kim, H.Y.; Kim, M.K.; Seo, D.; Lee, B.M.; Kim, K.B. Risk assessment of triclosan, a cosmetic preservative. Toxicol. Res. 2019, 35, 137–154. [Google Scholar] [CrossRef]
  92. Carbajo, J.B.; Perdigón-Melón, J.A.; Petre, A.L.; Rosal, R.; Letón, P.; García-Calvo, E. Personal care product preservatives: Risk assessment and mixture toxicities with an industrial wastewater. Water Res. 2015, 72, 174–185. [Google Scholar] [CrossRef]
  93. Crista, D.M.A.; Miranda, M.S.; Esteves da Silva, J.C.G.E. Degradation in chlorinated water of the UV filter 4-tert-butyl-4′-methoxydibenzoylmethane present in commercial sunscreens. Environ. Technol. 2015, 36, 1319–1326. [Google Scholar] [CrossRef]
  94. Trebše, P.; Polyakova, O.V.; Baranova, M.; Kralj, M.; Dolenc, D.; Sarakha, M.; Kutin, A.; Lebedev, A.T. Transformation of avobenzone in conditions of aquatic chlorination and UV-irradiation. Water Res. 2016, 101, 95–102. [Google Scholar] [CrossRef]
  95. Kalavrouziotis, I. Wastewater Reuse Planning in Agriculture: The Case of Aitoloakarnania, Western Greece. Water 2011, 3, 988–1004. [Google Scholar] [CrossRef] [Green Version]
  96. Directive 2000/60/EC of the European Parliament and of the Council of 23 October 2000 Establishing a Framework for Community Action in the Field of Water Policy. Available online: https://eur-lex.europa.eu/legal-content/en/ALL/?uri=CELEX:32000L0060 (accessed on 14 November 2019).
Figure 1. Chemical structure of most common parabens (ag) (PBs) used in cosmetics (drawn via ACD/Chemsketch).
Figure 1. Chemical structure of most common parabens (ag) (PBs) used in cosmetics (drawn via ACD/Chemsketch).
Water 11 02501 g001
Figure 2. Chemical structure of triclosan (drawn via ACD/Chemsketch).
Figure 2. Chemical structure of triclosan (drawn via ACD/Chemsketch).
Water 11 02501 g002
Figure 3. Percentage (%) of individual ingredients found in the 52 examined PCPs (a) preservatives in PCPs, Key: PE: Phenoxyethanol, BA: Benzyl alcohol, SD: Sodium benzoate, BS: Benzyl salicylate BHT: Butylated hydroxytoluene PS: Potassium sorbate E: Ethylhexylglycerin, B: Benzoic acid, M: Methylisothiazolinone, BE: Benzyl benzoate, ME: Methylchloroisothiazolinone, PBS: Parabens, CC: Cetrimonium chloride, D: Dehydroacetic acid, R: Rest, (b) antimicrobial agents in PCPs, Key: AD: Alcohol denatured, DMDM: DMDM hydantoin, AB: Alkyl benzoate, 2B: 2-bromo-2-nitropropane-1,3-diol A: Alcohol, SO: Saponaria officinalis extr, (c) perfum ingredients in PCPs, Key: PA: parfum, HC: Hexyl cinnamal, BS: Benzyl salicylate, BM: Lilial, L: Linalool A: Alpha-isomethyl ionone, LI: Limonene, C: Citronellol, CO: Coumarin, G: Geraniol, CI: Citral, R: Rest, (d) UV filters in PCPs, Key: BS: Benzyl salicylate, TI: Titanium dioxide, BM-DBM: Avobenzone, OMC: Octinoxate, EHS: Octisalate, OC: Octocrylene, BP4: Benzophenone, R: Rest. In bold are shown the ingredients analyzed in Section 2.1.1, Section 2.1.2, Section 2.1.3 and Section 2.1.4.
Figure 3. Percentage (%) of individual ingredients found in the 52 examined PCPs (a) preservatives in PCPs, Key: PE: Phenoxyethanol, BA: Benzyl alcohol, SD: Sodium benzoate, BS: Benzyl salicylate BHT: Butylated hydroxytoluene PS: Potassium sorbate E: Ethylhexylglycerin, B: Benzoic acid, M: Methylisothiazolinone, BE: Benzyl benzoate, ME: Methylchloroisothiazolinone, PBS: Parabens, CC: Cetrimonium chloride, D: Dehydroacetic acid, R: Rest, (b) antimicrobial agents in PCPs, Key: AD: Alcohol denatured, DMDM: DMDM hydantoin, AB: Alkyl benzoate, 2B: 2-bromo-2-nitropropane-1,3-diol A: Alcohol, SO: Saponaria officinalis extr, (c) perfum ingredients in PCPs, Key: PA: parfum, HC: Hexyl cinnamal, BS: Benzyl salicylate, BM: Lilial, L: Linalool A: Alpha-isomethyl ionone, LI: Limonene, C: Citronellol, CO: Coumarin, G: Geraniol, CI: Citral, R: Rest, (d) UV filters in PCPs, Key: BS: Benzyl salicylate, TI: Titanium dioxide, BM-DBM: Avobenzone, OMC: Octinoxate, EHS: Octisalate, OC: Octocrylene, BP4: Benzophenone, R: Rest. In bold are shown the ingredients analyzed in Section 2.1.1, Section 2.1.2, Section 2.1.3 and Section 2.1.4.
Water 11 02501 g003aWater 11 02501 g003b
Table 1. Occurrence of PBs in influents and effluents of wastewater treatment plants (WWTPs) worldwide.
Table 1. Occurrence of PBs in influents and effluents of wastewater treatment plants (WWTPs) worldwide.
RegionSubstrateMePB
ng L−1
EtPB
ng L−1
PrPB
ng L−1
BuPB
ng L−1
Reference
SwedenEffluentn.d.−300<100–200<100–300-[20]
Ontario, CanadaInfluent
Effluent
100–1470
20–30
20–270
<10
200–2430
<10–40
20–260
<10–10
[24]
SwitzerlandInfluent
Effluent
65–9880
4.6–423
2.2–719
<0.3–17
43–1540
<0.5–28
9.7–864
<0.2–12
[22]
Wales,
United Kingdom
Influent
Effluent
661–30,688
<3–155
192–3312
<0.6–69
<2–8286
<1–95
<2–1595
<1–2
[25]
Galicia,
Spain
Influent
Effluent
1926–5138
<n.d.–1.5
452–549
<n.d.
1147–1302
<n.d.
150–181
<n.d.–3.6
[26]
Northwest SpainInfluent
Effluent
290–10,000
6.1–50
250–1600
n.d–9.8
520–2800
n.d.–21
39–270
n.d.
[21]
Valencia,
Spain
Influent
Effluent
334 (av.)
11 (av.)
n.d.
72 (av.)
163 (av.)
n.d.
15 (av.)
n.d.
[27]
Pearl River Delta, ChinaInfluent
Effluent
1002–1140
5.1–7.6
156.2–166.2
0.9–1
499.7–579.8
7.2–10.6
14.9–26.8
0.3
[28]
Beijing,
China
Influent
Effluent
211–1002
2.14–15
32.9–220
0.08–0.70
287–605
0.04–0.99
16.6–35.5
0.02–0.12
[23]
New York,
USA
Influent
Effluent
18.3–320
0.14–1.73
0.50–66.80
0.14–1.47
8.19–113
0.14–490
3.15–112
0.14–3.55
[29]
Saidpur, Beur, Coimbatore, Udupi, Manipal,
India
Influent
Effluent
51–267
4.4–41
11.6–58.4
1.9–9.8
38.2–583
2.8–19.3
4.1–10.5
n.d.–2.9
[30]
n.d.: Not detected, av: Average. MePB: Methylparaben, EtPB: Ethylparaben, PrPB: Propylparaben, BuPB: Butylparaben.
Table 2. Occurrence of triclosan in influents and effluents of wastewater treatment plants (WWTPs) worldwide.
Table 2. Occurrence of triclosan in influents and effluents of wastewater treatment plants (WWTPs) worldwide.
RegionSubstrateConcentration ng L−1Reference
Agrinio,
Greece
Influent
Effluent
65.3–303
24.9–87
[48]
Ioannina, Ioannina hospital, Arta, Preveza, Agrinio, Grevena, Kozani, Veroia,
Greece
Influent
Effluent
n.d.–1742.5
n.d.–452.1
[49]
Athens, Chalkida, Mytilene, Herakleio, Nafplio, (also from an airport, a university and a hospital),
Greece
Influent
Effluent
170–23,900
<130–6880
[47]
Athens,
Greece
Influent
Effluent
328–1096
31–211
[50]
Dortmund
Germany
Influent
Effluent
1200 (av.)
51 (av.)
[31]
Colombus, Glendale, Loveland, West Union I, West Union II,
USA (OH)
Influent
Effluent
3830–16,600
240–2700
[51]
Tokyo (5 individual WWTPs),
Japan
Influent
Effluent
219–1020
26.6–330
[52]
California,
USA
Influent
Effluent
180–4400
n.d.–160
[53]
Jeonju,
S. Korea
Influent
Effluent
280–745
n.d.–29.6
[54]
n.d.: Not detected, av: Average.
Table 3. Categories and chemical formulas of most commonly used synthetic musks (drawn via ACD/Chemsketch).
Table 3. Categories and chemical formulas of most commonly used synthetic musks (drawn via ACD/Chemsketch).
CategoryChemical Name Chemical Formula
Nitro-musksmusk tibetene, MT Water 11 02501 i001
musk ambrette, MA Water 11 02501 i002
musk moskene, MM Water 11 02501 i003
musk ketone, MK Water 11 02501 i004
musk xylene, MX Water 11 02501 i005
Polycyclic muskstonalide, AHTN Water 11 02501 i006
galaxolide, HHCB Water 11 02501 i007
traseolide, ATII Water 11 02501 i008
cashmeran, DPMI Water 11 02501 i009
celestolide, ADBI Water 11 02501 i010
phantolide, AHMI Water 11 02501 i011
Macrocyclic musksmuscone Water 11 02501 i012
exaltolide Water 11 02501 i013
ambrettolide Water 11 02501 i014
Alicyclic muskscyclomusk Water 11 02501 i015
helvetolide Water 11 02501 i016
romandolide Water 11 02501 i017
Table 4. Οccurrence of synthetic musks in influents and effluents of WWTPs worldwide.
Table 4. Οccurrence of synthetic musks in influents and effluents of WWTPs worldwide.
RegionSubstrateHHCB
ng L−1
AHTN
ng L−1
MX
ng L−1
MK
ng L−1
Reference
Catalonia,
Spain
Influent
Effluent
818–45,091
1.5–900
852–49,904
2–75,555
n.d.–632
n.d
n.d.–4110
n.d.–465
[64]
Τexas,
USA
Influent
Effluent
4772–13,399
2989–10,525
509–2337
328–1754
n.d.
n.d.
n.d.–812
n.d.–177
[67]
Beijing,
China
Influent
Effluent
30.9–3039
30.4–685.6
28.6–1486.1
14.26–195.3
<1.2–22.95
n.d.
52.25–165.8
22.77–91.6
[68]
Shanghai,
China
Influent
Effluent
1478–2214
181.1–242.2
553.5–1037.7
46.7–88.3
63–164
n.d.–6.5
74.5–161.3
6.7–18.5
[69]
Mittlere,
Switzerland
Influent
Effluent
2290–6810
570–1030
1130–2000
190–500
--[70]
Ontario,
Canada
Influent
Effluent
246.7–567.5
138.8–234
47.2–136.7
24.7–62.8
10.8–16
4.3–7.6
12.5–15.5
8.3–8.5
[71]
Utrecht,
Netherlands
Influent
Effluent
1420–4300
1250–2220
540–1760
420–1200
--[72]
n.d.: Not detected, av: Average, HHCB: Galaxolide, AHTN: Tonalide, MX: Musk xylene, MK: Musk ketone. *HHCB AHTN: Polycyclic musks, MX MK: Nitro-musks.
Table 5. Categories and chemical formulas of most commonly used UV filters (drawn via ACD/Chemsketch).
Table 5. Categories and chemical formulas of most commonly used UV filters (drawn via ACD/Chemsketch).
CategoryChemical NameChemical Formula
Benzophenonesbenzophenone-1, BP1 Water 11 02501 i018
benzophenone-2, BP2 Water 11 02501 i019
benzophenone-3, BP3 Water 11 02501 i020
benzophenone-4, BP4 Water 11 02501 i021
p-aminobenzoic acid derivativesethylhexyl dimethyl PABA, OD- PABA Water 11 02501 i022
ethyl ester PABA, Et-PABA Water 11 02501 i023
salicylateshomosalate, HMS Water 11 02501 i024
octisalate, EHS, OS Water 11 02501 i025
cinnamatesethylhexyl methoxycinnamate, octinoxate, EHMC, OMC Water 11 02501 i026
camphor derivatives4-methylbenzylidene camphor, 4-MBC Water 11 02501 i027
3-benzylidene camphor, 3-BC
(banned)
Water 11 02501 i028
dibenzoylmethane derivativesavobenzone, BM-DBM Water 11 02501 i029
Crylene derivativesoctocrylene, OC Water 11 02501 i030
benzotriazoles1-hydroxybenzotriazole, HBT Water 11 02501 i031
UVP Water 11 02501 i032
UV326 Water 11 02501 i033
UV327 Water 11 02501 i034
Table 6. Occurrence of UV filters in influents and effluents of WWTPs worldwide.
Table 6. Occurrence of UV filters in influents and effluents of WWTPs worldwide.
RegionSubstrateBP3
ng L−1
OD-PABA
ng L−1
EHMC
ng L−1
OC
ng L−1
BM-DBM
ng L−1
4MBC
ng L−1
HMS
ng L−1
Reference
15 WWTPs
Portugal
Influent
Effluent
5.4–323.3
<10–136
12.2–418
<10
33–689.5
35–483.4
<10–689
125–357
<50–2935
<50–168.1
45.8–155
<50
-[83]
5 WWTPs
Hong Kong
China
Influent
Effluent
114–450.7
18.4–541.1
<0.31–259
<11–224
50–1134
<0.85–505.2
<66–131
< 5.91
35–1290.2
<0.44–1018.3
<3.5–350
<4.58–207.2
<39–1650
<3.75–154.2
[82]
5 WWTPs
Catalonia,
Spain
Influent
Effluent
75.6–306
7.71–34
----<33–48.3
n.d.–23.8
-[81]
2 WWTPs,
Wales,
United Kingdom
Influent
Effluent
<104,000–3,975,000
<80,000–2,196,000
------[25]
Adelaide,
Australia
Influent
Effluent
1059–3112
54–488
-106–319
<0.7–53
88–89
<3.4–73
-394–406
17–140
-[84]
8 WWTPs
Zurich,
Switzerland
Influent
Effluent
700–7800
<10–700
-500–19,000
<10–100
100–12,000
<10–300
-600–6500
60–2700
-[85]
PB3: Benzophenone-3, OD-PABA: Ethylhexyl dimethyl PABA, EHMC: Ethylhexyl methoxycinnamate, OC: Octocrylene, BM-DBM: Avobenzone, 4MBC: 4-methylbenzylidene camphor, HMS: Homosalate.
Table 7. Categories of collected personal care products (PCPs).
Table 7. Categories of collected personal care products (PCPs).
No of Brands/PCP CategoryNo of Brand
Shampoo6
Soap bar1
Liquid bath soap8
Liquid hand soap3
Face cleansing gel1
Face cleansing emulsion1
Face cleansing balsam1
Face cream3
Eye cream1
Body cream2
Hair conditioner2
Hand cream2
Body lotion1
Sunscreen3
Sun protecting hair oil1
Hair oil1
Hair mousse1
Hair gel1
Hair styling mud2
Stick deodorant1
Eau de toilette1
Eau de parfum3
Nail enamel1
Make-up removal wet tissues1
Lipstick1
Shaving mousse1
Insect repellent lotion1
Toothpaste 1

Share and Cite

MDPI and ACS Style

Emmanouil, C.; Bekyrou, M.; Psomopoulos, C.; Kungolos, A. An Insight into Ingredients of Toxicological Interest in Personal Care Products and A Small–Scale Sampling Survey of the Greek Market: Delineating a Potential Contamination Source for Water Resources. Water 2019, 11, 2501. https://doi.org/10.3390/w11122501

AMA Style

Emmanouil C, Bekyrou M, Psomopoulos C, Kungolos A. An Insight into Ingredients of Toxicological Interest in Personal Care Products and A Small–Scale Sampling Survey of the Greek Market: Delineating a Potential Contamination Source for Water Resources. Water. 2019; 11(12):2501. https://doi.org/10.3390/w11122501

Chicago/Turabian Style

Emmanouil, C., M. Bekyrou, C. Psomopoulos, and A. Kungolos. 2019. "An Insight into Ingredients of Toxicological Interest in Personal Care Products and A Small–Scale Sampling Survey of the Greek Market: Delineating a Potential Contamination Source for Water Resources" Water 11, no. 12: 2501. https://doi.org/10.3390/w11122501

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop