Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-27T22:01:30.928Z Has data issue: false hasContentIssue false

6 - Biocides from Marine Coatings

Published online by Cambridge University Press:  22 January 2021

Stephen de Mora
Affiliation:
Plymouth Marine Laboratory
Timothy Fileman
Affiliation:
Plymouth Marine Laboratory
Thomas Vance
Affiliation:
Plymouth Marine Laboratory
Get access

Summary

Colonization by fouling organisms is a problem that has challenged operators of ships since humans first took to the seas. Reducing or preventing the fouling of a ship’s hull is important to allow the vessel to pass efficiently through the water. For centuries, fouling has been controlled through the application of a coating that discourages fouling organisms. As early as the third century there are reports of the Greeks using tar and wax to coat ships’ hulls, and as early as the sixteenth century there are reports of copper sheathing or mixtures containing arsenic being used as anti-fouling (AF) coatings on wooden ships. Many of these and the AF solutions that followed were based on the presence of a toxin in the paint to deter organisms from colonizing the painted surface. Cuprous oxide has been used as a biocide since the early nineteenth century and continues to be a common component of modern AF products. The twentieth century saw the introduction of contact leaching AF coatings designed to increase the efficacy and active lifetime of the coating. Typically, these paints contained copper and zinc as the biocidal additives and would be released through dissolution of the painted surface or leaching from an insoluble paint matrix. A major advancement in AF technology was the introduction of self-polishing paints where organotin (OT; typically tributyltin (TBT)) biocides that were incorporated into the paint polymer would allow for controlled release of the biocide as the polymer surface was hydrolyzed. Environmental concerns regarding the use of TBT as an AF biocide saw a ban on its use and the introduction of new, primarily organic biocides, often used alongside other biocides such as copper oxide. A common feature of these coatings has been the release of a biocide(s) into the environment. While modern coatings now aim to be biocide free, AF biocides continue to be developed and introduced onto the market. This chapter provides a comprehensive overview of our understanding of the potential harm caused by biocides released from AF paints applied to ships.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2020

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Ali, H. R., Arifin, M. M., Sheikh, M. A., Shazili, M. A. N., Bachok, Z. (2013). Occurrence and distribution of antifouling biocide Irgarol-1051 in coastal waters of Peninsular Malaysia. Marine Pollution Bulletin 70, 253257.Google Scholar
Ali, H. R., Arifin, M. M., Sheikh, M. A. et al. (2014). Contamination of Diuron in coastal waters around Malaysian Peninsular. Marine Pollution Bulletin 85, 287291.CrossRefGoogle ScholarPubMed
Alzieu, C. (1991). Environmental problems caused by TBT in France: assessment, regulations, prospects. Marine Environmental Research 32, 717.Google Scholar
Alzieu, C. (2000). Environmental impact of TBT: the French experience. Science of the Total Environment 258, 99102.Google Scholar
Amey, R. L., Waldron, C. (2004). Efficacy and chemistry of BOROCIDETM Ptriphenylboron-pyridine, a non-metal antifouling biocide. In: Proceedings of the International Symposium on Antifouling Paint and Marine Environment. Tokyo: InSAfE, pp. 234243.Google Scholar
Anastasiou, T. I., Chatzinikolaou, E., Mandalakis, M., Arvanitidis, C. (2016). Imposex and organotin compounds in ports of the Mediterranean and the Atlantic: is the story over? Science of the Total Environment 569–570, 13151329.Google Scholar
Antizar-Ladislao, B. (2008). Environmental levels, toxicity and human exposure to tributyltin (TBT)-contaminated marine environment: a review. Environment International 34, 292308.Google Scholar
Arrhenius, Å., Backhaus, T., Hilvarsson, A. et al. (2014). A novel bioassay for evaluating the efficacy of biocides to inhibit settling and early establishment of marine biofilms. Marine Pollution Bulletin 87, 292299.CrossRefGoogle ScholarPubMed
Artifon, V., Castro, Í. B., Fillmann, G. (2016). Spatiotemporal appraisal of TBT contamination and imposex along a tropical bay (Todos os Santos Bay, Brazil). Environmental Science and Pollution Research International 23, 1604716055.Google Scholar
Bao, V. W. W., Koutsaftis, A., Leung, K. M. Y. (2008). Temperature-dependent toxicities of chlorothalonil and copper pyrithione to the marine copepod Tigriopus japonicus and dinoflagellate Pyrocystis lunula. Australasian Journal of Ecotoxicology 14, 4554.Google Scholar
Bao, V. W. W., Leung, K. M. Y., Qiu, J. W., Lam, M. H. W. (2011). Acute toxicities of five commonly used antifouling booster biocides to selected subtropical and cosmopolitan marine species. Marine Pollution Bulletin 62, 11471151.CrossRefGoogle ScholarPubMed
Barreiro, R., Gonzalez, R., Quintela, M., Ruiz, J. (2001). Imposex, organotin bioaccumulation and sterility of female Nassarius reticulatus in polluted areas of NW Spain. Marine Ecology Progress Series 218, 203212.CrossRefGoogle Scholar
Barroso, C. M., Moreira, M. (2002a). Imposex, female sterility and organotin contamination of the prosobranch Nassarius reticulatus from the Portuguese coast. Marine Ecology Progress Series, 230, 127135Google Scholar
Barroso, C. M., Moreira, M. H. (2002b). Spatial and temporal changes of TBT pollution along the Portuguese coast: inefficacy of the EEC Directive 89/677. Marine Pollution Bulletin 44, 480486.Google Scholar
Barroso, C., Reis-Henriques, M., Ferreira, M., Moreira, M. (2002). The effectiveness of some compounds derived from antifouling paints in promoting imposex in Nassarius reticulatus. Journal of the Marine Biological Association of the United Kingdom, 82(2), 249255.Google Scholar
Basheer, C., Tan, K. S., Lee, H. K. (2002). Organotin and Irgarol-1051 contamination in Singapore coastalwaters. Marine Pollution Bulletin 44, 697703.CrossRefGoogle Scholar
Batista, R. M., Castro, Í. B., Fillmann, G. (2016). Imposex and butyltin contamination still evident in Chile after TBT global ban. Science of the Total Environment, 566–567, 446453.CrossRefGoogle ScholarPubMed
Batista-Andrade, J. A., Caldas, S. S., de Oliveira Arias, J. L. et al. (2016). Antifouling booster biocides in coastal waters of Panama: first appraisal in one of the busiest shipping zones. Marine Pollution Bulletin 112, 415419.CrossRefGoogle ScholarPubMed
Biggs, T. W., D’Anna, H. (2012). Rapid increase in copper concentrations in a new marina, San Diego Bay. Marine Pollution Bulletin 64, 627635.Google Scholar
Biselli, S., Bester, K., Huhnerfuss, H., Fent, K. (2000). Concentrations of the antifouling compound Irgarol 1051 and of organotins inwater and sediments of German North and Baltic Sea Marinas. Marine Pollution Bulletin 40, 233243.CrossRefGoogle Scholar
Blunden, S. J., Chapman, A. H. (1982). The environmental degradation of organotin compounds – a review. Environmental Technology Letters 3, 267272.CrossRefGoogle Scholar
Bowman, J. C., Readman, J. W., Zhou, J. L. (2003). Seasonal variability in the concentrations of Irgarol 1051 in Brighton Marina, UK; including the impact of dredging. Marine Pollution Bulletin 46, 444451.Google Scholar
Boxall, A. B. A., Comber, S. D., Conrad, A. U., Howcroft, J., Zaman, N. (2000). Inputs, monitoring and fate modelling of antifouling biocides in UK estuaries. Marine Pollution Bulletin 40, 898905.Google Scholar
Bryan, G. W., Gibbs, P. E., Hummerstone, L. G., Burt, G. R. (1986). The decline of the gastropod Nucella lapillus around south-west England: evidence for the effect of tributyltin from antifouling paints. Journal of the Marine Biological Association of the United Kingdom 66(3), 611640.CrossRefGoogle Scholar
Bureau of Land Management (2005). Diuron Ecological Risk Assessment. All U.S. Government Documents (Utah Regional Depository). Paper 105. http://digitalcommons.usu.edu/govdocs/105Google Scholar
Burton, E. D., Phillips, I. R., Hawker, D. W. (2006). Tributyltin partitioning in sediments: effect of aging. Chemosphere 63, 7381.Google Scholar
Cacciatore, F., Noventa, S., Antonini, C. et al. (2018). Imposex in Nassarius nitidus (Jeffreys, 1867) as a possible investigative tool to monitor butyltin contamination according to the Water Framework Directive: a case study in the Venice Lagoon (Italy). Ecotoxicology and Environmental Safety 148, 10781089.Google Scholar
Cai, Z., Fun, Y., Ma, W. T., Lam, M., Tsui, J. (2006). LC-MS analysis of antifouling agent Irgarol 1051 and its decyclopropylated degradation product in seawater from marinas in Hong Kong). Talanta 70(1), 9196.Google Scholar
Caquet, T., Roucaute, M., Mazzella, N. et al. (2013). Risk assessment of herbicides and booster biocides along estuarine continuums in the Bay of Vilaine area (Brittany, France). Environmental Science and Pollution Research 20, 651666.Google Scholar
CAR (2014). Work Programme for Review of Active Substances in Biocidal Products Pursuant to Council Directive 98/8/EC – Copper Pyrithione (PT 21). Document IIIA 5. Sundbyberg: KEMI, Swedish Chemical Agency.Google Scholar
Carafa, R., Wollgast, J., Canuti, E. et al. (2007). Seasonal variations of selected herbicides and related metabolites in water, sediment, seaweed and clams in the Sacca di Goro coastal lagoon (Northern Adriatic). Chemosphere 69, 16251637.Google Scholar
Carbery, C., Owen, R., Frickers, T., Otero, E., Readman, R. (2006). Contamination of Caribbean coastal waters by the antifouling herbicide Irgarol 1051. Marine Pollution Bulletin 52, 635644.CrossRefGoogle ScholarPubMed
Carteau, D., Vallée-Réhel, K., Linossier, I. et al. (2014). Development of environmentally friendly antifouling paints usingbiodegradable polymer and lower toxic substances. Progress in Organic Coatings 77, 485493.Google Scholar
Cassi, R., Tolosa, I., deMora, S. (2008). A survey of antifoulants in sediments from ports and marinas along the French Mediterranean coast. Marine Pollution Bulletin 56, 19431948.Google Scholar
Castro, I. B., Westphal, E., Fillmann, G. (2011). Tintas anti-incrustantes de terceira geração: novos biocidas no ambiente aquático [Third-generation antifouling paints: new biocides in the aquatic environment]. Quimica Nova 34, 10211031.Google Scholar
CBP (2006). Assessment Report: Dichlofluanid, Product-type 8. Regulation (EU) No. 528/2012 Concerning the Making Available on the Market and Use of Biocidal Products, Inclusion of Active Substances in Annex I to Directive 98/8/EC. www.echa.europa.eu/documents/10162/17158507/consolidated_bpr_en.pdfGoogle Scholar
CBP (2009). Assessment Report: Tolylfluanid Product-type 8. Directive 98/8/EC Concerning the Placing of Biocidal Products on the Market, Inclusion of Active Substances in Annex I. https://circabc.europa.eu/sd/a/5fd06ea4-c46e-48b8-ab80-a5bb7106a807/Tolylfluanid%20assessment%20report_March_25_09.pdfGoogle Scholar
CBP (2011). Assessment Report: Cybutryne (PT21). Directive 98/8/EC Concerning the Placing of Biocidal Products on the Market, Inclusion of Active Substances in Annex I. http://dissemination.echa.europa.eu/Biocides/ActiveSubstances/1281-21/1281-21_Assessment_Report.pdfGoogle Scholar
CBP (2013). Assessment Report: Zineb, Product-type 21. Regulation (EU) No. 528/2012 Concerning the Making Available on the Market and Use of Biocidal Products, Evaluation of Active Substances. http://dissemination.echa.europa.eu/Biocides/ActiveSubstances/1409-21/1409-21_Assessment_Report.pdfGoogle Scholar
CBP (2014a). Assessment Report: 4,5-Dichloro-2-octyl-2H-isothiazol-3-one (DCOIT), Product-type 21. Regulation (EU) No. 528/2012 Concerning the Making Available on the Market and Use of Biocidal Products, Evaluation of Active Substances. https://circabc.europa.eu/sd/a/5d2b12c8-7690-4636-a962-ab277f4b183d/DCOIT%20-%20PT%2021%20(assessment%20report%20as%20finalised%20on%2013.03.2014).pdfGoogle Scholar
CBP (2014b). Assessment Report: Tolylfluanid Product-type 21. Regulation (EU) No. 528/2012 Concerning the Making Available on the Market and Use of Biocidal Products, Evaluation of Active Substances. https://echa.europa.eu/documents/10162/6dc15617-6986-8a61-3a9a-ef782a0d66f1Google Scholar
CBP (2014c). Assessment Report: Tralopyril. Regulation (EU) No. 528/2012 Concerning the Making Available on the Market and Use of Biocidal Products, Evaluation of Active Substances. https://echa.europa.eu/documents/10162/edf62568-dafb-73a2-51b7-b3118505dab3Google Scholar
CBP (2015a). Assessment Report: Copper Pyrithione, Product-type 21. Regulation (EU) No. 528/2012 Concerning the Making Available on the Market and Use of Biocidal Products, Evaluation of Active Substances. http://dissemination.echa.europa.eu/Biocides/ActiveSubstances/1275-21/1275-21_Assessment_Report.pdfGoogle Scholar
CBP (2015b). Assessment Report: Medetomidine, Product-type 21. Regulation (EU) No. 528/2012 Concerning the Making Available on the Market and Use of Biocidal Products, Evaluation of Active Substances. http://dissemination.echa.europa.eu/Biocides/ActiveSubstances/1327-21/1327-21_Assessment_Report.pdfGoogle Scholar
Champ, M. A. (2000). A review of organotin regulatory strategies, pending actions, related costs and benefits. Science of the Total Environment 258, 2171.Google Scholar
Clark, E. A., Sterritt, R. M., Lester, J. N. (1988). The fate of tributyltin in the aquatic environment. Environmental Science & Technology 22, 600604.Google Scholar
Commendatore, M. G., Franco, M. A., Gomes Costa, P. et al. (2015). Butyltins, polyaromatic hydrocarbons, organochlorine pesticides, and polychlorinated biphenyls in sediments and bivalve mollusks in a mid-latitude environment from the Patagonian coastal zone. Environmental Toxicology and Chemistry 34, 27502763.Google Scholar
Connelly, D. P., Readman, J. W., Knap, A. H., Davies, J. (2001). Contamination of the coastal waters of Bermuda by organotins and the triazine herbicide Irgarol 1051. Marine Pollution Bulletin 42, 409414.CrossRefGoogle ScholarPubMed
Daehne, D., Fürle, C., Thomsen, A., Watermann, B., Feibicke, M. (2017). Antifouling Biocides in German marinas – exposure assessment and calculation of national consumption and emission. Integrated Environmental Assessment and Management 13, 892905.Google Scholar
Dafforn, K. A., Lewis, J. A., Johnston, E. (2011). Antifouling strategies: history and regulation, ecological impacts and mitigation. Marine Pollution Bulletin 62, 453465.Google Scholar
Dahl, B., Blanck, H. (1996). Toxic effects of the antifouling agent Irgarol 1051 on periphyton communities in coastal water microcosms. Marine Pollution Bulletin 32, 342350.Google Scholar
de Mora, S. J., Pelletier, E. (1997). Environmental tributyltin research: past, present, future. Environmental Technology 18, 11691177.Google Scholar
Devilla, R. A., Brown, M. T., Donkin, M., Tarran, G. A., Aiken, J., Readman, J. W. (2005). Impact of antifouling booster biocides on single microalgal species and on a natural marine phytoplankton community. Marine Ecology Progress Series 286, 112.Google Scholar
Di Landa, G., Parrella, L., Avagliano, S. et al. (2009). Assessment of the potential ecological risks posed by antifouling booster biocides to the marine ecosystem of the Gulf of Napoli (Italy). Water, Air and Soil Pollution 200, 305321.Google Scholar
Diniz, L. G. R. (2016). Análise cromatográfica de biocidas anti-incrustantes em amostras de água de mar e tecido de moluscos. PhD thesis, São Carlos, Brazil.Google Scholar
Diniz, L. G. R., Jesus, M. S., Dominguez, L. A. E. et al. (2014). First appraisal of water contamination by antifouling booster biocide of 3rd generation at Itaqui Harbor (São Luiz – Maranhão – Brazil). Journal of the Brazilian Chemical Society 25, 380388.Google Scholar
EC (2003). Regulation (EC) No. 782/2003 of the European Parliament and of the Council of 14 April 2003 on the prohibition of organotin compounds on ships. OJ L 115, 1,11. http://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX:32003R0782Google Scholar
ECHA (2014). CLH Report for Medetomidine: Proposal for Harmonised Classification and Labelling. York: Chemicals Regulation Directorate.Google Scholar
Eguchi, S., Harino, H., Yamamoto, Y. (2009). Assessment of antifouling biocides contaminations in Maizuru Bay, Japan. Archives of Environmental Contamination and Toxicology 58, 684693.Google Scholar
Eklund, B., Elfström, M., Gallego, I., Bengtsson, B. E., Breitholtz, M. (2010). Biological and chemical characterization of harbour sediments from the Stockholm area. Journal of Soils and Sediments 10, 127141.Google Scholar
Erdelez, A., Furdek Turk, M., Štambuk, A., Župan, I., Peharda, M. (2017). Ecological quality status of the Adriatic coastal waters evaluated by the organotin pollution biomonitoring. Marine Pollution Bulletin 123, 313323.Google Scholar
Fent, K. (2006). Worldwide occurrence of organotins from antifouling paints and effects in the aquatic environment. Antifouling Paint Biocides 5, 71100.Google Scholar
Fernández-Alba, A. R., Hernando, M. D., Piedra, L., Chisti, Y. (2002). Toxicity evaluation of single and mixed antifouling biocides measured with acute toxicity bioassays. Analytica Chimica Acta 456, 303312.Google Scholar
Ferreira, M., Blanco, L., Garrido, A., Vieites, J. M., Cabado, A. G. (2013). In vitro approaches to evaluate toxicity induced by organotin compounds tributyltin (TBT), dibutyltin (DBT), and monobutyltin (MBT) in neuroblastoma cells. Journal of Agricultural and Food Chemistry 61, 41954203.Google Scholar
Ferrer, I., Ballesteros, B., Pilar Marco, M., Barceló, D. (1997). Pilot survey for determination of the antifouling agent Irgarol 1051 in enclosed seawater samples by a direct enzyme-linked immunosorbent assay and solid-phase extraction followed by liquid chromatography-diode array detection. Environmental Science and Technology 31, 35303535.Google Scholar
Gadd, G. M. (2000). Microbial interactions with tributyltin compounds: detoxification, accumulation, and environmental fate. Science of the Total Environment 258, 119127.CrossRefGoogle ScholarPubMed
Gervais, J. A., Luukinen, B., Buhl, K., Stone, D. (2008). Capsaicin Technical Fact Sheet. National Pesticide Information Center, Oregon State University. http://npic.orst.edu/factsheets/archive/Capsaicintech.html#modeGoogle Scholar
Gibbs, P. E., Bryan, G. W., Pascoe, P. L., Burt, G. R. (1987). The use of the dog-whelk, Nucella lapillus, as an indicator of tributyltin (TBT) contamination. Journal of the Marine Biological Association of the United Kingdom 67, 507523.Google Scholar
Gibson, C., Wilson, S. (2003). Imposex still evident in eastern Australia 10 years after tributyltin restrictions. Marine Environmental Research 55, 101112.Google Scholar
Gonzalez-Rey, M., Tapie, N., Le Menach, K. et al. (2015). Occurrence of pharmaceutical compounds and pesticides in aquatic systems. Marine Pollution Bulletin 96, 384400.Google Scholar
Government of Western Australia (2009). Antifouling Biocides in Perth Coastal Waters: A Snapshot at Select Areas of Vessel Activity. Water Science Technical Series. Perth: Government of Western Australia.Google Scholar
Guardiola, F. A., Cuesta, A., Meseguer, J., Esteban, A. M. (2012). Risks of using antifouling biocides in aquaculture. International Journal of Molecular Science 13, 15411560.CrossRefGoogle ScholarPubMed
Guitart, C., Sheppard, A., Frickers, T., Price, A. R. G., Readman, J. W. (2007). Negligible risks to corals from antifouling booster biocides and triazine herbicides in coastal waters of the Chagos Archipelago. Marine Pollution Bulletin 54, 226246.Google Scholar
Hall, L. W. Jr., Giddings, J. M., Solomon, K. R., Balcomb, R. (1999). An ecological risk assessment for the use of Irgarol 1051 as an algaecide for antifouling paints. Critical Reviews of Toxicology 29, 367437.Google Scholar
Hamwijk, C., Schouten, A., Foekema, E. M. et al. (2005). Monitoring of the booster biocide dichlofluanid in water and marine sediment of Greek marinas. Chemosphere 60, 13161324.Google Scholar
Harino, H., Langston, W. J. (2009). Degradation of alternative biocides in the aquatic environment. In: Arai, T. Harino, H. Ohji, M., Langston, W. J., eds., Ecotoxicology of Antifouling Biocides. Tokyo: Springer, pp. 397412Google Scholar
Harino, H., Kitano, M., Mori, Y. et al. (2005). Degradation of antifouling booster biocides in water. Journal of the Marine Biological Association of the United Kingdom 85, 3338.Google Scholar
Harino, H., Midorikawa, S. Arai, T. M. et al. (2006a). Concentrations of booster biocides in sediment and clams from Vietnam. Journal of the Marine Biological Association of the United Kingdom 86, 11631170.Google Scholar
Harino, H., Ohji, M. Wattayakorn, G. et al. (2006b). Occurrence of antifouling biocides in sediment and green mussels from Thailand. Archives of Environmental Contamination and Toxicology 51, 400407.Google Scholar
Harino, H, Yamamoto, Y, Eguchi, S. et al. (2007). Concentrations of antifouling biocides in sediment and mussel samples collected from Otsuchi Bay, Japan. Archives of Environmental Contamination and Toxicology 52, 179188.CrossRefGoogle ScholarPubMed
Harino, H., Arai, T. Ohji, M. Ismail, A. B., Miyazaki, N. (2009). Contamination profiles of antifouling biocides in selected coastal regions of Malaysia. Archives of Environmental Contamination and Toxicology 56, 468478.Google Scholar
Harino, H., Eguchi, S., Arai, T. et al. (2010). Antifouling biocides contamination in sediment of coastal waters from Japan. Coastal Marine Science 34, 230235.Google Scholar
Harino, H., Arifin, Z., Rumengan, I. F. M. et al. (2012). Distribution of antifouling biocides and perfluoroalkyl compounds in sediments from selected locations in the Indonesian coastal waters. Archives of Environmental Contamination and Toxicology 63, 1321.Google Scholar
Hernando, M. D., Ejerhoon, M., Fernandez-Alba, A. R., Chisti, Y. (2003). Combined toxicity effects of MTBE and pesticides measured with Vibrio fischeri and Daphnia magna bioassays. Water Research 37, 40914098.Google Scholar
Hoch, M. (2001). Organotin compounds in the environment – an overview. Applied Geochemistry 16, 719743.Google Scholar
Horiguch, T., Shiraishi, H., Shimizu, M., Morita, M. (1997). Imposex in sea snails, caused by organotin (tributyltin and triphenyltin) pollution in Japan: a survey. Applied Organometallic Chemistry 11, 451455.Google Scholar
IFOP (2013). Informe Final: Determinación y evaluación de los componentes presentes en las pinturas anti-incrustantes utilizadas em la acuicultura, sus efectos y la acumulación em sedimentos marinos de la X región de los Lagos. Valparaíso: Instituto de Fomento Pesquero.Google Scholar
IMO (2012). International Convention On the Control of Harmful Anti-Fouling Systems on Ships, 2001. Treaty Series n. 13. London: The Stationery Office.Google Scholar
International (2014). Ficha de Informação de Segurança de Produto Químico (FISP) – Micron Navigator Preto [Material Safety Datasheet (MSDS) – Micron Navigator Black]. www.yachtpaint.com/bra/diy/produtos/anti-incrustantes/micron-navigator.aspxGoogle Scholar
Johansson, P., Eriksson, K. M., Axelsson, L., Blanck, H. (2012). Effects of seven antifouling compounds on photosynthesis and inorganic carbon use in sugar kelp Saccharina latissimi (Linnaeus). Archives of Environmental Contamination and Toxicology 63, 365377.Google Scholar
Kaonga, C. C., Takeda, K., Sakugawa, H. (2016). Concentration and degradation of alternative biocides and an insecticide in surface waters and their major sinks in a semi-enclosed sea, Japan. Chemosphere, 145, 256264.Google Scholar
KEMI (2014). Work Programme for Review of Active Substances in Biocidal Products Pursuant to Council Directive 98/8/EC: Copper Pyrithione (PT 21). Final Competent Authority Report (CAR). Sundbyberg: KEMI, Swedish Chemical Agency.Google Scholar
Kim, N. S., Shim, W. J., Yim, H. U. et al. (2014). Assessment of TBT and organic booster biocide contamination in seawater from coastal areas of South Korea. Marine Pollution Bulletin 78, 201208.Google Scholar
Kim, N. S., Hong, S. H., An, J. G., Shin, K. H., Shim, W. J. (2015). Distribution of butyltins and alternative antifouling biocides in sediments from shipping and shipbuilding areas in South Korea. Marine Pollution Bulletin 95, 484490.Google Scholar
Kim, N. S., Hong, S. H., Shin, K. H., Shim, W. J. (2017). Imposex in Reishia clavigera as an indicator to assess recovery of TBT pollution after a total ban in South Korea. Archives of Environmental Contamination and Toxicology 73, 301309.Google Scholar
Knutson, S.Downs, C. A.Richmond, R. H. (2012). Concentrations of Irgarol in selected marinas of Oahu, Hawaii and effects on settlement of coral larval. Ecotoxicology 21(1), 18.Google Scholar
Konstantinou, I. K., Albanis, T. A. (2004). Worldwide occurrence and effects of antifouling paint booster biocides in the aquatic environment: a review. Environment International 30, 235-248.Google Scholar
Koutsaftis, A., Aoyama, I. (2007). Toxicity of four antifouling biocides and their mixtures on the brine shrimp Artemia salina. Science of the Total Environment 387, 166174.Google Scholar
Kurouani-Harani, H. (2003). Microbial and photolytic degradation of benzothiazoles in water and wastewater. Doctorate thesis. Fakultät III der Technischen Universität Berlin, Berlin, Germany.Google Scholar
Lam, K. H., Cai, Z., Wai, H. Y. et al. (2005). Identification of a new Irgarol-1051 related s-triazine species in coastal waters. Environmental Pollution 136, 221230.Google Scholar
Lam, N. H., Jeong, H.-H., Kang, S.-D. et al. (2017). Organotins and new antifouling biocides in water and sediments from three Korean Special Management Sea Areas following ten years of tributyltin regulation: contamination profiles and risk assessment. Marine Pollution Bulletin 121, 302312.Google Scholar
Lamoree, M. H., Swart, C. P., Van Der Horst, A., Van Hattum, B. (2002). Determination of diuron and the antifouling paint biocide Irgarol 1051 in Dutch marinas and coastal waters. Journal of Chromatography A 970, 183190.Google Scholar
Langston, W. J., Pope, N. D. (1995). Determinants of TBT adsorption and desorption in estuarine sediments. Marine Pollution Bulletin 31, 3243.Google Scholar
Langston, W. J., Pope, N. D., Davey, M. et al. (2015). Recovery from TBT pollution in English Channel environments: a problem solved? Marine Pollution Bulletin 95, 551564.Google Scholar
Laranjeiro, F., Sánchez-Marín, P., Barros, A. et al. (2016). Triphenyltin induces imposex in Nucella lapillus through an aphallic route. Aquatic Toxicology 175, 127131.Google Scholar
Laranjeiro, F., Sánchez-Marín, P., Oliveira, I. B., Galante-Oliveira, S., Barroso, C. M. (2018). Fifteen years of imposex and tributyltin pollution monitoring along the Portuguese coast. Environmental Pollution 232, 411421.Google Scholar
Lee, M., Kim, U., Lee, I., Choi, M., Oh, J. (2015). Assessment of organotin and tin-free antifouling paints contamination in the Korean coastal area. Marine Pollution Bulletin 99, 157165.Google Scholar
Lee, S., Ching, J., Won, H., Lee, D., Lee, Y. (2011). Analysis of antifouling agents after regulation of tributyltin compounds in Korea. Journal of Hazardous Materials 185, 13181325.Google Scholar
Lind, U., Rosenblad, M. A., Frank, L. H. et al. (2010). Octopamine receptors from the barnacle Balanus improvisus are activated by the alpha(2)-drenoceptor agonist medetomidine. Molecular Pharmacology 78, 237248.Google Scholar
Liu, D., Maguire, R. J., Lau, Y. L. et al. (1997). Transformation of the new antifouling compound Irgarol 1051 by Phanerochaete chrysosporium. Water Research 31, 23632369.CrossRefGoogle Scholar
Liu, D., Pacepavicius, G. J., Maguire, R. J. et al. (1999). Mercuric chloride-catalyzed hydrolysis of the new antifouling compound Irgarol 1051. Water Research 33, 155163.Google Scholar
Liu, S., Zhou, J., Ma, X. et al. (2016). Ecotoxicity and preliminary risk assessment of nonivamide as a promising marine antifoulant. Journal of Chemistry 2016, 14.Google Scholar
Liu, W., Zhao, J., Liu, Y. et al. (2015). Biocides in the Yangtze River of China: spatiotemporal distribution, mass load and risk assessment. Environmental Pollution 200, 5363.CrossRefGoogle ScholarPubMed
Lopes-Dos-Santos, R. M. A., Galante-Oliveira, S., Lopes, E., Almeida, C., Barroso, C. M. (2014). Assessment of imposex and butyltin concentrations in Gemophos viverratus (Kiener, 1834), from São Vicente, Republic of Cabo Verde (Africa). Environmental Science and Pollution Research 21, 1067110677.Google Scholar
Mackie, D. S., van den Berg, C. M. G., Readman, J. W. (2004). Determination of pyrithione in natural waters by cathodic stripping voltammetry. Analytica Chimica Acta 511, 4753.CrossRefGoogle Scholar
Marcheselli, M., Rustichelli, C., Mauri, M. (2010). Novel anti-fouling agent zinc pyrithione: determination, acute toxicity and bioaccumulation in marine mussels (Mytilus galloprovincialis). Environmental Toxicology Chemistry 29, 25832592.Google Scholar
Martínez, K., Barceló, D. (2001). Determination of antifouling pesticides and their degradation products in marine sediments by means of ultrasonic extraction and HPLC–APCI–MS. Fresenius Journal of Analytical Chemistry 370, 940945.Google Scholar
Martínez, K., Ferrer, I., Barceló, D. (2000). Part-per-trillion level determination of antifouling pesticides and their byproducts in seawater samples by off-line solid-phase extraction followed by high-performance liquid chromatography–atmospheric pressure chemical ionization mass spectrometry. Journal of Chromatography A 879, 2737.Google Scholar
Martins, S. E., Lillicrap, A., Fillmann, G., Thomas, K. (2018). Ecotoxicity of organic and organo-metallic antifouling co-biocides and implications for environmental hazard and risk assessments in aquatic ecosystems. Biofouling 34, 3452.CrossRefGoogle ScholarPubMed
Matthai, C., Guise, K., Coad, P., McCready, S., Taylor, S. (2009). Environmental status of sediments in the lower Hawkesbury–Nepean River, New South Wales. Australian Journal of Earth Science 56, 225243.Google Scholar
Matthiessen, P., Waldock, R., Thain, J. E., Waite, M. E., Scropehowe, S. (1995). Changes in periwinkle (Littorina littorea) populations following the ban on TBT-based antifoulings on small boats in the United Kingdom. Ecotoxicology and Environmental Safety 30, 180194.Google Scholar
Mattos, Y., Stotz, W. B., Romero, M. S. et al. (2017). Butyltin contamination in Northern Chilean coast: is there a potential risk for consumers? Science of the Total Environment 595, 209217.Google Scholar
Mezcua, M., Hernando, M. D., Piedra, L., Agüera, A., Fernandez-Alba, A. R. (2002). Chromatography-Mass Spectrometry and Toxicity Evaluation of Selected Contaminants in Seawater. Cromatographia 56, 199206.CrossRefGoogle Scholar
Minchin, D., Bauer, B., Oehlmann, J., Schulte-Oehlmann, U., Duggan, C. B. (1997). Biological indicators used to map organotin contamination from a fishing port, Killybegs, Ireland. Marine Pollution Bulletin 34, 235243.Google Scholar
Mochida, K., Ito, K., Harino, H., Kakuno, A., Fujii, K. (2006). Acute toxicity of pyrithione antifouling biocides and joint toxicity with copper to red sea bream (Pagrus major) and toy shrimp (Heptacarpus futilirostris). Environmental Toxicology Chemistry 25, 30583064.Google Scholar
Mochida, K., Onduka, T., Amano, H. et al. (2012). Use of species sensitivity distributions to predict no-effect concentrations of an antifouling biocide, pyridine triphenylborane, for marine organisms. Marine Pollution Bulletin 64, 28072814.Google Scholar
Mohr, S., Berghahn, R., Mailahn, W. et al. (2009). Toxic and accumulative potential of the antifouling biocide and TBT successor Irgarol on freshwater macrophytes: a pond mesocosm study. Environmental Science and Technology 43, 68386843.Google Scholar
Mukherjee, A., Mohan Rao, K. V., Ramesh, U. S. (2009). Predicted concentrations of biocides from antifouling paints in Visakhapatnam Harbour. Journal of Environmental Management 90, S51S59.Google Scholar
Muñoz, I., Bueno, M. J. M., Agüera, A., Fernández-Alba, A. R. (2010). Environmental and human health risk assessment of organic micro-pollutants occurring in a Spanish marine fish farm. Environmental Pollution 158, 18091816.Google Scholar
Nakajima, K., Yasuda, T. (1990). High-performance liquid chromatographic determination of zinc pyrithione in antidandruff preparations based on copper chelate formation. Journal of Chromatography 502, 379384.Google Scholar
New Zealand Environmental Protection Authority (2012). Antifouling Paints Reassessment – Preliminary Risk Assessment. Environmental Protection Authority, Te Mana Rauht Taiao. Wellington: New Zealand Government.Google Scholar
New Zealand Environmental Protection Authority (2013). Decision on the Application for Reassessment of Antifouling Paints (APP201051). Application for the Reassessment of a Group of Hazardous Substances, under Section 63 of the Hazardous Substances and New Organisms Act 1996. Wellington: New Zealand Government.Google Scholar
Nicolaus, E. E. M., Barry, J. (2015). Imposex in the dogwhelk (Nucella lapillus): 22-year monitoring around England and Wales. Environmental Monitoring and Assessment 187, 736.Google Scholar
NIVA (2012). Screening of Selected Alkylphenolic Compounds, Biocides, Rodenticides and Current Use Pesticides. Statlig Program for Forurensningsovervåking, Rapportnr. 1116/2012. Oslo: Klima og Forurensnings Direktoratet.Google Scholar
Nödler, K., Voutsa, D., Licha, T. (2014). Polar organic micropollutants in the coastal environment of different marine systems. Marine Pollution Bulletin 85, 5059.Google Scholar
Oehlmann, J., Markert, B., Stroben, E. et al. (1996). Tributyltin biomonitoring using prosobranchs as sentinel organisms. Fresenius Journal of Analytical Chemistry 354, 540545.Google Scholar
Okamura, H. (2002). Photodegradation of the antifouling compounds Irgarol 1051 and Diuron released from a commercial antifouling paint. Chemosphere 48, 4350.Google Scholar
Okamura, H., Mieno, H. (2006). Antifouling paint biocides. In: Konstantinou, I., ed., The Handbook of Environmental Chemistry. Heidelberg: Springer, pp. 201212.Google Scholar
Okamura, H., Sugiyama, Y. (2004). Photosensitized degradation of Irgarol 1051 in water. Chemosphere 57, 739743.Google Scholar
Okamura, H., Aoyama, I., Liu, D. et al. (2000). Fate and ecotoxicity of the new antifouling compound Irgarol in the aquatic environment. Water Research 34, 35233530.Google Scholar
Okamura, H., Watanabe, T., Aoyama, I., Hasobe, M. (2002). Toxicity evaluation of new antifouling compounds using suspension-cultured fish cells. Chemosphere 46, 945951.Google Scholar
Okamura, H., Aoyama, I., Ono, Y., Nishida, T. (2003). Antifouling herbicides in the coastal waters of western Japan. Marine Pollution Bulletin 47, 5967.Google Scholar
Okoro, H. K., Fatoki, O. S., Adekola, F. A., Ximba, B. J., Snyman, R. G. (2016). Spatio-temporal variation of organotin compounds in seawater and sediments from Cape Town harbour, South Africa using gas chromatography with flame photometric detector (GC-FPD). Arabian Journal of Chemistry 9, 95104.Google Scholar
Oliveira, I. B., Richardson, C. A., Sousa, A. C. et al. (2009). Spatial and temporal evolution of imposex in dogwhelk Nucella lapillus (L.) populations from north Wales, UK. Journal of Environmental Monitoring 11, 14621468.Google Scholar
Oliveira, I. B., Beiras, R., Thomas, K. V., Suter, M. J. F., Barroso, C. M. (2014). Acute toxicity of tralopyril, capsaicin and triphenylborane pyridine to marine invertebrates. Ecotoxicology 23, 13361344.Google Scholar
Omae, I. (2003). General aspects of tin-free antifouling paints. Chemical Reviews 103, 34313448.Google Scholar
OPRF (2010). Final Report on the Comprehensive Management against Biofouling to Minimize Risks on Marine Environment. Tokyo: Nippon Foundation.Google Scholar
Paz-Villarraga, C. A., Castro, Í. B., Miloslavich, P., Fillmann, G. (2015). Venezuelan Caribbean Sea under the threat of TBT. Chemosphere 119, 704710.CrossRefGoogle ScholarPubMed
Petracco, M., Camargo, R. M., Berenguel, T. A. et al. (2015). Evaluation of the use of Olivella minuta (Gastropoda, Olividae) and Hastula cinerea (Gastropoda, Terebridae) as TBT sentinels for sandy coastal habitats. Environmental Monitoring and Assessment 187, 440.Google Scholar
Price, A. R. G., Readman, J. W. (2013). Booster biocide antifoulants: is history repeating itself? Late Lessons From Early Warnings: Science, Precaution, Innovation. Part B: Emerging Lessons From Ecosystems. European Environment Agency (EEA). www.eea.europa.eu/publications/late-lessons-2/late-lessons-chapters/late-lessons-ii-chapter-12/viewGoogle Scholar
Readman, J. W., Kwong, L. L. W., Grondin, D. et al. (1993). Coastal water contamination from a triazine herbicide used in antifouling paints. Environmental Science & Technology 27, 19401942.Google Scholar
Reemtsma, T., Fiehn, O., Kalnowski, G., Jekel, M. (1995). Microbial transformations and biological effects of fungicide-derived benzothiazoles determined in industrial wastewater. Environmental Science & Technology 29, 478485.Google Scholar
Rodríguez, Á. S., Ferrera, Z. S., Rodríguez, J. J. S. (2011a). An evaluation of antifouling booster biocides in Gran Canaria coastal waters using SPE-LC MS/MS. International Journal of Environmental Analytical Chemistry 91(12), 11661177.Google Scholar
Rodríguez, Á. S., Ferrera, Z. S., Rodríguez, J. J. S. (2011b). A preliminary assessment of levels of antifouling booster biocides in harbours and marinas of the island of Gran Canaria, using SPE-HPLC. Environmental Chemistry Letters 9, 203208.Google Scholar
Ruiz, J. M., Quintela, M., Barreiro, R. (1998). Ubiquitous imposex and organotin bioaccumulation in gastropods Nucella lapillus from Galicia (NW Spain): a possible effect of nearshore shipping. Marine Ecology Progress Series 164, 237244.Google Scholar
Ruiz, J. M., Barreiro, R., Couceiro, L., Quintela, M. (2008). Decreased TBT pollution and changing bioaccumulation pattern in gastropods imply butyltin desorption from sediments. Chemosphere 73, 12531257.Google Scholar
Ruiz, J. M., Albaina, N., Carro, B., Barreiro, R. (2015). A combined whelk watch suggests repeated TBT desorption pulses. Science of the Total Environment 502, 167171.Google Scholar
Saleh, A., Molaei, S., Fumani, N., Abedi, E. (2016). Antifouling paint booster biocides (Irgarol 1051 and Diuron) in marinas and ports of Bushehr, Persian Gulf. Marine Pollution Bulletin 105, 367372.Google Scholar
Sánchez-Rodriguez, A., Sosa-Ferrera, Z., del Pino, A. S., Santana-Rodriguez, J. J. (2011). Probabilistic risk assessment of common booster biocides in surface waters of the harbours of Gran Canaria (Spain). Marine Pollution Bulletin 62, 985991.Google Scholar
Santillo, D., Johnston, P., Langston, W. J. (2001). Tributyltin (TBT) antifoulants: a tale of ships, snails and imposex, In: Harremoës, P., Gee, D., MacGarvin, M. et al., eds., Late Lessons from Early Warnings: The Precautionary Principle 1896–2000. Copenhagen: European Environmental Agency, pp. 135148.Google Scholar
Sapozhnikova, Y., Wirth, E., Schiff, K., Brown, J., Fulton, M. (2007). Antifouling pesticides in the coastal waters of Southern California. Marine Pollution Bulletin 54, 19621989.Google Scholar
Sapozhnikova, Y., Wirth, E., Schiff, K., Fulton, M. (2013). Antifouling biocides in water and sediments from California marinas. Marine Pollution Bulletin 69, 189194.Google Scholar
Schouten, A., Mol, H., Hamwijk, C. et al. (2005). Critical aspects in the determination of the antifouling compound dichlofluanid and its metabolite DMSA (N,N-dimethyl-N'-phenylsulfamide) in seawater and marine sediments. Chromatographia 62, 511517.Google Scholar
Schulte-Oehlmann, U., Tillmann, M., Markert, B. et al. (2000). Effects of endocrine disruptors on prosobranch snails (Mollusca: Gastropoda) in the laboratory. Part II: triphenyltin as a xeno-androgen. Ecotoxicology 9, 399412.Google Scholar
Seymour, M. D., Bailey, D. L. (1981). Thin-layer chromatography of pyrithiones. Journal of Chromatography 206, 301310.Google Scholar
Sheikh, M. A., Juma, F. S., Staehr, P. et al. (2016). Occurrence and distribution of antifouling biocide Irgarol-1051 in coral reef ecosystems, Zanzibar. Marine Pollution Bulletin 109, 586590.Google Scholar
Smith, B. S. (1971). Sexuality in the American mud snail, Nassarius obsoletus Say. Proceedings of the Malacological Society of London 39, 377378.Google Scholar
Sousa, A., Laranjeiro, F., Takahashi, S., Tanabe, S., Barroso, C. M. (2009). Imposex and organotin prevalence in a European post-legislative scenario: temporal trends from 2003 to 2008. Chemosphere 77, 566573.Google Scholar
Sousa, A. A., Pastorinho, M. R., Takahashi, S., Tanabe, S. (2014). History on organotin compounds, from snails to humans. Environmental Chemistry Letters 12, 117137.Google Scholar
Stewart, C., de Mora, S. J. (1990). A review of the degradation of tri(n-butyl)tin in the marine environment. Environmental Techology 11, 565570.Google Scholar
Stewart, M., Aherns, M., Olsen, G. (2009). Analysis of Chemicals of Emerging Environmental Concern in Auckland’s Aquatic Sediments. Prepared by NIWA for Auckland Regional Council. Auckland Regional Council Technical Report 2009/021. Auckland: Auckland Research Council.Google Scholar
Stroben, E., Oehlmann, J., Fioroni, P. (1992). The morphological expression of imposex in Hinia reticulata (Gastropoda: Buccinidae): a potential indicator of tributultin pollution. Marine Biology 113, 625636.Google Scholar
Takahashi, K. (2009). Release rate of biocides from antifouling paints. In: Arai, T., Harino, H., Ohji, M., Langston, W. J., eds., Ecotoxicology of Antifouling Biocides. Tokyo: Springer, pp. 322.Google Scholar
Thomas, K. V. (1998). Determination of selected antifouling booster biocides by high-performance liquid chromatography–atmospheric pressure chemical ionisation mass spectrometry. Journal of Chromatography A 825(1), 2935.Google Scholar
Thomas, K. V. (2001). The environmental fate and behaviour of antifouling paint booster biocides: a review. Biofouling 17, 7386.Google Scholar
Thomas, K. V. (2009). The use of broad-spectrum organic biocides in marine antifouling paints. In: Hellio, C., Yebra, D, eds., Advances in Marine Antifouling Coatings and Technologies. Cambridge: Woodhead Publishing Limited, pp. 522553.Google Scholar
Thomas, K. V., Brooks, S. (2010). The environmental fate and effects of antifouling paint biocides. Biofouling 26(1), 7388.Google Scholar
Thomas, K. V., Langford, K. (2009a). Monitoring of alternative biocides: Europe and USA. In: Arai, T., Harino, H., Ohji, M., Langston, W. J., eds., Ecotoxicology of Antifouling Biocides. Tokyo: Springer, pp. 331344.Google Scholar
Thomas, K. V., Langford, K. (2009b). The analysis of antifouling paint biocides in water, sediment and biota. In: Arai, T., Harino, H., Ohji, M., Langston, W. J., eds., Ecotoxicology of Antifouling Biocides. Tokyo: Springer, pp. 311327.Google Scholar
Thomas, K. V., Blake, S., Waldock, M. (2000). Antifouling paint booster biocide contamination in UK marine sediments. Marine Pollution Bulletin 40, 739745.Google Scholar
Thomas, K. V., Fileman, T. W., Readman, J. W., Waldock, M. J. (2001). Antifouling paint booster biocides in the UK coastal environment and potential risks of biological effects. Marine Pollution Bulletin 42, 677688.Google Scholar
Thomas, K. V., McHugh, M., Waldock, M. (2002). Antifouling paint booster biocides in UK coastal waters: inputs, occurrence and environmental fate. Science of the Total Environment 293, 117127.Google Scholar
Thomas, K. V., McHugh, M., Hilton, M., Waldock, M. (2003). Increased persistence of antifouling paint biocides when associated with paint particles. Environmental Pollution 123, 153161.Google Scholar
Titley-O’Neal, C. P., Munkittrick, K. R., Macdonald, B. A. (2011). The effects of organotin on female gastropods. Journal of Environmental Monitoring 13, 23602388.Google Scholar
Tolosa, I., Readman, J. W., Blaevoet, A. et al. (1996). Contamination of Mediterranean (CSte d’Azur) coastal waters by organotins and Irgarol 1051 used in antifouling paints. Marine Pollution Bulletin 32, 335341.Google Scholar
Tornero, V., Hanke, G. (2016). Chemical contaminants entering the marine environment from sea-based sources: a review with a focus on European seas. Marine Pollution Bulletin 112, 1738.Google Scholar
Tsunemasa, N. (2013). Impact of antifouling biocide on marine environment. In: Özhan, E., ed., Proceedings of the Global Congress on ICM: Lessons Learned to Address New Challenges. MEDCOAST 2013 Joint Conference, Marmaris, pp. 10651075.Google Scholar
Tsunemasa, N., Yamazaki, H. (2014). Concentration of antifouling biocides and metals in sediment core samples in the northern part of Hiroshima Bay. International Journal of Molecular Science 15, 999110004.Google Scholar
Turner, A., Glegg, G. (2014). TBT-based antifouling paints remain on sale. Marine Pollution Bulletin 88, 398400.Google Scholar
Turquet, J., Quiniou, F., Delesmon, R., Durand, G. (2010). ERICOR Evaluation du risque « pesticides » pour les récifs coralliens de La Réunion [ERICOR Risk Assessment of pesticides to coral reefs de La Reúnion]. Republique Française. www.programmepesticides.fr/content/download/4531/41722/version/2/file/Turquet_synthese.pdfGoogle Scholar
US EPA (2000–2018). [ECOTOX] ECOTOXicology Knowlegdebase. https://cfpub.epa.gov/ecotoxGoogle Scholar
US EPA (2003). Reregistration Eligibility Decision (RE) for Ziram. Washington, DC: Environmental Protection Agency.Google Scholar
US EPA (2005a). Reregistration Eligibility Decision (RE) for Mancozeb. Washington, DC: Environmental Protection Agency.Google Scholar
US EPA (2005b). Reregistration Eligibility Decision (RE) for Maneb. Washington, DC: Environmental Protection Agency.Google Scholar
US EPA (2006). Reregistration Eligibility Decision for 2-(Thiocyanomethylthio)- benzothiazole (TCMTB). Washington, DC: Environmental Protection Agency.Google Scholar
US EPA (2007). Reregistration Eligibility Decision for 2-Octyl-3 (2H)-isothiazolone (OIT). Washington, DC: Environmental Protection Agency.Google Scholar
Van Scoy, A., Tjeerdema, R. (2014). Environmental fate and toxicology of chlorothalonil. Reviews of Environmental Contamination and Toxicology 232, 89105.Google Scholar
van Wezel, A. P., van Vlaardingen, P. (2004). Environmental risk limits for antifouling substances. Aquatic Toxicology 66, 427444.Google Scholar
Walker, W., Cripe, C., Pritchard, P., Bourquin, A. (1988). Biological and abiotic degradation of xenobiotic compounds in in vitro estuarine water and sediment/water systems. Chemosphere 17, 22552270.Google Scholar
Wang, J., Shi, T., Yang, X., Han, W., Zhou, Y. (2014). Environmental risk assessment on capsaicin used as active substance for antifouling system on ships. Chemosphere 104, 8590.Google Scholar
Wells, F. E., Keesing, J. K., Brearley, A. (2017). Recovery of marine Conus (Mollusca: Caenogastropoda) from imposex at Rottnest Island, Western Australia, over a quarter of a century. Marine Pollution Bulletin 123, 182187.Google Scholar
Wendt, I., Arrhenius, Å., Backhaus, T. et al. (2013). Effects of five antifouling biocides on settlement and growth of zoospores from the marine macroalga Ulva lactuta L. Bulletin of Environmental Contamination and Toxicology 91, 426432.Google Scholar
Wendt, I., Backhaus, T., Blanck, H., Arrhenius, Å. (2016). The toxicity of the three antifouling biocides DCOIT, TPBP and medetomidine to the marine pelagic copepod Acartia tonsa. Ecotoxicology 25, 871879.Google Scholar
WFD (2012). Proposed EQS for Water Framework Directive Annex VIII Substances: Chlorothalonil (For Consultation). London: Water Framework Directive – United Kingdom Technical Advisory Group (WFD-UKTAG).Google Scholar
WHO (1990). Environmental health criteria for tributyltin compounds. Environmental Health Criteria. World Health Organization. www.inchem.org/documents/ehc/ehc/ehc116.htmGoogle Scholar
Willis, K. J., Woods, C. M. C. (2011). Managing invasive Styela clava populations: inhibiting larval recruitment with medetomidine. Aquatics Invasions 6, 511514.Google Scholar
Woldegiorgis, A., Remberger, M., Kaj, L. et al. (2007). Results from the Swedish Screening Programme 2006 Subreport 3: Zinc Pyrithione and Irgarol 1051. www.ivl.se/webdav/files/Rapporter/B1764.pdfGoogle Scholar
Xu, H., Lu, A., Yu, H., Sun, J., Shen, M. (2013). Distribution of Diuron in coastal seawater and sediments from West Sea area of Zhoushan Island. Open Journal of Marine Science 3, 140147.Google Scholar
Xu, Q., Barrios, C. A., Cutright, T., Newby, B. Z. (2005). Evaluation of toxicity of capsaicin and zosteric acid and their potential application as antifoulants. Environmental Toxicology 20, 467474.Google Scholar
Xu, S. (2000). Environmental Fate of Ethylenethiourea. Sacramento, CA: California Department of Pesticide Regulation.Google Scholar
Yamada, H. (2007). Behaviour, occurrence, and aquatic toxicity of new antifouling biocides and preliminary assessment of risk to aquatic ecosystems. Bulletin of Fisheries Research Agencie 21, 3135.Google Scholar
Zhang, A. Q., Leung, K. M. Y., Kwok, K. W. H., Bao, V. W. W., Lam, M. H. W. (2008). Toxicities of antifouling biocide Irgarol 1051 and its major degraded product to marine primary producers. Marine Pollution Bulletin 57, 575586.Google Scholar
Zhou, J., Yang, C., Wang, J. et al. (2013). Toxic effects of environment-friendly antifouling nonivamide on Phaeodactylum tricornutum. Environmental Toxicology Chemistry 32, 802809.Google Scholar
Zhou, X. J., Okamura, H., Nagata, S. (2006). Remarkable synergistic effects in antifouling chemicals against Vibrio fischeri by bioluminescent assay. Journal of Health Science 52, 243251.Google Scholar
Zhou, X. J., Okamura, H., Nagata, S. (2007). Abiotic degradation of triphenylborane pyridine (TPBP) antifouling agent in water. Chemosphere 67, 19041910.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×