Statlig program for forurensningsovervåking Rapportnr. 1067/2010

TA Environmental screening of selected ”new” 2625 brominated flame retardants and selected 2010 polyfluorinated compounds 2009

Utført av DNV i samarbeid med NGI, Universitetet i Umeå, Universitetet i Örebro og Molab AS

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Foreword

Det Norske Veritas AS (DNV), the Norwegian Geotechnical Institute (NGI), University of Umeå, University of Örebro and Molab AS have on behalf of the Climate and Pollution Agency carried out the screening survey of "new" brominated flame retardants and selected polyfluorinated compounds in 2009.

In this screening project several samples in different environmental matrices have been taken in different parts of Norway. This includes sediment from receiving waters and waste water treatment plants, water from waste disposal sites and waste water treatment plants, air samples, biota samples, soil samples and seepage water from fire fighting training grounds. The results from all samples are presented and discussed in this report.

Especially a thanks to University of Umeå for a tremendous job with all work related to development of analytical methods and analysis of the “new” brominated flame retardants and also to the University of Örebro for analysing the polyfluorinated compounds.

In general a special thanks to all which have contributed in this project and especially to:

Det Norske Veritas AS (DNV) Tormod Glette, Amund Ulfsnes, Anders Bergslien, Christian Volan, Marte Braathen for field work and logistic, Sam Arne Nøland for verification and control and Gjermund Gravir for GIS related work. Also thanks to Jens Laugesen for support and discussions.

Norwegian Geotechnical Institute (NGI) Hans Peter Arp for field work and reporting and Arne Pettersen for logistic.

University of Umeå Jenny Rattfelt Nyholm, Roman Grabic and Patrik Andersson for analytical work and reporting.

University of Örebro Anna Kärrman and Kristin Elgh-Dalgren for analytical work and reporting.

Molab AS Marco Skibnes Venzi for performing the air sampling and reporting.

Acknowledgements Thank you to Bård Nordbø at the Climate and Pollution Agency for good co operation and clear communication.

Also thanks to all representatives at the different plants included in this screening. Without good will and co operation from all of them this kind of project would be difficult to execute.

Høvik, 12 April 2010-04-13

Thomas Møskeland, Project leader (DNV)

1 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Content

1. Summary...... 3

2. Sammendrag...... 5

3. Introduction...... 8 3.1 Flame retardants...... 8 3.2 Brominated Flame Retardants...... 9 3.2.1 Background...... 9 3.2.2 Production...... 10 3.2.3 Environmental release, transport and persistence ...... 10 3.2.4 Toxicity and ecotoxicology...... 11 3.2.5 Regulations...... 11 3.2.6 Current levels of BFRs in the Norwegian environment and other parts of the world12 3.2.7 Shift in focus towards emerging “new” BFRs...... 13 3.3 Polyfluorinated compounds ...... 14 3.3.1 Background...... 14 3.3.2 Production...... 15 3.3.3 Environmental Release, Transport and Persistence ...... 16 3.3.4 Toxicology and ecotoxicology...... 17 3.3.5 Regulations...... 18 3.3.6 Environmental levels (published data)...... 19 3.3.7 Current levels in the Norwegian environment ...... 19 3.4 Background and purpose of the study...... 21

4. Description of substances included in the screening ...... 22 4.1 Newly prioritized brominated flame retardants ...... 22 4.2 Polyfluorinated compounds ...... 23

5. Material and methods...... 24 5.1 Description of sampling sites...... 24 5.1.1 Drammen area...... 25 5.1.2 Hokksund area...... 26 5.1.3 Lillehammer area ...... 27 5.1.4 Tromsø area...... 27 5.1.5 Bergen area ...... 28 5.1.6 Haugesund area...... 28 5.2 Sampling and sample treatment...... 29 5.2.1 Drammen area...... 30 5.2.2 Hokksund area...... 35 5.2.3 Lillehammer area ...... 37 5.2.4 Tromsø area...... 39 5.2.5 Haugesund area...... 41 5.2.6 Bergen area ...... 43 5.3 Chemical analysis ...... 47 5.3.1 Brominated flame retardants...... 47 5.3.2 Polyflourinated compounds ...... 50

1 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

6. Results and discussion ...... 55 6.1 Brominated flame retardants...... 57 6.2 Polyfluorinated compounds ...... 73

7. Conclusions ...... 88 7.1 Brominated flame retardants (BFRs)...... 88 7.2 Perfluorinated compounds (PFCs)...... 89

8. References ...... 91

9. Appendix I – Analytical results ...... 109

10. Appendix II – PNEC values for BFRs...... 122

11. Appendix III – Description of “new” BFRs included in the screening123

12. Appendix IV – Description of PFCs included in the screening ...... 150

2 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

1. Summary

Det Norske Veritas AS (DNV), the Norwegian Geotechnical Institute (NGI), University of Umeå, University of Örebro and Molab AS have on behalf of the Climate and Pollution Agency carried out the screening survey of "new" brominated flame retardants and selected polyfluorinated compounds in 2009. In this study, decaBDE, 14 “new” priority brominated flame retardants (BFR), selected polyfluorinated compounds (with focus on PFSAs (PFBS, PFHxS, PFOS, PFDS and 6:2 FTS) were chosen for screening in various samples throughout Norway.

Results A summary of the results are presented in the two tables below.

Overview of results for the investigated BFRs. +: Detected, O: detected in single replicate and/or very close to detection limit -: not detected Compound Sediment Sediment Sludge waste Waste Seepage Biological Air Air receiving waste water facility water water material outdoor indoor water disposal PBT + o - o + - - - PBEB + - - - + - - - HBB + o + + + o + - BTBPE + o + + + - - - DBDPE + - + + + o - - DPTE ------TBPA ------TBP o - - - - + - - ATE ------TBBPAAE + + - + + - - - BTBPI - - - - + - - - EHTBB ------TBBPA-DBPE - - - o + o - - BEHTBP ------Overview of results for the investigated PFCs. +: detected, O: detected in single replicate and/or very close to detection limit, -: not detected. Compound Soil Sediment Water Blue mussel Crab Fish liver 6:2 FTS + + + NQ NQ NQ PFBS + o + - - - PFHxS + + + - + + PFOS + + + - + + PFDS + + - - + + PFPeA + + + - - - PFHxA + + + - - o PFHpA + + + - - - PFOA + + + - o o PFNA o o + - o + PFDA NQ NQ + - - + PFUnDA + + + - + + PFDoDA + + - - o + PFTrDA + + - + + + PFTeDA - - - - - o

3 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Risk assessment of results In order to assess whether the investigated compounds poses an environmental concern or not some general criteria were used:

(i) If the compound was not detected or only detected in samples not taken in the receiving environment it is assessed to be of no or little environmental concern. Included in this category are for example water, sediment and sludge from waste water facilities and waste disposal sites.

(ii) If the compound was detected in receiving environment it is assessed as being of moderate environmental concern. This is nuanced based on comparison with limits for negative effects such as predicted no effect concentrations (PNEC).

(iii) If the compound was identified in biological material it is automatically assessed as being of environmental concern. Detection of substances in needles is not considered to represent biological material. It’s assumed that the pollutants are associated to the waxes on the surface of the needles. In this regards needles are considered as passive samplers.

One should be aware that the assessment should be interpreted with care partly because it’s based on simple criteria and partly because it is based on few samples for most of the investigated compounds.

Conclusions Based on this very general risk assessment the investigated compounds are classified as follows:

No or little environmental concern BFRs: Bis(2-ethylhexyl)tetrabromophtalate (TBPH), 2-etylhexyl-2,3,4,5tetrabromobenzoate (EHTBB), 2,4,6-tribromophenylether (ATE), Tetrabromophtalicanhydride (TBPA), 2,3- dibromopropyl-2,4,6-tribromophenyl ether (DPTE)

PFCs: None identified

Moderate environmental concern BFRs : Hexabromobenzene (HBB), Tetrabromobisphenol A bis(2,3-dibromophenylether) (TBBPA-DBPE), ethylene bis(tetrabromophtalimide) (BTBPI), tetrabromobisphenol A dialyllether (TBBPA-AE), pentabromoethylbenzene (PBEB), pentabromotoulene (PBT)

PFCs: Perfluorobutane sulfonate (PFBS)

Environmental concern BFRs: 2,4,6-tribromophenol (TBP), decabromodiphenylethane (DBDPE) and 1,2 bis(2,4,6- tribromophenoxy)ethane (BTBPE)

PFCs: 1H,1H,2H,2H-tetrahydrofluorooctane sulfonate (6:2 FTS), perfluorooctane sulfonate (PFOS), perfluorodecane sulfonate (PFDS), perfluorohexane sulfonate (PFHxS)

4 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

2. Sammendrag

Bakgrunn På vegne av Klima og forurensningsdirektoratet har Det Norske Veritas AS (DNV), Norges Geotekniske Institutt (NGI), Universitetet i Umeå, Universitetet i Örebro og Molab AS gjennomført screeningundersøkelsen av ”nye” bromerte flammehemmere og polyfluorerte forbindelser i 2009.

I denne studien er decabromodiphenyl eter (decaBDE,) 14 ”nye” prioriterte bromerte flammehemmere og utvalgte polyfluorerte organiske forbindelser (med hovedfokus på PFBS, PFHxS, PFOS, PFDS og 6:2 FTS) undersøkt i ulike prøvematriser i hele Norge. Det viktigste målet med denne studien er å utvide dagens kunnskap om disse potensielt skadelige stoffene med hensyn til deres forekomst i miljøet, om de utgjør en miljørisiko og om utviklingen over forekomsten i miljøet over tid (overvåking).

Dagens kunnskap tilsier at decaBDE er den vanligste bromerte flammehemmeren og PFOS de vanligste perfluorerte forbindelsen i norsk miljø, og to av forbindelser som gjennomgår en "utfasing". Overvåkning av disse to forbindelsene er viktig i forhold til om utfasing vil føre til lavere konsentrasjoner i miljøet av disse to stoffene fremover.

Angående de 14 ”nye” prioriterte bromerte flammehemmerene (BFR) og 6:2 FTS, har forekomsten av disse i norsk miljø ikke vært undersøkt tidligere og kun begrensede mengder data eksisterer i litteraturen på tilstedeværelsen av disse forbindelsene i miljøet. Derfor er det viktig å finne ut om de er tilstede, og hvis de er det om de utgjør en risiko for miljøet. De 14 BFR ble valgt ut av på bakgrunn av en tidligere studie i regi av KLIF (2008) og følgende kriterier ble lagt til grunn:

• Produksjon volum (HPV eller LPV) • Bruk av produktet (additiv, reaktive midler eller polymer) • Potensial for langtransport (Long Range Transport - LRT) • Bioakkumuleringspotensiale (BAP) • Persistens (lite nedbrytbar og kan være i miljøet over lang tid) • Miljø nivåer • Miljø transportprosesser

For å gjøre denne undersøkelsen så meningsfylt og så informativ som mulig er et bredt spekter av miljøprøver tatt på forskjellige steder over hele Norge. Prøvene omfatter sedimenter, vann og slam fra avløpsanlegg, deponier og resipienter, i tillegg til biota (fisk, skjell og krabbe) fra resipienter. Jord og sedimenter har blitt prøvetatt fra brannøvingsfelt. I tillegg er det blitt tatt luftprøver, uteluft fra urbane områder og inneluft fra kontor- og butikklokaler.

Resultater En oppsummering av resultatene er presentert tabellene under.

5 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Oversikt over resultatene for 14 “nye” bromerte flammhemmere. +: detektert, O: detektert i et replikat og/eller veldig nær deteksjonsgrensen, -: ikke detektert. Forbindelse Sediment Sediment Slam avløps Sige- Biologisk Uteluft Inneluft resipient avfalsanlegg rense- vann vann materiale anlegg PBT + o - o + - - - PBEB + - - - + - - - HBB + o + + + o + - BTBPE + o + + + - - - DBDPE + - + + + o - - DPTE ------TBPA ------TBP o - - - - + - - ATE ------TBBPAAE + + - + + - - - BTBPI - - - - + - - - EHTBB ------TBBPA-DBPE - - - o + o - - BEHTBP ------

Oversikt over resultatene for polyfluorerte forbindelser. +: detektert, O: detektert i et replikat og/eller veldig nær deteksjonsgrensen, -: ikke detektert. Forbindelse Jord Sediment Vann Blåskjell Krabbe Fiskelever 6:2 FTS + + + NQ NQ NQ PFBS + o + - - - PFHxS + + + - + + PFOS + + + - + + PFDS + + - - + + PFPeA + + + - - - PFHxA + + + - - o PFHpA + + + - - - PFOA + + + - o o PFNA o o + - o + PFDA NQ NQ + - - + PFUnDA + + + - + + PFDoDA + + - - o + PFTrDA + + - + + + PFTeDA - - - - - o

Risikovurdering av resultater For å vurdere om de undersøkte forbindelsene utgjør en miljømessig bekymring eller ikke ble noen generelle kriterier lagt til grunn. Det er ingen til få relevante data på effekt konsentrasjoner (PNEC, NOEC verdier og lignende) for veldig mange av de undersøkte forbindelsene som kan benyttes i den generelle risikovurderingen. Derfor er følgende relativt enkle kriterier brukt:

6 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

(i) Hvis forbindelsen ikke ble detektert i noen prøver eller bare ble detekter i prøver som ikke var tatt i resipientene eller biologisk materiale ble den vurdert å utgjøre ingen til liten miljømessig bekymring. Inkludert i denne kategorien er for eksempel detekterte forbindelser i vann, sediment og slam fra rense- og avfallsanlegg.

(ii) Hvis forbindelsen ble detektert i resipientene er den vurdert å utgjøre en moderat miljømessig bekymring. Denne vurderingen er nyansert basert på tilgjengelige grenseverdier for negative effekter (eksempelvis PNEC verdier).

(iii) Hvis forbindelsen ble identifisert i biologisk materiale, ble den automatisk vurdert å utgjøre en miljømessig bekymring. Barnåler anses ikke å representere biologisk materiale da det er lagt til grunn at forbindelsene er festet til voks på overflaten av nålene, så her er barnåler betraktet som passive prøvetakere.

Det må nevnes at vurderingene bør tolkes med forsiktighet blant annet fordi de er basert på enkle kriterier og dels fordi de er basert på få prøver for de fleste av de undersøkte forbindelsene.

Ingen eller liten miljømessig bekymring BFR: Bis(2-ethylhexyl)tetrabromophtalate (TBPH), 2-etylhexyl-2,3,4,5tetrabromobenzoate (EHTBB), 2,4,6-tribromophenylether (ATE), Tetrabromophtalicanhydride (TBPA), 2,3- dibromopropyl-2,4,6-tribromophenyl ether (DPTE)

PFC: Ingen identifisert

Moderat miljømessig bekymring BFRs : Hexabromobenzene (HBB), Tetrabromobisphenol A bis(2,3-dibromophenylether) (TBBPA-DBPE), ethylene bis(tetrabromophtalimide) (BTBPI), tetrabromobisphenol A dialyllether (TBBPA-AE), pentabromoethylbenzene (PBEB), pentabromotoulene (PBT)

PFC: Perfluorobutane sulfonate (PFBS)

Miljømessig bekymering BFR: 2,4,6-tribromophenol (TBP), decabromodiphenylethane (DBDPE) and 1,2 bis(2,4,6- tribromophenoxy)ethane (BTBPE)

PFC: 1H,1H,2H,2H-tetrahydrofluorooctane sulfonate (6:2 FTS), perfluorooctane sulfonate (PFOS), perfluorodecane sulfonate (PFDS), perfluorohexane sulfonate (PFHxS)

7 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

3. Introduction

Worldwide many millions of synthetic chemicals have been identified in the environment. Many of them are a concern from a human health and environmental perspective, as above certain threshold levels they can exhibit toxic effects to humans and organisms as well as deleterious effects to ecosystems. The Norwegian government, represented by the Climate and Pollution Agency (former known as Norwegian Pollution Control Authority), are working systematically to identify chemicals which pose a risk to humans and the environment. Part of this work involves yearly screening investigations of the distribution of selected emerging pollutants in the environment. This report deals with the screening investigation carried out in 2009, where the distribution of 14 “new” brominated flame retardants(BFRs), 1 legacy BFR, and 5 polyfluorinated compounds (PFCs) where investigated in different environmental compartments throughout Norway. Of special note, two of the chemicals considered in this year’s screening investigation are in the process of being phased out in Norway, namely decabrominated diphenyl ether (decaBDE) and perfluorooctansulfonate (PFOS). Thus, it is of interest to see if their phasing out is being reflected by decreasing concentrations in the environment.

3.1 Flame retardants

During the course of the twentieth century, manufactures began to move away from traditional material such as wood and metal towards new, engineered materials used plastics building materials and furniture. Additionally, in the textile industry, engineered textiles with unique properties replaced natural textiles, primarily based on wool and cotton. Many of these new materials were more flammable than the materials they replaced. At the same time there was an increasing emphasis on fire-safety by regulators due to the obvious safety concerns of household appliances, building materials and clothing catching fire. This caused the rapid development of flame retardants and the flame retardant industry, which made safe the use of many of the new materials as well pre-existing materials that society has come to rely on. The basic tasks of flame retardants are to minimise both the chances of ignition and the rate of combustion, ideally without compromising the materials they are being applied to or posing any health risks themselves.

There are three major categories of flame retardants: inorganic, organohalogenated and organophosphate compounds. Inorganic flame retardants represent the largest fraction of total flame retardants in use in Europe, organohalogenated flame retardants are most used in other parts of the world, such as Asia. Organohalogenated flame retardants are primarily based on bromine and chlorine. Global consumption figures for 2005 (Fink et al., 2005) are presented in Figure 1.

8 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Figure 1 Use of various flame retardants in different regions (Fink et al., 2005). Based on this survey, brominated flame retardants are the second most used in the world, and the third most used in Europe. Brominated and chlorinated are in the group organohalogenated flame retardants, aluminium hydroxide and antimony trioxide are in the group inorganic flame retardants and organosphosphorous are one major group.

3.2 Brominated Flame Retardants

3.2.1 Background Brominated Flame Retardants (BFRs) comprise a diverse variety of organic compounds, which as their name suggests, are linked by the fact that they contain bromine and are considered commercial flame retardants. Currently, several different types of BFRs exist on the market. Three common molecular-structure categories of BFRs generally assigned are: 1) aromatic, 2) cycloaliphatic and 3) aliphatic; however, there is much more variety amongst BFR structures than these three categories imply. BFRs range in polarity from apolar to ionizable, in size from just a few atoms to macro-molecular, and in reactivity from inert to polymerizing (Andersson et al., 2006, KLIF, 2009a). There are two main economic reasons why such a large variety of BFRs exist on the market. The first reason is that often only very specific BFRs are suitable for fireproofing certain materials (i.e. niche usages), in terms of effectively binding to the material and not influencing the material’s commercial properties. The other reason is that many of the initially used BFRs were deemed to have direct impacts on human and environment health (see below) and thus have prompted the use of alternatives.

BFRs are designed to be either reactive or additive. Reactive BFRs are chemically bonded to materials or even incorporated as an occasional monomer into the polymeric structure of a plastic. Additive BFRs are only physically-sorbed to the material, either by being adsorbed to the surface or absorbed into the matrix of the material.

BFRs are effective flame retardants because when they are heated to combustion temperatures they release bromine radicals, which catalytically bind hydrogen and hydroxyl radicals in the combustion gas, forming water that dilutes the combustion gas and preventing

9 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) hydroxyl radicals from participating in further combustive reactions. This is the same principle also of other halogenated flame retardants; however, fluorine and chlorine radicals are generally not as economically effective (in terms of cost and amount of chemical needed) in this reaction as bromine radicals released from BFRs.

3.2.2 Production BFRs were first used commercially in 1965 (Vonderheide et al., 2008). Their production dramatically increased since then due to the continuous discovery of new, specific uses for individual BFRs, and especially upon the ban of polychlorinated biphenyls (PCBs) in the late 1970s - early 1980s as BFRs were a suitable replacement (Vonderheide et al., 2008). Though many different BFRs were being manufactured, the first generation of BFRs to be produced on a massive scale consisted of polybrominated diphenyl ethers (PBDEs), polybrominated biphenyls (PBBs), tetrabromobisphenol A (TBBPA) and hexabromocyclododecane (HBCD). Of these, only TBBPA was used as a reactive, chemically binding BFR, and the rest were left to absorb and adsorb to the materials. The BFRs PBDE, HBCD, TBBPA and PBB are here referred to as “legacy” BFRs. It is worth noting that other BFRs were produced in the 1970s and 1980s too, however, not to the same massive scale as the legacy BFRs.

3.2.3 Environmental release, transport and persistence One of the first reported environmental disasters related to these compounds was when a BFR mixture called Firemaster TM , comprised mainly of hexabrominated-biphenyl (a PBB), was accidentally mixed into animal feed in 1974, causing diverse harmful effects to livestock as well as exposure to humans of contaminated food (WHO, 1994). As the 1990s progressed, when global production of BFRs exceeded 150 000 metric tons/year (de Wit, 2002) the presence of BFRs in the environment became a rapidly increasing concern. Globally, many BFRs were found to be leaking in trace amounts from the products they were applied to. At hot spots, such as landfills of electric waste , elevated and even toxic concentrations of BFRs are commonly found in leachate and nearby water supplies (e.g. Leung et al., 2007).

Not only in hotspots are BFRs a concern. Like many industrially produced organohalogenated compounds (such as PCBs and many pesticides), BFRs can be found in elevated concentrations within biota and environmental samples in remote locations, far from any sources, where they can potentially exhibit toxic effects. This phenomenon of organic chemicals being transported long distances in the environment is referred to as “Long Range Transport” (LRT). In order for a molecule to exhibit LRT, it has to a) be persistent in the environment (i.e. resistant to transformation reactions when exposed to the environment), and b) be able to be transported spontaneously by environmental mechanisms (e.g. wind transport, water currents, etc.).

Whether or not a molecule is transported in the environment depends on if it has the right range of physical-chemical properties that allow it to volatilize, as well as adsorb or absorb to various, mobile environmental media. For instance, small organohalogenated compounds, such as methylene bromide, are quite volatile and thus can spread in the environment as a vapour in the atmosphere. Polar compounds, (e.g. bromophenols) are soluble in water, and thus can readily be transported by water currents. Very large, apolar compounds (e.g. decaBDE) are very insoluble in water and usually exhibit low volatility, and thus such compounds are not transported in water substantially as a dissolved molecule, nor in the air as a vapour. However, such molecules can nevertheless be transported within air and water by sorbing to aerosols and suspended sediments that are transported in air and water, as well as into organisms. If the molecules have a particular high affinity for lipids and proteins in

10 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) biota compared to the surrounding medium, it is common to find elevated amounts of these molecules in biota, despite lack of detection in the surrounding medium (a phenomena referred to as bioaccumulation). Transport in the environment can also be facilitated through the food chain, such as from algae to fish to bird (a process referred to as biomagnifications).

Why BFRs and other organohalogens are environmental persistent is largely due to the strong nature of carbon-halogen bonds. Breaking these bonds requires a relatively substantial amount of energy, compared to what is available under ambient conditions, and thus they occur only very slowly in the environment. Nevertheless they do occur in the environment and at appreciable levels, such as by photolytic and biological degradation pathways (La Guardia et al., 2007). These reactions typically involve the parent BFR being debrominated or oxidized into another brominated compound. As an example decaBDE can degrade into other, less brominated PBDEs (e.g. La Guardia et al., 2007; Schenker et al., 2008b). This example is particularly important, as the presence of decaBDE itself in the environment is generally argued to be benign (as current research indicates it exhibits low toxicity), the formation of ecotoxic daughter products has caused concern regarding the presence of this chemical in the environment.

3.2.4 Toxicity and ecotoxicology There has been a lot of research studying the toxicity of legacy BFRs in humans, wildlife and ecosystems. Here a main summary of some reviews on the subject (e.g. Darnerud, 2003; de Wit, 2002; Hites, 2004) which have collectively set the framework for ongoing research will be given. Initially, most studies were available for PBBs and PBDEs. Key findings for these molecules were that different congeners (both PBBs and PBDEs consist of 209 individual molecules each) seem to be more toxic than others. For instance, pentabromodiphenylethers (pentaBDEs) seem to be substantially more toxic at the low concentrations than decaBDE at much higher concentrations. This is poignant, as many initial PBDE commercial mixtures contained an abundance of pentaBDEs. Well-known toxic effects are related to endocrine disruption, particularly in the thyroid and in immunological hormones, which causes impairments in growth, neuro-development and reproductive development to a broad array of organisms including humans. Toxic effects on kidneys and livers have also been identified and studied, as well as carcinogenesis and harmful mutagenetic and epigenetic influences. Overall, there seems to be a clear cut case for concern of the presence of tetra- to nona- brominated PBDEs in the environment, as at measured environmental concentrations, toxic and ecotoxicological effects are possible (and very likely occur at hot spots). Based on research so far, decaBDE, HBCD and TBBPA appear to be relatively benign at current environmental concentration levels, compared to PBB and less brominated PBDE. However, this is not totally verified yet, and it does not apply to degradation products of these compounds, especially for decaBDE, which as stated earlier, can debrominate to form more toxic daughter products. Toxicological and ecotoxicological studies in this areas have flourished in the past years, though still knowledge gaps exist, and identifying safe exposure levels of these compounds is still a focus of active research. Toxicological studies on the “new BFRs”, such as analysed in this study, is only just emerging.

3.2.5 Regulations As a consequence of BFRs possibly having potential deleterious effects to ecosystems and to humans, many governments along with industry decided to ban or phase out certain BFRs of high risk. The first BFR to experience industrial phase-out were PBBs (prompted by the Firemaster TM incident mentioned earlier), which co-occurred with bans of PCBs during the

11 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) late 1970s and early 1980s (WHO, 1994). Starting at around 2003, a wave of bans in many countries on pentaBDE mix and octaBDE mix (mixtures of BDEs dominated by pentabromodiphenyl ethers and octobromodiphenylethers, respectively) emerged. The European Union banned these chemicals in 2003 (EU, 2003) and in 2009 these were included in the priority list by the United Nation’s Environment Program’s Stockholm Convention on Persistent Organic Pollutants (UNEP, 2009). Since the beginning of the century and until the present day, the BFRs that are by far the most commonly produced in Europe are decaBDE, TBBPA and HBCD. Figures from 2003 show that these chemicals are being produced in 56418, 145113 and 21951 metric tonnes/year (Andersson et al., 2006). Estimated emissions of these BFRs in Norway are presented in Figure 2 (KLIF, 2009b , see http://www.miljostatus.no/Tema/Kjemikalier/Noen-farlige-kjemikalier/Bromerte- flammehemmere/ for Norwegian regulations)

Figure 2 2007 Emission levels of BFRs in Norway

3.2.6 Current levels of BFRs in the Norwegian environment and other parts of the world In recent years, KLIF has done several recent screening of the presence of legacy BFRs in diverse environmental samples in Norway (e.g. KLIF, 2004, 2005a, 2007a, 2008c). A review of previous investigations and studies of levels of PBDE in Norway was recently compiled (KLIF, 2008c), and is summarized here. Regarding air levels (vapour and particles), ΣPBDE can range from 0.3-14 pg/m 3, with the decaBDE being the dominant component. In soils and mosses decaBDE is also the dominant component, with levels ranging up to 4.4 ng/g d.w in soils and 9 ng/g d.w in mosses. Water samples, mainly from the Mjøsa region, were reported to range from 19 to 820 pg/l, again with decaBDE being the most common (note, these levels are relatively high). In fresh and salt water sediments, decaBDE too is the most dominating compound, with levels up to the µg/g d.w range. A sediment core in Oslo harbor (sampled 2006) reported concentrations that penta-mix, octa-mix BDEs and HBCD are decreasing, and that concentrations of decaBDE are increasing (KLIF, 2008b). Correspondingly, decaBDE is also the most dominating PBDE reported in landfill and sewage sludge samples (though lighter weight PBDEs are typically found in leachate). In biota, such as fish and birds, however, this dominance of decaBDE is not reflected, as biota in general appears to be

12 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) contaminated with various levels of diverse BDEs. Although decaBDE is expected to be strongly attached to soil and sediments and therefore was not expected to bio accumulate, investigations have shown that also decaBDE bio accumulates.

There are several reviews of the emission, distribution and environmental occurrence of legacy BFRs throughout the world, though by far the most amounts of data exists for the Arctic, Europe and Japan (de Wit, 2002; Law et al., 2008; Watanabe and Sakai, 2003). More data has appeared in recent years for China, though to our knowledge no review of this data has been performed yet. Compound class specific reviews on environmental levels were done for PBDE (Hites, 2004; Vonderheide et al., 2008), HBCD (Covaci et al., 2006), and TBBPA (Covaci et al., 2009). General observations that can be made from screens carried out in Norway (from the previous KLIF reviews), Europe and Japan is that concentrations of decaBDE and HBCD are increasing in biota and environmental samples, whereas levels of pentaBDE and octaBDE are decreasing (e.g KLIF, 2008c; Law et al., 2008; Minh et al., 2007; Covaci et al., 2006). Overall, trends in regulation and industry production do seem to be reflected with corresponding changes in environmental concentrations, particularly in sediment and biota. Unsurprisingly, levels of all BFRs are higher nearest source zones in all media (air, soil, sediment and biota). There is little doubt that PBDEs and HBCD are capable of rapid LRT and biomagnification (Breivik et al., 2006; Covaci et al., 2006; Hites, 2004). Levels of TBBPA are lower and are not commonly present in environmental samples outside of local source zones (e.g. production facilities, Covaci et al., 2009). The biggest uncertainties and concerns commonly brought up in these investigations are 1) little data is still available for transformation products as well as transformation pathways of decaBDE, TBBPA and HBCD, thus there is not enough evidence yet to conclude that generally benign concentrations in certain environments are not an issue (particularly for decaBDE); 2) more knowledge is needed on identifying specific uptake pathway mechanisms in various organisms (e.g. food chain, ingestion, breathing, lipid partitioning, membrane diffusion, etc) in order to make more realistic risk assessments; and 3) more data is needed on emerging BFRs.

3.2.7 Shift in focus towards emerging “new” BFRs Based on the information above, of the three contemporary mass-produced BDEs, decaBDE is generating the most concern. Increasing regulations are being enforced, mainly because decaBDE is often the most abundant BFR present in the environment and there are concerns of its degradation products. In 2008 Norway banned decaBDE, while Sweden, Canada and parts of the USA proposed strict regulations on what it can be used for (KLIF, 2008a; Vonderheide et al., 2008). Currently, there is some momentum in Norway and other countries to take a tighter regulatory line on TBBPA and HBCD (KLIF, 2009b, see also http://www.miljostatus.no/Tema/Kjemikalier/Noen-farlige-kjemikalier/Bromerte- flammehemmere/ ).

Though the possibility of increasingly stringent regulations on decaBDE, HBCD and TBBPA is being highly contested, especially by the BFR industry who argue these chemicals are relatively benign to the environment ( http://www.besf.com and http://www.ebfrip.org/ ). This trend is invariably putting pressure on industry to find alternatives, such as other BFRs that can be demonstrated to exhibit less potentially harmful environmental consequences. Four methods of approach by the industry for finding alternatives include the development of BFRs that 1) is itself benign and is rapidly degraded in the environment to benign, irrelevant

13 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) fragments (such as Br-salts or other benign Br-fragments), 2) bind additively chemically to materials, like TBBPA or stronger, 3) be so hydrophobic and persistent that they will be even less soluble and volatile than existing BFRs (e.g. stable brominated polymers), 4) some combination of the above. A less sustainable approach, but another one that may be economically favorable in the short term, is to choose a random BFR in which little is known, but currently does not happen to fall onto the radar of scientists and regulators.

Because of regulatory and commercial pressures, studies are now focusing on identifying new BFRs in the environment, including degradation products of legacy BFRs.

3.3 Polyfluorinated compounds

3.3.1 Background Polyfluorinated compounds (PFCs) are a large group of chemicals that consist of a perfluorinated alkyl tail (i.e. an alkyl-chain in which all hydrogens are replaced with fluorines) with an organic functional group at one end. The general formula for these compounds is F(CF 2)xR. Two important subsets of PFCs exist, perfluoroalkyl compounds in which the head group contains no C-H bonds and fluorotelomer (FT) compounds in which the R-group contains an even-numbed alkyl-chain (general formula F(CF 2)x(CH 2-CH 2)yR and F(CF 2)x(CH=CH) yR).

In industry, these molecules are commonly used as building blocks to form fluorinated polymers (i.e. perfluoralkylpolymers), which are PFC monomers linked by a hydrocarbon backbone. Note that fluorinated polymers are different than the more commonly known fluoropolymers, such as polytetrafluoroethylene (i.e. Teflon TM ), though certain PFCs are used to assist in the manufacturing process of fluoropolymers and thus may appear as residues. PFC-generated fluoropolymers are used as additives for a vast array of materials to either lower their surface tension (e.g. in hydraulic fluids, photographic emulsifiers and paints), or as a coating to make materials more stain and water repellent (e.g. in carpets, textiles, adhesives). A unique and important application is specialised aqueous film forming foams (AFFFs) to extinguish oil and jet fuel fires, as PFCs, being powerful surfactants, can facilitate mixture of water and oil, thus facilitating flame extinguishment and eliminating the dispersion of burning oils. As in the case of BFRs, PFCs concentrations in environmental samples have been found to be increasing and are ubiquitous in the environment, and several toxic effects have been identified. This has caused increased regulations, as well as changes in industrial production.

Of the various PFCs, there are two types that have been the most utilized by industry, and have attracted the most attention and concern in recent years. These are the perfluorocarboxylic acids (PFCAs) (general formula F(CF 2)xCOOH) and perfluorosulfonic acids (PFSAs) (F(CF 2)xS(O 3)H). These acids are readily ionized, and also exist in negatively

14 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) charged ions (by the loss of the proton) or a salt, referred to perfluorocarboxylates and perfluorosulfonates, respectively. Of these two types of PFCs, two individual molecules have attracted the most environmental concern: perfluorooctanoic acid (PFOA, F(CF2) 7COOH) and perfluorooctosulfonic acids (PFOS, F(CF 2)8S(O 3)H).

3.3.2 Production PFCs were first produced in 1946, though mass production did not begin until the 1960s and 1970s, when they became more commonly used as additives in AFFFs and other products (Stock et al., 2010). The most abundantly produced chemicals were clearly PFCAs and the precursor to PFSAs, PSFs (perfluorosulfonyl fluorides, general formula F(CF 2)xS(O 3)F). The two individual molecules that were the most abundantly produced were PFOA and POSF (perfluorooctane sulfonyl, F(CF 2)8S(O 3)F). POSF was used to make PFOS and other products, for conversion to PFOS, simply the fluorine on the -S(O 3)F group was cleaved (Paul et al., 2009; Prevedouros et al., 2006). Though PFOA and POSF were the most predominately produced PFCAs, PFSs and PFSAs of other chain lengths were generated in substantial quantities. Production of these molecules peaked around the year 2000, when the company 3M announced that it was phasing out POSF-based products. At this peak, global POSF emissions are estimated at 4500 tonnes / year (Paul et al., 2009), and PFOA based products at over 300 tonnes/year (Prevedouros et al., 2006). By 2001-2, 3M phased out production of POSF products, and global production levels fell. Estimates for POSF in 2005 are circa 1000 tonnes/year, mainly due to ongoing production in Asia (Paul et al., 2009). In western countries, production of butyl PFCA and PFSA (perfluorobutylcarboxylic acid and perfluorobutylsulfonic acid) have increased, as they are currently thought to be less toxic than PFOA, PFOS, and other long-chain PFCAs and PFSAs (Renner, 2006; Stock et al., 2010).

15 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Figure 3 Production volumes (tonnes/year) of various perfluorinated products, including perfluorooctane sulfonyl (POSF, the main precursor to PFOS), perfluorooctanoic acid, perfluorononanoic acid, and fluorotelomer products (Paul et al., 2009; Prevedouros et al., 2006). POSF experienced a massive drop in production in the year 2000 when it was phased out by 3M, though production continues in Asia. PFCA products were not as produced on the same scale as POSF, though fluorotelomers were, which degrade to PFSAs and PFCAs in the environment. Little data is available for production in recent years, but it is widely speculated smaller chain POSF and PFCA production has increased substantially (Renner, 2006; Stock et al., 2010).

3.3.3 Environmental Release, Transport and Persistence PFCs, particularly PFCA and PFSA, have shown to be exhibit both persistence and LRT (long range transport) in the environment, as confirmed by their ubiquitous presence in environmental samples (Stock et al., 2010). However, the mechanisms of how these chemicals are transported in the environment have eluded scientists, as PFCs have completely different chemical properties than many of the previous chemicals to which environmental persistence and LRT models have been developed for. As opposed to being hydrophobic apolar compounds (like BFRs and PCBs), these compounds are ionic, strong acid, surfactants. The pKa (or acid-dissociation constant) of these compounds is estimated to be near 0 for PFCAs and around -3 for PFSA (Campbell et al., 2009; Goss, 2008). As far as their surfactant properties are concerned, these compounds are amongst some of the most powerful surfactants known. The perfluoralkyl tail is one of the most hydrophobic molecular - - fragments possible, similarly the anionic/acid functional groups (CO 2 , SO 3 ) are some of the most hydrophilic functional groups known. As a result, PFCAs and PFSA have a strong affinity for water surfaces, with the hydrophilic head group liking to be in water whereas the rest of the molecule prefers being outside of the water. These molecules are therefore likely to be transported substantially in the environment by water surfaces (e.g. by dispersion on lake surfaces, sorption to clouds and rain droplets), as has been debated and discussed in the recent literature (Arp and Goss, 2009; Goss and Arp, 2009). The uniqueness of PFC’s chemical properties and what they mean for environmental transport remains an active area of research. Key findings in this direction are that PFCA and PFAS are readily water soluble at low concentrations, form aggregates and micelles at higher concentrations (Arp and Goss, 2009; Cheng et al., 2009), and in the environment associate at interfaces (Psillakis et al., 2009), sorb somewhat but not readily to soils-water and sediment-water interfaces (Higgins

16 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) and Luthy, 2006, 2007), and in biota prefer sorption to proteins than to lipids (Stock et al., 2010), and are transferred into biota by bioaccumulation and biomagnification (Stock et al., 2010).

We note here that knowledge of both physical-chemical parameters and environmental transport pathways are only recently starting to be understood. Thus, reference to hypothesis and physical-chemical data for these compounds from a few years ago can look quite different from what is emerging in the literature today; hence, several statements have been made in the previous paragraph that may appear in contrast to earlier KLIF reports (e.g. in KLIF, 2007b).

To deal with the unique environmental transport and partitioning processes, researchers need an additional set of physical-chemical parameters and models to account for the ionic and surfactant nature of these compounds, such as the pKa (the acid-dissociation constant), surface-water sorption coefficients, and the critical micelle and aggregate-formation concentrations. Though, parameters that are also necessary for apolar, neutral compounds (like the BFRs looked at) are also relevant. Some of these relevant physical-chemical parameters for compounds considered in this screening are provided in Chapter 4, though most parameters remain unknown.

Regarding the persistence of PFC, based on the strong nature of the carbon fluorine bond (the strongest bonds possible in organic molecules), these compounds are extremely resilient to environmental transformations. It appears that the most stable of all PFCs in the environment are PFCAs and PFSAs, which may even be regarded as environmental transformational end products. This is said not only because they are extremely stable (to our knowledge, degradation of these compounds in the environment or under ambient conditions has yet to be observed; though, slow, immeasurable degradation is likely occurring), but also because all other PFCs that have been manufactured appear to be transformed to PFCAs and PFSAs in the atmosphere and elsewhere in the environment (Armitage et al., 2009b; Schenker et al., 2008a; Wallington et al., 2006). The two classes of PFCs that are considered the most substantial precursors of PFCAs and PFSAs are fluorotelomer alcohols (FTOHs) and fluorotelomer sulfonaminds (FSAs) and fluorotelomer sulphonic acids (FTS); however, several hundreds of manufactured PFCs are considered to be capable of conversion into PFCAs and PFSAs (Stock et al., 2010). As a result, the presence of PFCAs and PFSAs are not only due to direct emissions of these compounds, but are also due to indirect transformation of many other PFCs. As a result, multimedia models that account for the occurrence of these molecules consider both these direct and indirect sources of PFCAs and PFSAs (e.g. Armitage et al., 2009a; Armitage et al., 2009b; Schenker et al., 2008a; Wania, 2007).

3.3.4 Toxicology and ecotoxicology Most of the toxicological studies of PFCs have focussed specifically on PFOA and PFOS, and are only recently being expanded to other PFCs. Selected findings reported in recent reviews (Lau et al., 2007; Lau et al., 2004) will be summarised here. For the majority of organisms tested, PFOA and PFSA seem to be readily absorbed through oral pathways (ingestion, breathing), and poorly eliminated. The general explanation for this is that they are not metabolized substantially (due to their stability), and tend to accumulate in the liver, blood serum and kidney; likely because they bind to certain proteins, such as β-lipoproteins, albumin and liver fatty acid binding proteins (and not because they accumulate in lipids).

17 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

These compounds have also been identified as peroxisome proliferators, which can lead to a variety of toxicological effects to the liver, including carcinomas. PFOS and PFOA have also been found to inhibit cell growth by inhibiting intercellular communication, which can lead to cancer growth. Chronic toxicity studies have linked PFOA exposure to tumours in rats and other organisms (though a carcinogenic link has yet to be proven for humans) (Rosen et al., 2009). Recently, PFOA and PFOS have also been found to be endocrine disruptors in humans, being able to compete with thyroxin to the human thyroid hormone transport protein transthyretin (Weiss et al., 2009). In general, it appears from evidence so far that longer chain PFOA and PFAS are much more toxic than shorter chained ones PFCA and PFAS, which has prompted a recent shift in industry to favouring more shorter-chain molecules such as PFButS (Renner, 2006). It should be noted here that one of the molecules considered in this investigation is a fluorotelomer sulphonic acid (FTS), namely 1H,1H,2H,2H- tetrahydroperfluorooctane sulfonate (6:2 FTS, aka THPFOS). Little toxicological information exists for these molecules, however, in the case of telomere-carboxylic acids it has been argued that these are more toxic than PFCAs (Phillips et al., 2007).

In the Norwegian regulation of classification, labelling of dangerous chemicals PFOS is classified as posing a danger for development of cancer, may cause damage to foetus, dangerous by inhalation or contact to the skin, poisonous: severe danger to health by long term influence by swallowing, hazardous to children getting breast milk, poisonous to water living organisms and may cause negative long term effects in water.

There also exists a suggestion for classification of PFOA in the EU (after initiative from KLIF): posing a danger for development of cancer, poisonous: severe danger to health by long term influence by inhalation and swallowing, dangerous if inhaled or swallowed and eye irritating.

3.3.5 Regulations The ubiquitous occurrence and toxicity of PFOS and PFOA has led to an increasing number of governmental regulations, ranging from maximum tolerable intake values, to manufacturing restrictions to outright bans. The U.S. EPA provides a webpage that tabulates recent regulatory developments within the U.S. and elsewhere (http://www.epa.gov/oppt/pfoa/index.html ), and Norwegian Climate and Pollution Control Authority keep an updated webpage of their action plan of these chemicals (http://www.klif.no/no/Tema/Kjemikalier/PFOS-og-andre-PFCs ).

Key regulations regarding PFOS, is that United Nation’s Stockholm Convention on persistent organic pollutants classified it on their list of persistent organic pollutants (UNEP, 2009). It was banned outright (manufacture, use, sale) in Canada in 2008 (Canada, 2008). The EU has set maximum allowable human intake values and maximum values allowed in market goods (EFSA, 2008; EU, 2006), and the U.S. has set maximum tolerable concentration in drinking water (0.2 µg/l for PFOS). In 2005, Norway has set a goal to phase out or severely reduce emissions of PFOS by 2010, and has banned them in AFFFs, textiles and preservatives (KLIF, 2005b, 2008e). In Norway a ban of PFOS and PFOS related compounds in textiles, impregnation compounds and fire foam is given by law trough the regulations relating to restrictions on the manufacture, import, export, sale and use of chemicals and other products hazardous to health and the environment (Product Regulations).

18 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Regarding PFOA, the U.S. EPA launched in 2006 a campaign to achieve a 95% reduction in PFOA in 2010 compared to the 2000 baseline (EPA, 2006) as well as regulated the maximum drinking water concentration to 0.4 µg/l (EPA, 2009). The EU has established guidelines for maximum allowable human intake values (EFSA, 2008). Norway has set a goal to severely reduce or eliminate PFOA emissions by 2010, with completely eliminating them by 2020, and is reviewing a ban in certain consumer products (KLIF, 2005b, 2007b, 2008e).

3.3.6 Environmental levels (published data) Studies can be found that have looked at PFCA levels worldwide in diverse environmental matrices, biota (including wildlife) and humans. A review of the various trends can be found by Stock et al. (Stock et al., 2010), with key ranges summarized here. Keep in mind these levels reflect the ubiquitous presence of these compounds in the environment, and are not only confined to hotspots. Individual PFCs are generally found in the air between <0.1 pg/m 3 - 1 ng/m 3, in precipitation between <0.1 ng/l – 100 ng/l, in groundwater near air force bases (where AFFFs are applied) between undectable to 100 mg/l, in surface water between <1 ng/l – 1 µg/l, and in surface sediments <0.1 ng – 100 ng/l. In wildlife, generally PFOS is the most dominant species, with values ranging from <1 – 36 µg/g w.w, opposed to PFOA which is generally << 20 µg/g w.w.

Currently, one of the most intriguing elements of environmental exposure limits is whether reductions of emissions of PFOS are manifested in reduced amounts in arctic wildlife. Some reports states that there is a correlation between emission reductions and wildlife concentrations (Butt et al., 2007; Hart et al., 2009), and others are reporting the opposite trend (Bossi et al., 2005; Dietz et al., 2008; Holmstrom et al., 2005). One modelling study on this concluded this noticeable response to reduction emission in wildlife concentration will take several years to manifest; however, this was unable to be definitely concluded, due to uncertainties about the exact mechanisms of environmental transport and bioaccumulation pathways (Armitage et al., 2009b). Thus, to best understand what influence the decreasing emissions and increasing regulations are having on PFCA levels, further monitoring of PFCA levels in the environment and more screening studies, such as this one, are needed.

3.3.7 Current levels in the Norwegian environment In Norway, KLIF has commissioned several screens of perfluorinated compounds, primarily PFCAs and PFOS, in diverse environmental and biota samples in 2004 (KLIF, 2005a), 2006 (KLIF, 2007a), and has done literature reviews studies compiling all data and reports on (KLIF, 2008c).

In air samples the most prevalent compounds are PFOS (0.02 – 1.72 pg/m 3), PFOA and PFBA (0 – 4 pg/m 3), and the precursor 8:2 FTOH (8:2 fluorotelomer alcohol) gave high readings in Oslo (10 – 60 pg/m 3). In water samples, PFOS and PFOA levels ranged from 0.2 – 1.41 and 5 – 8 ng/l, respectively. In fresh water sediments PFOS was found in most samples (0.02 - 3.62 ng/g d.w), PFOA is only found sporadically (up to 1 ng/g d.w), as well as other PFCAs and PFSAs. A study in Oslo harbour found PFOS and PFDS most abundant in surface sediments, with levels between 0.1 – 0.7 ng/g d.w. In salt water sediment samples, PFOS was also most abundant, with concentrations ranging from 0.01 – 58.9 ng/g d.w. Other PFCA and PFAS seemed only sporadically present.

19 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

In shell fish most total analysed PFC concentrations were over 1 ng/g dw with PFOS and the PFOS precursor PFOSA (perfluorooctane sulfonamide) being the most dominant. In crustaceans, PFOS ranged from 1 – 10 ng/g d.w, and various PFCAs were found (with the most abundant being PFUnA, perfluoroundecanoic acid, at levels up to 2.5 ng/g d.w). In both fresh water and salt water fish, PFOS was clearly the most abundant, with levels ranging from up to 57 ng/g w.w and various PFCA and PFOSA could often be spotted (with individual concentrations generally << 10 ng/g d.w). In fish, the concentrations were greater in the liver than in whole fish samples. In birds, PFOS was the most dominant by far, with concentrations in plasma and livers being on the order of 100 ng/g w.w. Regarding human samples in Norway, plasma levels of females were measured and were found to be highest in PFOS and PFOA, with median levels around 3.7 and 15.5 ng/g w.w in two subsequent studies for PFOS, and 6.8 and 2.3 ng/g w.w for PFOA.

Regarding potential hotspots in Norway, in landfill effluents PFOS, PFHxsS (perfluorohexanoic sulfonate) and PFOA are the most abundant, ranging 0.6 – 197 ng/l, 0.1 – 143 ng/l and 4.5 – 516 ng/l respectively. In landfill sediments, PFOS and PFHxS were the most commonly measured, with concentrations of 0.22 – 10 ng/g d.w and 0.1 – 1 ng/g d.w respectively. PFOA was found in some samples (< 10 ng/g d.w). In waste water treatment plants, PFOS was quantified in all sewage samples (0.64 – 6.34 ng/l), and PFOA was found in high concentrations in some samples (up to 17 ng/l). Similar findings were for effluent, in that PFOS was generally found in all samples (1 – 18 ng/l), whereas PFOA and PFHexA was found only sporadically but in comparable concentrations when found (up to 22 ng/l). Sludge samples had lower concentrations than sewage samples, with PFOS and PFOA concentrations around 0.5 ng/l with trace amounts of other PFCAs. In storm water samples, PFOS was the only analyte commonly found.

The largest hotspots in Norway are most likely fire training facilities at airports and oil facilities, due to usage of AFFFs (KLIF, 2008d). Soils from Norwegian fire training facilities are contaminated with a variety of PFCs, though usually most with PFOS (with concentrations often > 5 µg/g d.w near the source, the highest was 48 µg/g at a ditch near Rygge airport), about 10 – 10000 times higher than levels found in sewage sludge. PFCs from these facilities were also found to leach vertically into the groundwater, and to spread horizontally, with soils 70 m away from the training facilities being 7-8 times higher than the background concentration. Groundwater and stream water concentrations at these sights were also elevated, and were amongst the highest groundwater concentrations ever reported for these compounds at the time (concentration PFOS 2- 69 ng/l, sum 3 – 117 ng/l), again orders of magnitude higher than levels found in sewage effluent water. These elevated concentrations were also evident in sediment and biota near the facilities. Despite it’s likely that these sites are amongst the most PFC contaminated sites in Norway, these nearby areas are low to moderately polluted according to existing risk-criteria set by KLIF (KLIF, 2007c, 2009c). Sediment levels fit into the “Good” criteria, the most concentrated water levels fit into the “Moderate” category (i.e. chronic in terms of long term exposure), and only some of the soils had concentrations that went above the “normal value” maximum of 20 µg/g d.w. Similarly for organisms analysed (earth worms, molluscs), the concentrations of the organisms were in the lower range of exposure to toxic risk, based on the approaches used in the report (KLIF, 2008d).

As is evident from this brief overview of the presence of perfluorinated compounds in Norway, PFOS is clearly the most abundant, both in terms of frequency of where it is found and concentration, followed by other PFCs, particularly PFHexS and PFOA.

20 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

3.4 Background and purpose of the study

In this study, decaBDE, 14 new priority BFRs, 4 PFSAs (PFBS, PFHxS, PFOS, PFDS) and one FTS (6:2 FTS) were chosen for screening in various samples throughout Norway. The main the goal of this study is to expand current knowledge of these potentially harmful substances with respect to their occurrence in the environment and possible temporal trends. decaBDE and PFOS are the most abundant BFRs and PFCs in the Norwegian environment, respectively, and two of the compounds that are undergoing a “phase-out” transition. Carefully monitoring their presence is important with respect to if stricter regulations will cause lower environmental concentrations of these two substances.

Regarding the 14 “new priority” BFRs and 6:2 FTS, these chemicals have never been selectively screened for in Norway, and only limited data on the presence of these compounds is available in the literature. Therefore it’s important to identify if they are present and if they they are a concern. The 14 BFRs were prioritised in an earlier study that assessed which “new” BFRs could be present in Norway at a substantial level, and were selected from a list of known, manufactured BFRs, based on the following criteria:

• Production volume (HPV or LPV) • Product usage (additive, reactive intermediate or polymer) • Long range transport potential (LRT potential) • Bioaccumulation potential (BAP) • Persistence • Environmental levels • Environmental transport processes

In a screening investigation a wide range of environmental compartments in different places throughout Norway is sampled. Many of the sampled locations are considered to be hot spots meaning that the probability of detecting the compounds here is considered to be relatively high. The samples include sediments, water and sludge from waste water facilities, waste disposal sites and receiving waters, as well as biota from receiving waters and land. Additionally, air samples have been taken, including urban air and indoor air. The localities are located in eastern, western and northern parts of Norway (more details can be found in Chapter 5).

21 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

4. Description of substances included in the screening 4.1 Newly prioritized brominated flame retardants

A detailed description for the 14 new BFRs are given in Appendix III, by reproducing information given on these compounds in an earlier report (KLIF, 2009).

A summary of the structures looked at is presented in the following figure (note that TBB is also referred to as EHTBB):

22 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

4.2 Polyfluorinated compounds

The PFCs looked in this screening include PFSAs of various chain-length and 6:2 FTS. Some introductory information of these particular compounds was provided in Chapter 3.3 and references therein.

A summary of the structures within the scope of this screening is presented in the figure below. In addition other PFSAs has been included when presenting the results. A more detailed description of the compounds is presented in Appendix IV.

23 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5. Material and methods 5.1 Description of sampling sites

One of the purposes of a screening investigation is to sample several environmental matrices with a wide geographical distribution and at locations where it’s expected to find the compounds investigated (hot spots). A general program was worked out by KLIF stating that samples should cover different environmental matrices from eastern, western and northern parts of Norway. Regarding waste water treatment plants they should have different cleaning technologies (mechanical, chemical and biological) in order to see if this had any consequences for the results. Therefore, samples have been collected from waste water treatment plants, waste disposal sites, recycling plants, receiving water, biota and air. The samples have been taken from the eastern, western and northern part of Norway. An overview of the sampling localities and the matrices are presented in the table below.

Table 1 Summary of samples; main area, sampling location, sample category and sample matrices. Area Station Category Matrix Analyses Drammen Solumstrand RA water treatment plant water BFR Drammen Solumstrand RA water treatment plant sludge BFR Drammen Drammensfjord receiving water sediment BFR Drammen Drammensfjord receiving water blue mussel BFR Drammen Drammensfjord receiving water fish liver BFR Drammen Drammensfjord receiving water crab BFR Drammen Lindum Ressurs og waste disposal site water BFR Gjenvinning Drammen Lindum Ressurs og waste disposal site sediment BFR Gjenvinning Drammen Drammen city city urban air BFR Drammen Elkjøp shop/store indoor air BFR Drammen/Hurum Hurum Energigjenvinning incineration of waste moss BFR Drammen/Hurum Hurum Energigjenvinning incineration of waste pine needle BFR Hokksund Hellik Teigen AS metal recycling water BFR facility Hokksund Hellik Teigen AS metal recycling sediment BFR facility Lillehammer Lillehammer RA water treatment plant water BFR Lillehammer Lillehammer RA water treatment plant sludge BFR Lillehammer Mjøsa receiving water sediment BFR Lillehammer Mjøsa receiving water fish liver BFR Lillehammer Losna receiving water sediment BFR (stream) Tromsø Langnes RA water treatment plant water BFR Tromsø Langnes RA water treatment plant sludge BFR Tromsø Sandnessundet receiving water sediment BFR Tromsø Sandnessundet receiving water blue mussel BFR Tromsø Sandnessundet receiving water fish liver BFR Tromsø Sandnessundet receiving water crab BFR Bergen Puddefjorden receiving water fish liver BFR Haugesund RES-Q fire fighting facility water PFC Haugesund RES-Q fire fighting facility sediment PFC Haugesund Bleivika receiving water sediment PFC Haugesund Bleivika receiving water fish liver PFC Haugesund Bleivika receiving water blue mussel PFC

24 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 1 continues Haugesund Bleivika receiving water crab PFC Bergen Flesland airport Airport/ fire fighting water PFC facility Bergen Flesland airport Airport/ fire fighting soil PFC facility Bergen Langavatnet receiving water sediment PFC Bergen Langavatnet receiving water fish liver PFC Bergen Flesland brygge receiving water blue mussel PFC Bergen Flesland brygge receiving water crab PFC

5.1.1 Drammen area A detailed description of the sampling sites is given below.

Solumstrand RA Solumstrand waste water treatment plant is located in Drammen municipality and receives waste water from large areas of Drammen such as Bragernes, Konnerud, Strømsø, Fjell, Tangen, Nøsted and Solum. Waste water that is not going to Solumstrand is led to Muusøya waste water treatment plant. Solumstrand mainly receives water from households (about 70 %). About 30 % of the waste water that passes trough Solumstrand is mainly from the industry, but the plant also receives waste water from Lindum waste disposal site and a hospital (Buskerud HF). Waste water is treated chemically and the cleaning process can be divided as follows: grid, sand- and fat trap, flocculation and sedimentation. The sludge is dewatered by use of centrifuge.

In 2009 the plant treated a water volume of about 8.95 millions m 3. The plant is dimensioned for about 60 000 PE (population equivalent) and a maximum water volume of 3000 m 3/h. Retention time for the water in the plant is about 4 hours.

Discharge water is lead out in the Drammensfjord outside Solumstrand at about 40 m water depth and 200 m from the shoreline.

Lindum Ressurs & Gjenvinning Lindum Ressurs & Gjenvinning is located in Drammen municipality. The facility receives various kinds of wastes from the Drammen region. The waste that cause run off to seepage water can be categorized as slightly polluted masses, residual waste, waste from sand trap and grid, sand from sand blasting, compost, waste put permanently in bio cell and waste to industry bio cell (degradable waste and sludge). A total of about 170 000 m 3 or 230 000 tonnes waste were delivered to the site in 2009. The seepage water is not aerated nor is it treated by any sedimentation (sedimentation basin), but the seepage water is pumped directly to Solumstrand waste water treatment plant.

Drammensfjorden Drammensfjorden is a typical sill fjord with a very narrow sill against the outer part of Oslofjord in the area of Svelvik. Inside the sill the fjord is formed as a deep basin with water depth over 60 m in most parts, and a maximum depth of approximately 120 m. The bottom water is mostly anoxic, meaning that there is limited water exchange with the outer part of the Oslofjord. The fjord receives waste water from Solumstrand RA (water treatment plant) and Lindum Ressurs & Gjenvinning (waste disposal site).

25 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

The city of Drammen has a population of about 62,000, and is located next to the fjord, in the area where the Drammenselva flows into the fjord. Drammenselva is a relatively large river with an average yearly flow of 250 – 300 m 3/s. This flow amounts to about 97 % of the fresh water going into the fjord. The second main source of fresh water to the fjord is the Lierelva, which contributes to about 2 % of the fresh water going into the fjord. Drammenselva flows through several small communities’ mainly agricultural areas, but also various types of industry, including wood processing. Lierelva mainly flows through agricultural areas.

Drammen city and Elkjøp store BFR is sampled from indoor air in Elkjøp Drammen and outdoor air is sampled from the centre of Drammen city.

Some details of Drammen are given above. Elkjøp is Scandinavia's largest trading company for consumer electronics and electrical appliances, and had a turnover of NOK 17.6 billion in 2007/08. Operating income was NOK 1.028 billion.

The Group has established retail operations in Norway, Sweden, Denmark, Finland, Iceland and the Faroe Islands with the main business being related to huge department stores. All of the 233 department stores, including the store in Drammen, in the Nordic countries are mainly supplied from its own distribution business, with a central warehouse in Jönköping (Sweden). Approximately 6,000 employees are employed in Elkjøp group, which is owned by British DSG International plc., One of Europe's largest retailers in consumer electronics.

Hurum Energigjenvinning Hurum Energigjenvinning KS is a plant for incineration of waste. It is located on Hurum in the vicinity of the Drammen fjord. The plant is based on the Energos technology. Energos plants are built in modules with one or more processing line in parallel to meet customer’s requirements regarding energy production and fuel processing capacity. The energy content of the fuel is converted into electricity and/or heat delivery for local use, e.g. district heating or industrial applications. The plant at Hurum was started in 2001, and treats approximately 40,000 tonnes domestic and industrial waste each year. The annual energy production is approximately 90 Gwh. Identifying emissions from the plant are particles, inorganic gasses (such as CO), acid gasses (such as SOx, HCl and NOx), heavy metals, dioxins and PAH’s.

5.1.2 Hokksund area

Hellik Teigen AS Hellik Teigen AS is a car demolishing site located at Losmoen in Hokksund. It receives and recycles iron and other metals from scrapped cars, household appliances and raw metal materials such as steel, cast iron, stainless steel and copper, aluminium, brass and different types of alloys. Hellik Teigen also receives computer and electronic waste, wood and rubber. According to staff members, the plant primarily receives the cars after the main environmental contaminants have been removed. All water run-off at the site is collected and treated (sand trap, oil skimmer, flotation, venturi scrubber) before it is discharged to the small river Loselva. At low water levels in the river (which is influenced by the tidal amplitude) the discharge pipe is well above the water level. Loselva flows into the upper part of Drammenselva which in turn flows into the Drammen fjord.

26 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5.1.3 Lillehammer area

Lillehammer RA Lillehammer water treatment plant was established in 1977 as a mechanical and chemical plant for treatment of waste water, dimensioned for 50 000 PE. In general, the process steps are as follows: removal by grid and sand trap, primary sedimentation and secondary and final sedimentation tank/basin, sludge treatment with gravitational thickening, drainer and centrifuges. In 1993, the plant was expanded to include the Kaldnes Moving Bed biofilm process for removal of nitrogen. In addition, the capacity of the plant was upgraded to 70 000 3 3 PE. The volume of the basin in the plant is about 11,000 m , Q dim = 1200 m /h and Q max = 1900 m 3/h.

Lillehammer water treatment plant receives water from the municipality of Gausdal, some parts of Øyer and Ringsaker, and the municipality of Lillehammer. Besides treatment of domestic waste water, the plant also receives waste water from the food industry, textile industry, laundries, hospital, treatment plant for wet organic waste and waste disposal sites. The water treatment plant has it own sewage tank, and receives sewage from the above- mentioned municipalities as well as thickened sludge from another water treatment plant. The reported volume of treated water in 2008 was 6.118.969 m 3. The requirement for the plant is removal of 95 % of phosphorus (<0.25 mg/l), 70 % BOF, 75 % KOF, and 70 % nitrogen. Treated water is discharged in Mjøsa (lake) about 200 m from land and at 20 m water depth. Dewatered sludge is transported to an intermunicipal sludge treatment plant where it is treated further.

Mjøsa and Losna Mjøsa is Norway’s largest lake with a surface area of 365 km 2. The lake is about 117 km long and the deepest part is officially measured to 468 m. The largest river flows into Mjøsa in the northern part, near Lillehammer, and is named Gudbrandsdalslågen. Losna is in fact a part of Gudbrandalslågen. Several other rivers flow into Mjøsa, such as Rinna, Vismunda, Stokkeelva, Hunnselva, Leanelva in the west,and Moelva, Brumunda, Svartelva and Starelva in the east. The only river that flows out of Mjøsa is Vorma (in the southern end). Mjøsa is known for its population of large trout and lake herring which are popular among anglers.

5.1.4 Tromsø area

Langnes RA Langnes waste water treatment plant is located in Tromsø in northern Norway. Waste water is cleaned mechanically by four Maskozoll coarse screens with 1 mm openings and then by 2 Hydrotech screens with 0.12 mm openings. Waste water going into the plant comes mainly from households. Treated volume of water was in 2009 about 32 798 74 m 3. Sludge is treated by a thickening tank and polymer is added before it is run trough the coarse screens for dewatering. The plant is dimensioned for about 15 000 PE. There is no certain good numbers of water retention times because the water flows straight trough screens and is led trough 530 m of 630 mm PE pipeline before it is discharged at 20 m water depth in Sandnessundet.

27 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Sandnessundet Sandnessundet is a strait between Tromsøya and Kvaløya in northern Norway. The strait is about 14 km long and is dominated by rather strong tidal currents with a speed of up to 2 – 5 knots. Waste water from Langnes RA is discharged to Sandnessundet.

5.1.5 Bergen area

Flesland airport Flesland airport is located in the western part of Norway near the city of Bergen. The airport was included in this investigation due to the existing fire drill area at the airport. The fire drill area has been used for many years on a regular basis. In a fire drill exercise, various kinds of fire fighting powders and foam are used. The existing fire drill area is connected to an oil separator. Seepage water is drained to a stream which leads to Langavatnet, a lake located just adjacent to the airport.

Langavatnet Langavatnet is located within the Flesland airport area. The lake receives seepage water from Flesland airport. Langavatnet drains to the sea through Fleslandselva which flows into the sea near Flesland brygge

Flesland brygge Flesland brygge is located just west of Flesland airport, and is the area where Fleslandselva flows into the sea.

Puddefjorden Puddefjorden is an approximately 3.5 km large part of Byfjorden in Bergen. It is located close to Bergen city and is surrounded by various kinds of industries. Puddefjorden is a polluted fjord and is included in the plan for remediation of polluted sediments in Bergen. Puddefjorden is located northeast of Flesland airport.

5.1.6 Haugesund area

Res-Q Res-Q is located near Bleivik to the north of Haugesund city. Res-Q is a facility where fire and explosion safety courses and drills are held. It is the only facility in Norway that regularly hosts courses where fire foam with FTS (as a substitution in fire foam) is used. Run-off from the facility is lead to two sedimentation basins before discharge to sea in Bleivika at about 15 m depth. An oil skimmer and centrifugation is used for cleaning the run- off. Sediment from the sediment basins is removed twice a year and handled as hazardous waste. PFCs were earlier found in sediment from the facility (SFT, 2007).

Bleivika Bleivika is the receiving water for the discharges from Res-Q. The area is partly sheltered by some small island, but the discharge point from Res-Q is located in a more exposed area outside the Islands at 15 m depth.

28 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5.2 Sampling and sample treatment

In general the sample strategy presented in Table 2 was followed. More detailed description of the sampling is given in the chapters below.

Table 2 Overview of general sampling strategy. Type of sampling Strategy Waste water Composite sample of inlet and outlet water from 1 week in 3 periods => 6 samples. About 7 litres were sampled in glass bottles wrapped in aluminium foil (BFR) and 1 litre in PE bottles (PFC) Sludge One sample from 1 week in 3 periods => 3 samples. Stored in glass bottles wrapped in aluminium foil (BFR) and PE containers (PFC) Seepage water 3 replica of a composite sample from 1 week => 3 samples. About 7 litres were sampled in glass bottles wrapped in aluminium foil (BFR) and 1 litre in PE bottles (PFC) Sediment- disposal site 3 replica of a composite sample from 1 week => 3 samples. Stored in glass bottles wrapped in aluminium foil (BFR) and PE containers (PFC) Moss and pine needles Last year’s growth from 4 stations in different direction and distance from plant => 4 samples. 1 litre of sample from each station was packed in aluminium foil. Samples of pine needles and stair step moss ( Hylocómium splendens ) Soil 10 samples in increasing distance from the fire drill site => 10 samples. Sample stored in PE container (PFC only) Sediment – receiving water 3 replicas from 3 stations with increasing distance from discharge point => 9 samples. Stored in glass bottles wrapped in aluminium foil (BFR) and PE containers (PFC) Fish liver 20 fish from receiving waters analysed as 4 sub samples of 5 fish => 4 samples. Fish was caught by local fishermen after instruction from DNV. Material worked up at DNV’s Biology laboratory and stored in glass bottles wrapped in aluminium foil (BFR) and PE containers (PFC) Blue mussel Composite sample of 30 shells from 3 stations in receiving water = 3 samples. Shells were sampled by local fishermen. Material worked up at DNV’s Biology laboratory and stored in glass bottles wrapped in aluminium foil (BFR) and PE containers (PFC) Crab 20 crabs from receiving waters analysed as 4 sub samples of 5 crab meat => 4 samples. Crabs where caughtby local fishermen. Material worked up at DNV’s Biology laboratory and stored in glass bottles wrapped in aluminium foil (BFR) and PE containers (PFC) Air 1 week of sampling from 3 different periods => 3 samples

The following general remarks are given regarding sampling:

● Clean gloves and suit were used for each sampling. ● No cosmetics were allowed to be used prior to or during sampling. ● Watches and other electronic equipment were not allowed used during sampling. ● The samples were stored in a cool and dark place in the time between sampling. All samples were frozen when they arrived at DNV office prior to analysis. ● All samples were kept frozen until arrival at the analysis laboratory

29 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5.2.1 Drammen area The sampling locations are presented in Figure 4 and Figure 5.

Figure 4 Overview of sampling locations in the inner Drammensfjord and city of Drammen, Drammen area.

30 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Figure 5 Overview of sampling locations around Hurum Energigjenvinning, Drammen area.

31 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Solumstrand RA Information about sampling at Solumstrand water treatment plant is given in Table 3.

Table 3 Sampling information, Solumstrand waste water treatment plant. Sample Sampling period Description Waste water – period 1 09.07.09 – 15.07.09 Inlet water sampled before treatment Waste water – period 2 16.07.09 – 22.07.09 (before grid). Outlet water sampled in basin Waste water – period 3 23.07.09 – 29.07.09 after all cleaning steps. All samples taken at around 14:00. Sludge – period 1 09.07.09 Sludge taken from container under silo. Sludge – period 2 16.07.09 Sample taken at the start of each water Sludge – period 3 23.07.09 sampling period at around 14:00.

Drammensfjorden All sediment samples were collected with a Van veen grab (surface area 0.025 m 2). Surface sediments (0-2 cm) were sampled. All sampling coordinates were logged by a hand held GPS.

Table 4 Sampling information for sediment localities, Drammensfjorden. All coordinates in UTM32, WGS84. Station North East Depth (m) Sampling period Drsol-01 6620184.00 572002.00 80 1.10.2009 Drsol-02 6620283.00 571662.00 57 1.10.2009 Drsol-03 6620113.00 571399.00 25 1.10.2009

An overview of where the biota samples were collected is presented in Table 5. There were some difficulties in finding relevant sampling locations for blue mussel, so only two out of three planned stations were sampled. The fish caught was cod ( Gadus morhua ) and the crab was edible crab ( Cancer pagurus ). Fish were caught with fish traps, blue mussels were caught by towing a scrape, and crabs were caught by the use of lobster pot. Size data for the specimens caught is presented in Table 6.

Table 5 Sampling information for biota localities, Drammensfjorden. All coordinates in UTM32, WGS84. Station North East Sampling period Description Fish 1) 6619670.00 571716.00 16.06.2009 – 19.06.2009 Approximately 300 m south of 03.07.2009 – 04.07.2009 Solumstrand waste water 15.10.2009 – 16.10.2009 treatment plant. Crab 1) 6619670.00 571716.00 03.07.2009 – 07.07.2009 Approximately 300 m south of 15.10.2009 – 16.10.2009 Solumstrand waste water treatment plant. Blue mussel -1 6619670.00 571716.00 07.07.2009 300 m south of Solumstrand Blue mussel -2 6621105.57 571239.42 15.10.2009 – 16.10.2009 In the area of Nøstodden north of Solumstrand waste water treatment plant 1) Different sampling periods, pooled samples made of individuals from the same sampling period.

32 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 6 Sample data biota, Drammensfjorden Species Average weight/size Standard deviation Cod length 40.1 cm 5.2 cm Cod weight 849 g 394 g Edible crab 2 15.8 cm 0.85 cm Blue mussel 2 6.1 cm 0.9 cm

Drammen city and Elkjøp store During the summer and autumn of 2009 a total of 6 samples of outdoor and indoor air were collected from the two sites in Drammen. To ensure potentially detectable levels of the pollutants, each sampling was set to a continuous 7 days period. Field blanks from each of the sites were also sent to the laboratory for analysis.

After sampling, all the filters were separately wrapped in aluminum foil and stored at -16 °C, except for brief periods during transportation.

There was no absorbent for sampling the gas phase in the actual setup, thus only particle bound compounds were sampled.

For indoor sampling a basic air sampling pump (SKC, Ltd) was used for indoor air sampling, exposing a glass microfiber filter (Whatman GF/F, Ø 25 mm) with a constant flow of 1.0 L/min over the 7 days period. The sampling inlet was mounted at about 1.5 m above floor. Sampling details are presented in Table 7.

Table 7 Sampling of indoor air, schedule and air volume Location Start Stop Total Avg. flow Total air volume (L) time (L/min) (min) Drammen (El-kjøp) 11.08.2009 18.08.2009 10.560 1 10.560 (10:30) (18:30) Drammen (El-kjøp) 14.09.2009 21.09.2009 10.180 1 10.180 (11:30) (13:10) Drammen (El-kjøp) 13.10.2009 16.10.2009 4.150 1 4.150 (14:00) (11:10)

The reported total air volumes are accurate within a relative uncertainty of ±5 %.

The sampling of outdoor air is conducted with a high volume pump with a PM 10 inlet (Sierra instruments, Inc), exposing a glass microfibre filter (Whatman EPM 2000; 20.3 x 25.4 cm). The inlet is approximately 1.5 m above ground level. Sampling details are presented in Table 8.

Table 8 Sampling of outdoor air, schedule and air volume Location Start (time) Stop (time) Total time Avg. flow Total air (min) (L/min) volume (L) Drammen (City 2.910 100 291.000 centre) 11.08.2009 (13:30) 13.08.2009 (14:00) Drammen (City 6.669 100 666.900 centre) 14.09.2009 (11:15) 18.09.2009 (12:00) Drammen (City 1.310 150 196.500 centre) 14.10.2009 (14:10) 15.10.2009 (12:00)

33 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

The high volume pump had an initial sampling flow of 150 L/min. Due to the long sampling time the flow was reduced during the period due to clotting of the filter by particles. The pump flow was measured at the end of sampling, and the reported flow was the linear average of the initial and end flow. Since the flow reduction in reality was not linear, the reported total air volumes are subject to a high degree of uncertainty. There should be considered a relative uncertainty of ±50 % for the reported total air volumes.

Some of the sampling time totals differs from the specified 7 days period. This was due to different technical issues with the sampling equipment.

Hurum Energigjenvinning All samples were taken in different directions from the plant. For details see Table 9.

Table 9 Sampling information for moss and pine needles localities, Hurum Energigjennvinning. All coordinates in UTM32, WGS84. Station North East Sampling period Description 1 6600545.08 586920.66 28.09.2009 North east about 100 m from plant 2 6600556.43 586881.36 28.09.2009 Eastern boundary 3 6600430.57 586737.72 28.09.2009 South west about 100 m from plant 4 6600488.48 586736.95 28.09.2009 North west about 100 m from plant

34 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5.2.2 Hokksund area The sampling location is presented in Figure 6.

Figure 6 Sampling location at Hellik Teigen, Hokkusnd area.

35 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Sampling information for Hellik Teigen AS in Hokksund is presented in Table 10. Sediment samples were taken with a hand held corer. Sediment in the 0-2 cm layer was sampled.

Table 10 Sampling information sediment and seepage water, Hellik Teigen AS. All coordinates in UTM32, WGS84. Station North East Sampling period Description Seepage water 6625014.36 552973.83 8.07.2009 – 15.07.2009 Samples of seepage water were taken at the discharge point for seepage water in Loselva. Sediment 6625014.36 552973.83 8.07.2009 – 15.07.2009 Samples taken with hand held corer about 10 m from discharge point of seepage water.

36 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5.2.3 Lillehammer area An overview of the locations of the sampling stations in the Lillehammer area is presented in Figure 7 below.

Figure 7 Overview of sampling locations, Lillehammer area

37 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Lillehammer RA Information about sampling at Lillehammer water treatment plant is given in Table 11.

Table 11 Sampling information, Lillehammer waste water treatment plant. Sample Sampling period Description Waste water – period 1 19.08.2009 – 25.08.2009 Samples of inlet water taken from drain Waste water – period 2 16.08.2009 – 01.09.2009 valve before all cleaning steps. Outlet water Waste water – period 3 02.09.2009 – 08.09.2009 taken from drain valve after cleaning. Sludge – period 1 19.08.2009 – 25.08.2009 Sludge taken from silo.

Sludge – period 2 16.08.2009 – 01.09.2009 Sludge – period 3 02.09.2009 – 08.09.2009

Mjøsa and Losna All sediment samples were collected with a Van veen grab (surface area 0,025 m 2). Sediments in the 0-2 cm layer were sampled.

Table 12 Sampling information for sediment localities, Mjøsa and Losna. All coordinates in UTM32, WGS84. Station North East Depth (m) Sampling period MLI-1 7517632.98 364068.84 37 19.08.2009 MLI-2 7514859.93 363712.70 53 19.08.2009 MLI-3 7511620.53 363734.93 94 19.08.2009 Losna-1 7551580.05 361658.46 27 18.08.2009

An overview of where the fish samples (trout) are taken is presented in the table below.

Table 13 Sampling information for where in Mjøsa trout was caught. Station North East Sampling period Description Øyresvika 6773657,54 577471,87 06.09.2009 – 12.09.2009 In the area of Øyresvika

Size data for the trout that was caught is presented in the table below.

Table 14 Sample data trout, Mjøsa. Species Average weight/size Standard deviation Trout length 50 cm 4,5 cm Trout weight 2279 g 682 g

38 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5.2.4 Tromsø area An overview of the sampling stations in Tromsø is presented in Figure 8.

Figure 8 Overview of sampling locations, Tromsø area

39 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Langnes RA Information about sampling at Langnes water treatment plant is given in Table 15.

Table 15 Sampling information, Langnes waste water treatment plant. Sample Sampling period Description Waste water – period 1 07.09.2009 – 13.09.2009 Samples of inlet ater were from water tap Waste water – period 2 14.09.2009 – 20.09.2009 after first screen. Otlett water taken form Waste water – period 4 1) 20.10.2009 – 27.10.2009 water tap after all cleaning steps (all screens). Sludge – period 1 08.09.2009 Samples taken from sludge filter. Sludge – period 2 15.09.2009 Sludge – period 3 22.09.2009 1) Sampled again because samples from period 3 where lost

Sandnessundet All sediment samples were collected with a Van veen grab (surface area 0.025 m 2). Sediments in the 0-2 cm layer were taken out. All stations coordinates were logged by a hand held GPS. In Sandnessundet there are relatively strong tidal currents, which mean that the sediments are rather coarse. As a result of this the samples taken in the vicinity of the discharge point for waste water from Langnes waste water facility was coarse (station SAS-1c). The two other stations were moved to locations closer to the shore where finer sediment was found.

Table 16 Sampling information for sediment localities, Sandnessundet. All coordinates in UTM32, WGS84. Station North East Depth Sampling period Description (m) SAS-1c 7761581.51 881550.09 20 07.09.2009 Coarse sand/shell sand SAS-4b 7761873.75 881987.40 5 07.09.2009 Finer sediments SAS-6b 7762083.52 882093.34 5 07.09.2009 Finer sediments

An overview of where and when the biota samples were taken is presented in Table 17 below. Cod were caught by use of traditional fishing gear, blue mussels were collected by diving and crabs were caught by use of a lobster pot.

There was not possible to catch edible crab in Sandnessundet. The specimen which was abundant enough to use were crabs of the genus Hyas (spider crab and lyre crab). Also, it was difficult to collect blue mussels from relevant locations so only one location of blue mussels was found.

Table 17 Sampling information for biota, Sandnessundet. Station North East Sampling period Description Fish 7762567.45 882009.23 08.10.2009 – 15.10.2009 In Sandnessundet Crab 7761794.14 882141.67 08.10.2009 – 15.10.2009 In Sandnessundet Blue mussel 7762281.20 882329.67 08.10.2009 – 15.10.2009 In Sandnessundet

Size data of the biological samples are given in the table below.

40 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 18 Sample data biota, Sandnessundet. Species Average weight/size Standard deviation Cod weight 1110 g 606 g Cod length 43 cm 7.5 cm Crab 4.2 cm 0.6 cm Blue mussel 5.6 cm 0.3 cm

5.2.5 Haugesund area An overview of the sampling stations in the Haugesund area is presented in Figure 9.

Figure 9 Overview of sampling locations, Haugesund area

41 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Res-Q The sedimentation basins at Res-Q had been emptied and cleaned during the summer which made it difficult to sample the sedimentation basins as planned. Therefore alternative sampling locations was chosen as described in Table 19.

Table 19 Sampling information, Res-Q fire fighting facility. Sample Sampling period Description Seepage water 04.09.2009 – 14.09.2009 Water samples taken from water tap after cleaning steps. ResQ-4 (2 replica of sediment) 04.09.2009 Samples taken from a drainage basin (about 2 m deep) near the shoreline ResQ-5 (3 replica of sediments 04.09.2009 Samples taken from bore hole just outside the exercise area. This bore hole drains both sedimentation basins inside the Res- Q area which were originally planned to be sampled .

Bleivika Sediment sampling by use of Van veen grab was not possible due to very coarse sediment (stones and hard bottom). Sediments had to be sampled by diving. All samples were taken in the area of the discharge point for seepage water from Res-Q. Station information is given in Table 20.

Table 20 Sampling information for sediment localities, Bleivika. All coordinates in UTM32, WGS84. Station North East Depth Sampling period Description (m) ResQ-1 6598929.22 286540.97 15 11.10.2009 Relatively coarse sediment with shells ResQ-2 6598984.22 286627.93 15 11.10.2009 Relatively coarse sediment with shells ResQ-3 6599006.2 286510.04 32 11.10.2009 Relatively coarse sediment with shells

An overview of where and when the biota samples were taken is presented in Table 21 below. There were difficulties in sampling enough cod so a mix of species where chosen; one sample containing saithe ( Pollachius virens ), one sample containing pollock ( Pollachius pollachius ) and a third sample containing 2 cods and 2 saithes (Cod and crab were caught by use of fish net, blue mussel was collected from boat).

Table 21 Sampling information for biota, Bleivika. Station North East Sampling period Description Fish 6599506.30 286174.84 October 2009 Bleivik area Crab 6599506.30 286174.84 October 2009 Bleivik area Blue mussel-1 6602138.18 286388.46 October 2009 Area of Sørasund Blue mussel-2 6602238.54 287606.11 October 2009 Area of Smørsund Blue mussel-3 6602949.91 286884.07 October 2009 Area of Hidlesvågen

Size data of the biological samples are given in the table below.

42 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 22 Sample data biota, Sandnessundet. Species Average weight/size Standard deviation Fish weight 451 g 301 g Fish length 30 cm 6.3 cm Crab 14 cm 1.3 cm Blue mussel-1 5 cm 0.7 cm Blue mussel-2 4 cm 0.3 cm Blue mussel-3 4.8 cm 0.8 cm

5.2.6 Bergen area An overview of the sampling stations at Flesland airport and Puddefjorden, both in the Bergen area are presented in Figure 10 and Figure 11.

43 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Figure 10 Overview of sampling stations at Flesland airport, Bergen area. Note: Soil sample symbol show center position for the soil sample at the fire fighting training ground

44 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Figure 11 Overview of sampling station for fish in Puddefjorden, Bergen area

Flesland airport At the fire fighting training area soil was sampled by use of an earth auger in increasing distance from the training site (centre, 10, 20, 30, 40, 50, 75, 100, 150 and 200 m respectively). Five samples from 0-5 cm of the soil top layer was sampled and analysed as a pooled sample.

Table 23 Sampling information for soil and seepage water, Flesland airport. Sample North East Sampling period Description Soil 6691131.04 291775.93 03.09.2009 10 samples collected in increasing distance from centre position Seepage water 6691092.57 291667.94 09.11 – 16.11.2009 Seepage water drains to a stream that goes out in Langavatn. Samples were taken from bore hole close to Langavatn by use of pump.

45 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Langavatn, Flesland brygge and Puddefjorden Langavatn is the receiving water for seepage water from Flesland airport’s fire fighting drill site. Langavatn is a freshwater lake where sediments and fish (trout) were sampled. Langavatn then drains to Flesland brygge (marine receiving waters) where blue mussels and edible crabs were sampled.

Some rearrangements in the sampling program had to be done because it was not possible to catch trout in the Losna river (Mjøsa area). Puddefjorden was chosen as a new locality for fish.

Table 24 Sampling information for sediment localities, Langavatn. All coordinates in UTM32, WGS84. Station North East Depth (m) Sampling period Langavatn -1 6691010.53 291307.65 2 03.09.2009 Langavatn -2 6691028.68 291146.95 4 03.09.2009 Langavatn -3 6690831.71 291136.75 3 03.09.2009

Table 25 Sampling information for biota, Langavatn, Flesland brygge and Puddefjorden. All coordinates in UTM32, WGS84. Station North East Sampling period Fish - Langavatn 6691114.58 291242.01 12.11.2009 – 23.11.2009 Fish - Puddefjorden 6700873.82 295871.98 02.11.2009 – 12.11.2009 Crab 6690684.66 290449.47 12.11.2009 – 23.11.2009 Blue mussel-1 6690706.03 290423.84 12.11.2009 – 23.11.2009 Blue mussel-2 6690584.80 290479.38 12.11.2009 – 23.11.2009 Blue mussel-3 6690554.89 290478.31 12.11.2009 – 23.11.2009

46 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

5.3 Chemical analysis

5.3.1 Brominated flame retardants

Chemicals Analytical standards ( 12 C) HBB, PBT, PBEB, TBP, ATE, DBPE, TBPH, EHTBB, BTBPE and DBDPE were bought from Wellington Laboratories (Guelph, ON, Canada) and TBBPA- AE was bought from Chiron (Trondheim, Norway).These where all of high purity (>98 %). Technical grade TBPA was bought from Sigma-Aldrich, BTBPI was bought from Chiron (Trondheim, Norway), and TBBPA-DBPE from TCI (Tokyo, Japan). 13 C-Labeled internal standards of HBB, TBP, and BTBPE were bought from Wellington Laboratories (Guelph, ON, Canada) and 13 C-labeled BDE 28 and 209 were bought from Cambridge Isotope Laboratories (Andover, MA, USA). Labeled ( 13 C) PCB 208 was used as recovery standard and was bought from Cambridge Isotope Laboratories (Andover, MA, USA).

Methanol (SupraSolv), (SeccoSolv) and ethyl acetate (LiChrosolv) were bought from Merck, Germany, hexane (Picograde) and acetone (Picograde) was bought from Promochem, dichloromethane (glass distilled), cyclohexane (glass distilled) and toluene (glass distilled) were bought from Fluka, Switzerland

Extraction and clean-up All samples were stored in freezer (-20 ºC) before pre-treatment and extraction, and all glassware were washed before use and heated to 550 ºC. Individual extraction and clean-up procedures were developed for the various types of matrixes. Sediment, sludge, and moss were freeze-dried prior to extraction. After freeze-drying, the dry content were checked after placing sub samples in 105 ºC overnight. For sediment and sludge the loss of ignition was recorded after heating the sub samples to 550 ºC for two hours. Sediment (n=36, dry weight, dw = 8.95-28.6 g), sludge (n=9, dw=3.32-7.60 g), and moss samples (n=4, dw=3.08-5.32 g) were spiked with internal standards and extracted in Soxhlet apparatus with methanol as solvent overnight for at least 15 hours. The extracts were divided by weight in two parts; the first for analysis of HBB, PBT, PBEB, BTBPE and DBDPE with gas chromatography interfaced to mass spectrometry (GC-MS) and the second for analysis of TBP, ATE, DBPE, TBPA, TBPH, EHTBB, TBBPA-DBPE, TBBPA-AE, and BTBPI with high pressure liquid chromatography atmospheric pressure photo ionization tandem mass spectrometry (HPLC- APPI-MS/MS).

Prior to GC-MS analysis the sub samples were evaporated until almost dryness and redissolved in hexane before clean-up on a multi-layer silica column. The silica columns were packed with glass wool, 3 g KOH-silica, 3 g neutral silica, 6 g of 40% (w/w) H2SO 4silica, and 3 g Na 2SO 4, rinsed with solvent, and eluted with 60 ml of a mixture of hexane and dichloromethane (1:1). Activated copper were added to sediment and sludge samples for reduction of sulphur. The copper was activated with hydrochloric acid and thereafter washed with water, methanol, and hexane. Copper was added to the samples until reaction with sulphur ceased. The samples were reduced in volume to approximately 100 µl toluene and recovery standard was added.

Prior to HPLC-APPI-MS/MS analysis samples were evaporated until approximately 1 ml methanol. Some samples contained water and were evaporated until dryness and redissolved

47 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) in methanol. Samples that contained precipitations were centrifuged and the supernatant were transferred to a clean vial. Selected samples were additionally cleaned on an open neutral silica column. In that case the samples were evaporated until almost dryness and redissolved in hexane. Silica columns were packed with glass wool, 5 g neutral silica, and 3 g Na 2SO 4, rinsed with solvent, and eluted with 60 ml hexane:dichloromethane (1:1). Samples were evaporated until dryness and redissolved in1 ml methanol.

Water samples (n=24, v=1.9-7.5 l) were filtrated through glass microfibre filters (GF/B, Whatman) followed by 0.45 µm nylon membrane filters (Sartourius, Goettingen, Germany), and 90 mm SPE disks (ENVI-18 Dsk, Supelco, Bellafonte, USA). Internal standards were added before filtration, and the samples were acidified with hydrochloric acid to pH 3. The filters and the SPE disks were air dried in a fume hood and extracted and cleaned-up as described above. The filtrate was discarded. Air filters (n=8) were spiked with internal standards and extracted as described above. No clean-up was considered necessary for indoor air samples, while outdoor air samples were cleaned-up as described above.

Needle samples (n=4, wet weight, ww=34-60 g) were extracted with dichloromethane in an ultrasonic bath for 10 minutes. The procedure was repeated two times and the extracts were combined. The amount of extracted waxes was determined gravimetrically. The samples were cleaned-up as described above. Prior to HPLC-APPI-MS/MS analysis all needle samples were cleaned-up with neutral silica.

Fish liver (n=15, ww=1.0-3.8 g), crab (n=7, ww=2.6-12 g), and mussel samples (n=3, ww=12-24 g) were homogenized in Na 2SO 4, and extracted on an open column with acetone:hexane (5:2) and hexane:diethyl ether (9:1). Internal standards were added and the lipid contents were determined gravimetrically. The extracted samples were dissolved in cyclohexane:ethyl acetate (3:1) and lipids were reduced with gel permeation chromatography. The gel permeation chromatography column (1.5 cm inner diameter) was wet-packed in-house to a height of 40 cm with 25 g of SX-3 Bio-beads (Bio-Rad Laboratories, Hercules, CA, USA) that had been pre-swollen in cyclohexane:ethyl acetate (3:1) for 2 h. The flow rate was set to 2 ml/min, and the fractions containing the BFRs were collected between 20 and 50 min. The BFR fraction was divided in two parts; the one for analysis with GC-MS and the second for analysis with HPLC-APPI-MS/MS. The sub samples for HPLC-APPI-MS/MS analysis evaporated until dryness and redissolved in 1 ml methanol. The GC-MS sub samples were additionally cleaned on a H 2SO 4-silica column (6 g of 40% H 2SO 4-silica) eluted with 60 ml hexane:dichloromethane (1:1). The samples were reduced in volume to approximately 100 µl toluene and recovery standard was added.

Chemical analysis HBB, PBEB, PBT, BTBPE, and DBDPE were analyzed with GC-MS in electron capture negative ionization (ECNI) mode. A MSD 5975 quadrupole GC-MS system (Agilent) was used for the analyses. The samples were injected using a programmable temperature vaporizing (PTV) injector operated in pulsed splitless mode with the following temperature program: 80 ºC for 0.10 min followed by heating at 720 ºC/min to 300 ºC (held for 2 min). The pressure pulse was set to 15.8 psi for 1.80 min. For separation of the BFRs a 15 m Rxi- 5sil MS column (0.25 mm i.d. × 0.10 µm film thickness; Restec, USA) was used, with helium as the mobile phase (flow rate 1.5 ml/min), and the following oven temperature program: 80 ºC (held for 3 min) rising to 240 ºC at 25 ºC/min, then to 315 ºC at 10 ºC/min

48 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

(held for 5.5 min). The ion source temperature was set to 230ºC and methane was employed as the reagent gas. Ions were recorded in SIM mode and are shown in Table 26.

TBP, ATE, DBPE, TBPA, TBPH, EHTBB, TBBPA-DBPE, TBBPA-AE, and BTBPI were analyzed with HPLC-APPI-MS/MS. 20 µl of samples were injected on a Hypersil Gold Phenyl column (50×2.1 mm, particle phase 3 µm, pre-column 10×2.1 mm). A gradient of water and methanol was used for separation of compounds. Monitored parents and daughter ions are shown in Table 27.

Table 26 Retention time and monitored ions during GC-MS analysis Compound Retention time (min) Monitored ions (m/z) PBT 7.94 485.6, 487.6 PBEB 8.17 499.6, 501.6 HBB 8.87 549.5, 547.5 BTBPE 14.19 329.8, 331.8 DBDPE 21.09 79, 81

Table 27 Retention time and monitored ions during HPLC-MS/MS analysis Compound Retention time (min) Parent ions (m/z) Daughter ions (m/z) TBP 6.74 250.8 79, 81 TBPA 6.33 463.6 382.9, 384.9 ATE 8.34 290.8 79, 81 DPTE 7.70 290.8 79, 81 EHTBB 8.80 356.7, 358.7 312.9, 314.9 TBPH 9.65 512.8, 514.8 469.1, 471.1 TBBPA-DBPE 9.54 544.9, 542.9 409.0, 424.2 TBBPA-AE 8.64 542.8, 544.8 502.2, 504.2 BTBPI 10.43 951.4, 953.4 951.4, 953.4

Quality assurance and control Laboratory blanks were analyzed in parallel to the samples to ensure that contamination during homogenization, clean-up and instrumental analysis did not significantly influence the results.

A BFR was considered detected if its signal-to-noise ratio was > 3. The limit of detection (LOD) was based on the signal to noise in the quantification standard. The limit of quantification (LOQ) was set to ten times the signal-to-noise ratio. If a BFR was present in a laboratory blank the LOQ was set to three times the level detected in the blank.

Before the extraction and clean-up of samples the recovery of the analytes was tested. This was done by adding known amounts of the 14 BFRs of interest to three different matrices. The tested matrices were dry artificial OECD soil, cod liver oil, and river water from Umeå (Umeälven). Duplicates of each matrix were analysed. The recoveries (shown in Figure 12) from dry soil was relatively high for all BFRs (37 % for ATE, >48 % for all other BFRs) and the recoveries from fish oil were high for all BFRs (38% for BTBPI, >44% for all other BFRs) except for TBPA that was not detected. The recoveries from spiked water were relatively high for PBT, PBEB, HBB, and BTBPE, but very low for several other BFRs. In this test the water was not acidified before extraction which could explain the low recovery

49 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) of TBP. In real samples the water was pH adjusted to 3 before extraction which probably increased the recovery of TBP.

120% Fish oil Soil 100% Water

80%

60%

40% Recovery(%)

20%

0%

B E E E B H E PE P T -AE P B PBT B HBB TBP A T TBPI P TB DPTE PA DB H TBP B TBPA B DBD B E TB BPA- B T

Figure 12 Average recoveries for the 14 BFRs during extraction, clean-up and chemical analysis in three tested matrices.

13 C-labeled TBP, BDE 28, HBB, BTBPE and BDE 209 were added to the samples in order to compensate for losses during clean-up. 13 C-labeled TBP was used as internal standard for ATE and TBP, 13 C HBB was used for HBB, PBT and PBEB, 13 C-labeled BTBPE was used for DBPE, TBPA, TBPH, EHTBB, TBBPA-DBPE, TBBPA-AE, and 13 C-labeled BDE 209 was used for BTBPI and DBDPE. The recoveries of the internal standards were calculated after GC-MS analysis and average recoveries for each matrix are shown in Table 28.

Table 28 Average recovery of internal standards after GC-MS analysis Compound Sediment Water Sludge Air Fish, crab and Moss and samples samples samples samples mussel samples needle (n=36) (n=24) (n=9) (n=8) (n=25) samples (n=8) BDE 28 91 % 72 % 119 % 78 % 75 % 86 % HBB 72 % 56 % 68 % 45 % 74 % 74 % BTBPE 84 % 46 % 80 % 49 % 71 % 66 % BDE 209 53 % 46 % 102 % 20 % 74 % 95 %

5.3.2 Polyflourinated compounds

Chemicals Ammonium acetate (>99%, pa for HPLC) and n-hexane (Pestanal) were purchased from Fluka (Steinheim, Germany), HPLC-grade methanol (MeOH) was from Fisher Scientific (Leicestershire, UK). Acetonitrile (AcN) and water were from Lab-scan (Sowinskiego, Poland). Ammonium hydroxide (NH 4OH) 25%, sodium hydroxide (NaOH) p.a., sodium acetate p.a., hydrochloride acid (HCl) and glacial (100%) were all purchased from E. Merck (Darmstadt, Germany). Native linear perfluorinated sulfonates (potassium

50 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) perfluorobutanesulfonate (PFBS), sodium perfluorohexanesulfonate (PFHxS), sodium perfluorooctanesulfonate (PFOS), and sodium perfluorodecanesulfonate (PFDS)) and perfluorinated carboxylates (pentanoic- (PFPeA), hexanoic- (PFHxA), heptanoic- (PFHpA), octanoic- (PFOA), nonanoic- (PFNA), decanoic- (PFDA), undecanoic- (PFUnDA), dodecanoic- (PFDoDA), tridecanoic- (PFTrDA), and tetradecanoic acid (PFTeDA)) were from Wellington Laboratories (Guelph, Canada). Labeled standards were also from 18 13 13 13 13 Wellington Laboratories( O2PFHxS, C4PFOS, C2PFHxA, C4PFOA, C8PFOA, 13 13 13 C5PFNA, C2PFDA, C2PFUnDA). 7H -perfluoroheptanoic acid (7H-PFHpA) (98%) was purchased from ABCR (Karlsruhe, Germany) and 1H,1H,2H,2H -perfluorooctane sulfonate (6:2 FTS) (purity not given by supplier) was from Interchim (Montlucon, France).

Extraction and clean-up Soil, dry sediment and biota were stored at -20°C until analysis. Wet sediment was stored at 4°C until drying. Soil and sediment samples were air dried prior to extraction. Wet samples were weighed into methanol-washed PP-tubes, corresponding to a dry weight of approximately 1 g, and were left in a vented hood for 4-5 days. Dried samples were thereafter weighed and the water content was calculated before extraction. In addition, the dry weight was controlled by 24h drying at 105°C. All biological samples were homogenized before extraction (Ultra-Turrax, IKA). From the homogenate, 1 g of sample was taken in the analytical procedure.

The dried/homogenized soil/sediment/biological samples were added 50µL of a 0.2 ng/mL 13 internal standard (IS) mixture consisting of C4-labeled sulfonates (PFHxS and PFOS) and carboxylates (PFHxA, PFOA, PFNA, PFDA, PFUnDA). Thereafter, 0.4 mL of a 0.2 M NaOH (in methanol) solution was added to the samples and left for 30 min, in order to degrade the organic structures in the matrices. Extraction was performed using 4 ml of AcN, first in an ultrasonic bath for 15 min and thereafter on a shaking table for 15 min. The samples were neutralized (20µL of a 4M HCl solution in methanol), centrifuged and the supernatant decanted. The extraction was repeated once more and the two extracts were combined. n-hexane (corresponding to a volume of 2:1 sample extract:hexane) was added and the mixture were thoroughly shaken for 30 seconds. The step was repeated twice with hexane being removed between each time. Thereafter, the AcN-extract was transferred to a new tube, prepared with 50 mg dispersive carbon (Supelclean ENVI-Carb (20/400 mesh), Supelco Bellefonte, PA) and 100 µL glacial acetic acid and thoroughly shaken for 30 sec. The carbon was removed by filtration using a 0.2 µm GHP membrane (Pall, East Hills, NY, USA), and the solution was collected in LC-vials and reduced in volume by a gentle stream of nitrogen gas. Recovery standards (RS), 5µLof a 0.4 mg/mL 7H -PFHpA solution and 10µL 13 of a 0.2 mg/ml C8-PFOA solution, were added to the final extract together with 2mM ammonium acetate (aq). Blank samples, using empty extraction tubes, and field blanks where performed in parallel with each batch of samples, and were treated in exactly the same manner as the other samples.

Water samples (200-500 mL) were stored at 4°C until analysis and filtered through glass microfiber filters (GF/B, Whatman) before extraction using Oasis WAX (6cc/150mg, Milford, MA, USA) according to standard method ISO 25101 (International Organisation for standardisation). The cartridges were conditioned with 4 ml of 0.1% NH 4OH/methanol solution, 4 ml methanol, and 4 ml water. Same internal standards were added to the water as for the soil/sediment/biota, before vaccum was used to run through the water samples at a flow rate of approximately 1 drop per second. Sodium acetate buffer (4 mL, 0.025 M) was added after the sample, and the eluate was discarded. After drying the cartridges using

51 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) vacuum suction 4 ml methanol was added and discarded and the analytes were then eluted with 4 ml of 0.1 % NH 4OH/methanol solution at a rate of one drop per second. The eluates were collected, filtrated and evaporated to suitable volume with a gentle stream of nitrogen gas. Recovery standards and ammonium acetate were added as for the soil/sediment/biota samples. Extraction and field blank samples were prepared with HPLC-grade water (Fisher Scientific) and were treated exactly in the same way as the samples.

Chemical analysis Sample extracts were analyzed twice, once as described and once diluted five times. Analysis was performed using an Acquity UPLC coupled to a Quattro Premier XE (Waters Corporation, Midford, US) with an atmospheric electrospray interface operating in negative ion mode (ES-MS/MS). Capillary voltage was set to 0.6 kV, source and desolvation gas (N 2, 950L/hr) temperatures were 150 and 450°C. Cone voltages and collision energies were optimized for each transition. Multiple reaction monitoring was used monitoring the product ions given in table 1. Concentration of the analytes in the samples was calculated using internal standard quantification. The internal standard closest in retention time was used for those compounds that did not have a corresponding labeled internal standard (PFBS, PFDS, 6:2 FTS, PFPeA, PFHpA, PFTrDA, PFTeDA). Only the linear isomer was quantified when isomers were present. Separation was performed on an Acquity BEH C18 2.1 x 50 mm, 1.7 µm kept at 50°C. An extra guard column (PFC isolator, Waters Corporation, Midford, US) was inserted between the pump and injector to remove any PFC originating from the LC system. Injection volume was 10 µL and the flow rate was set to 400 µl/min. A gradient program was employed delivering mobile phases consisted of 2 mM ammonium acetate in methanol, and 2 mM ammonium acetate in water.

Table 29 Retention time and monitored ions during LC/MS/MS analysis Compound Retention time Precursor ion Product ions (min) (m/z) (m/z) PFBuS 2.25 299.15 79.6, 98.6* PFHxS 3.45 399.04 79.6, 98.6*, 118.9 PFOS 4.16 498.70 79.7, 98.7* PFDS 4.68 598.84 98.8* 6:2 FTS 3.80 426.97 80.7, 407.0* PFPeA 2.03 262.97 218.9* PFHxA 2.85 313.13 268.7*, 118.7 7H -PFHpA (RS) 2.57 345.03 281.0 PFHpA 3.40 363.18 319.0*, 168.0 PFOA 3.81 412.9 218.9, 368.9* PFNA 4.15 463.04 418.9*, 218.9 PFDA 4.44 513.04 468.7*, 168.8 PFUnDA 4.69 562.78 518.7*, 319.0 PFDoDA 4.89 612.91 568.7*, 168.8 PFTrDA 5.07 662.84 618.6*, 268.9 PFTeDA 5.22 713.0 668.7*, 168.9 *Ion used for quantification

52 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Quality assurance and control Qualifier transitions were monitored for each compound to verify the identity of the quantified peaks (exception for PFPeA and PFDS) according to EU Council Directive 96/23/EC (European Commiosion 2002). A result with dissimilarity >50% between qualifier and quantification ions have been marked with an asterisk in the results tables. Blank methanol/water injections were carried out repeatedly during the analysis to monitor possible contamination from the instrument. Extraction blanks and field blanks were also monitored. Trace levels of 6:2 FTS (<0.3 ng/g) and PFOS (<0.1 ng/g) were found in the extraction blanks and was withdrawn from the concentration found in the samples. Field blanks did not contain levels above the extraction blank levels. The limit of detection (LOD) was set to three times the noise level. If trace levels were found in extraction blanks the LOD was set to twice the blank signal. The PFC LOD´s varies since the background noise and levels in extraction blanks varied between matrices and also to some extent between runs.

The recovery of the individual polyfluorinated compunds was studied by spiking known concentrations of native compounds to non- or low contaminated samples (Figure 13). For this, non-contaminated sediment from lake Hjälmaren (Sweden), laboratory produced water and two low contaminated samples from RESQ (blue mussel from station 3 and fish liver replicate 3) were utilized. Unacceptable recoveries were obtained for 6:2 FTS in fish and mussel (400-500%) and for PFDA in soil and PFTeDA in biota (9-15%) and these compounds were therefore not quantified in mentioned matrices.

Figure 13 Average recoveries of polyfluorinated compounds added to water, fish liver, mussel and sediment.

The recovery of the internal standards added to each sample was calculated to assess the recovery of the analytes, and to monitor possible ion signal effects (Table 30). Recoveries between 50-150% are generally considered acceptable, whereas results based on recoveries both below and above this criterion are considered less reliable. The recovery can be >100% if there are compounds present in the sample matrix that enhances the ionization of the analytes in question. In addition, a recovery >100% can be obtained considering the uncertainty of the method . In this study signal effects were seen for soil (low recoveries) and for PFNA in mussel and crab (high recoveries).

53 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 30 Average recoveries of internal standards Internal standard Soil Sediment Water Mussel Crab Fish liver (n=10) (n=17) (n=6) (n=6) (n=8) (n=7) 13 C-PFHxS 60 67 48 82 57 68 13 C-PFOS 31 50 60 89 103 47 13 C-PFHxA 39 64 78 104 107 91 13 C-PFOA 52 66 91 100 80 86 13 C-PFNA 25 52 55 161 158 67 13 C-PFDA 68 84 76 112 124 83 13 C-PFUnDA 25 47 59 117 99 55

Reproducibility and precision of the chemical analysis were estimated by analyzing samples (sludge and fish muscle) originating from an international interlaboratory comparison study (ILS). Reproducibility expressed as relative standard deviation (RSD) was <15% for compounds above 1 ng/g fish and up to 30% for compounds present at lower levels. In sludge the reproducibility was <35% but up to 55% for PFDS and compounds with low levels. Precision estimated as deviation from the median results obtained in the ILS was for biota <15% except for compounds with levels < 1 ng/g for which up to 50% deviation was seen. For sludge the deviation from the ILS median was <31% and up to 65% for levels < 1 ng/g.

54 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

6. Results and discussion

In order to compare the findings in this screening it’s considered important to try to compare measured concentrations with toxicity data from the literature where such data exists. Below a description and presentation of the values used are given.

For most of the investigated new BFRs there are not enough toxicity data available for determination of a predicted no-effect concentration (PNEC). In these cases read-across, or the transfer of the hazard profile of one substance to another with a similar structure, can be used for filling data gaps. There is, of course, uncertainties related to such an approximation. There are established PNECs available for PBDEs, hexabromocyclododecane (HBCD) and for tetrabromobisphenol A (TBBPA) (European Union 2001, 2003, 2007, Swedish Chemical Inspectorate 2006). Here, these PNECs were transferred to the new investigated BFRs by categorization of the new BFRs as follows. New BFRs with a molecular weight below 700 g/mol are considered as similar to pentaBDE, except for TBBPA-AE which together with TBBPA-DBPE was considered to be most similar to TBBPA. TBPH was the only new BFR with a molecular weight between 700 and 900 g/mol, and were considered as similar to octaBDE. DBDPE and BTBPI have a molecular weight above 900 g/mol and were considered to be most similar to decaBDE. None of the new BFRs were judged to be most similar to hexabromocyclododecane as none of these have a cycloaliphatic structure.

Table 31 Adapted PNEC values for “new” BFRs based on PNEC values for “old” BFRs. Se also Appendix II.

BFR similar to PNEC water PNEC sed HBB pentaBDE 0.53 µg/l 1.55 mg/kg dw PBT pentaBDE 0.53 µg/l 1.55 mg/kg dw TBP pentaBDE 0.53 µg/l 1.55 mg/kg dw PBEB pentaBDE 0.53 µg/l 1.55 mg/kg dw ATE pentaBDE 0.53 µg/l 1.55 mg/kg dw DPTE pentaBDE 0.53 µg/l 1.55 mg/kg dw TBPA pentaBDE 0.53 µg/l 1.55 mg/kg dw EHTBB pentaBDE 0.53 µg/l 1.55 mg/kg dw BTBPE pentaBDE 0.53 µg/l 1.55 mg/kg dw TBPH octaBDE ≥0.2 µg/l ≥ 127 mg/kg dw DBDPE decaBDE ≥0.2 µg/l ≥ 127 mg/kg dw BTBPI decaBDE ≥0.2 µg/l ≥ 127 mg/kg dw TBBPA-AE TBBPA 0.25 µg/l 0.054 mg/kg ww TBBPA-DBPE TBBPA 0.25 µg/l 0.054 mg/kg ww

For PFOS much information can be find in the literature but for the other PFCs investigated there is much less data. A summary of the rewied literature are presented in the table and text below.

55 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 32 Effect levels of selected PFCs in the litterature. PNEC/NOEC PFC Matrix Test organism Value PNEC PFOS Fresh water Water living organisms 25 µg/L PNEC PFOS Fresh water sediment Sediment living organisms 67 µg/kg PNEC PFOS Salt water Water living organisms 2.5 µg/L PNEC PFOS Salt water sediment Sediment living organisms 6.7 µg/kg NOEC PFOS Soil Earth worm, reproduction 10 mg/kg NOAEL PFBS Food intake Rat 200 mg/kg/day NOAEC PFBS Food intake Quail 900 mg/kg w.w. feed US-EPA 1) PFBS Rat >500 mg/kg EU 2) PFBS Rat >2000mg/kg NOEL PFBS Rat >1000 mg/kg/day NOAEL 6:2 FTS Liver weight increase 30 ppm EC10 6:2 FTS Soil Earth worm, conc in soil 21 mg/kg LD50 6:2 FTS >500 mg/kg

1) Acute toxicity classification, 2) Acute toxicity classification

PFHxS: Rats given up to 10 mg/kg/day of PFHxS did not display any reproductive or developmental toxicity. Final liver concentration in the rats after exposure was up to 594 µg/g (42 days exposure).

PFDS : No toxicological data found.

Below only detected compounds are discussed. Regarding environmental concern the following criteria have in principle been applied:

(i) If the compound was not detected or only detected in samples not taken in the receiving environment it is assessed to be of no or little environmental concern. Included in this category are for example water, sediment and sludge from waste water facilities and waste disposal sites.

(ii) If the compound was detected in recieving environment it is assessed as being of moderate environmental concern. This is nuanced based on the PNEC and NOEC values presented in Table 31 and Table 32.

(iii) If the compound was identified in biological samples it is automatically assessed as being of environmental concern. Detection of substances in needles is not considered to represent biological samples. It’s assumed that the pollutants are associated to the waxes on the surface of the needles. In this regards needles are considered as passive air samplers.

Also, in the tables in Chapter 6.1 and 0 results for each detected compound are shown. In these tables LOD and LOQ is shown for BFR’s but not for PFC’s. The reason for this is that LOD and LOQ for PFCs vary among different samples and matrices (analytical disturbances).

56 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

6.1 Brominated flame retardants

In Table 33 an overview of detected investigated “new” BFRs is presented and a presentation of the concentration of the investigated BFRs in the different environmental compartments is presented in Figure 14, expect for air where only HBB was detected in outdoor air from Drammen (0.0013, 0.0014 and 0.0102 ng/m 3 in the 3 samples).

Table 33 Overview of positive detections of the investigated new BFR’s. +: Detected, O: detected in single replicate and/or very close to detection limit -: not detected Compound Sediment Sediment Sludge waste Waste Seepage Biological Air Air receiving waste water facility water water material outdoor indoor water disposal PBT + o - o + - - - PBEB + - - - + - - - HBB + o + + + o + - BTBPE + o + + + - - - DBDPE + - + + + o - - DPTE ------TBPA ------TBP o - - - - + - - ATE ------TBBPAAE + + - + + - - - BTBPI - - - - + - - - EHTBB ------TBBPA-DBPE - - - o + o - - BEHTBP ------

Findings in sludge

DBDPE Lillehammer RA BTBPE HBB DBDPE Solumstrand RA BTBPE HBB DBDPE Langnes RA BTBPE HBB

0 1 2 3 4 5 6 7 8 910 ng/g d.w sludge

Figure 14 Concentrations of detected BFR’s in the different environmental compartments with indication of minimum and maximum values. For biological material concentrations are given in wet weight for all samples except for moss samples where concentrations is given on dry weight basis.

57 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Findings in sediment

HBB Lindum

PBT DBDPE HBB Hellik Teigen PBEB Lindum TBBPAAE BTBPE PBT Hellik Teigen TBBPAAE BTBPE TBP Mjøsa HBB Mjøsa Drammen TBBPAAE BTBPE TBP Tromsö

0,0 1,0 2,0 3,0 4,0 5,0 HBB Tromsö ng/g sediment

PBT

0,0 0,2 0,4 0,6 0,8 1,0 1,2 1,4 1,6 1,8 2,0 ng/g sediment

Findings in water

Lindum HBB PBT BTBPI TBBPAAE Hellik Teigen HBB PBEB PBT

Lillehammer-outlet HBB Lindum BTBPE TBBPA_DBPE Hellik Teigen BTBPE DBDPE Lillehammer-inlet HBB Lillehammer-outlet PBT Lillehammer-inlet BTBPE Solumstrand-outlet HBB Solumstrand-outlet PBT Solumstrand-inlet DBDPE TBBPAAE Solumstrand-inlet HBB Tromsø-outlet TBBPA_DBPE Tromsø-inlet TBBPAAE Tromsø-outlet HBB 0 50 100 150 200 ng/l water Tromsø-inlet HBB

0 5 10 15 20 25 30 35 40 ng/l water

Findings in biological material

Hurum-moss TBP

HBB Tromsø-crab TBBPA_DBPE Tromsø-fish Drammen-crab Hurum-pine DBDPE 0 20 40 60 80 100 120 140 HBB ng/g biological material

TBP Drammen-mussel

0,0 0,2 0,4 0,6 0,8 1,0 1,2 1,4 1,6 ng/g biological material

Figure 14 continues

58 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Tetrabromobisphenol A bis (2,3 – dibromopropyl-ether) - TBBPA – DBPE

Results from this study A summary of positive detections of TBBPA – DBPE are presented in Table 34. TBBPA - DBPE was detected in seepage water (average 80 ng/l) from a car demolishing site in Hokksund area (Hellik Teigen), in one influent waste water sample from Langenes RA (18 ng/l) and in pine needles just north of Hurum Energigjenvinning (incineration of waste in the Drammen area) at a concentration close to the detection limit. Note that since the recovery of TBBPA - DBPE in spiked water (as can be seen in Figure 12) were very low there are great uncertainties related to the quantified concentrations in water in the present study.

Table 34 Positive detections of TBBPA – DBPE. Location Water ng/l Pine ng/g w.w. Langenes RA 18 n.a. Hellik Teigen 15.8 – 159.6 n.a. Hurum Energigjenvinning n.a. 0.16 LOQ 1.28 0.4 LOD 0.38 0.1 n.a: Not analysed

Comparison to previous studies In Pearl River Delta (China) TBBPA-DBPE levels were measured in sediment (<1.5 – 2300 ng/g d.w), in air (241-1240 pg/m3), sewage sludge (238 – 8946 ng/g d.w) and in farmland soil (17 – 60 ng / g d.w). TBBPA – DBPE was not detected (detection limit 1.5 ng/g dw) in fish, birds and dust (Shi et al., 2009).

In the present study TBBPA-DBPE was not detected in sewage sludge, sediment or air, however, the presence of TBBPA-DBPE in pine needles indicates current or recent presence of TBBPA-DBPE in air in Hurum.

It was detected in biological material (pine) but there are no investigations in the cited literature where pine is included.

Toxicological effects Most toxicological studies in the cited literature are studies on mice and considered little relevant for the findings here. Compared to an adapted PNEC water of 250 ng/l (se Table 31) the levels here are lower.

Fate It is assumed that that TBBPA-DBPE does not accumulate in fish (carp, WHO 1995) and that it is a relatively short half life due to hydrolysis. The elimination product TBBPA, bis(bromopropenyl ether), might be the more prevalent compound in sediments in a similar manner as DDE is for DDT (Rahm et al., 2005).

Concluding remark The substance has one of the lowest water ’s of the BFR’s included in this study and a relatively high logP ow (octanol water partitioning coefficient, high values suggest that a

59 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) compound is hydrophobic meaning here that it will sorb to particles to a high degree). This in combination with the possibility that the compound may not be long lived in the environment due to hydrolysis one may expect TBBPA – DBPE not to be very prevalent in the environment.

Because of the fact that the compound is found in waste water at Langenes RA and in seepage water from Hellik Teigen there is sources for this compound in these two areas. It is interesting that the compound is not detected in sediment or sludge and sediment from the same localities where it’s detected in water (Langnenes RA and Hellik Teigen) considering a relatively high logP ow . In this regard it may be interesting to analyse the elimination product TBBPA bis(bromopropenyl ether).

TBBPA-DBPE is detected in water and pine needles and is found in samples both in northern Norway (Tromsø area) and south eastern Norway (Drammen and Hokksund area). It is detected in seepage water draining receiving water (Loselva). Overall TBBPA-DBPE is assessed to be of moderate environmental concern.

Pentabromoethylbenzene - PBEB Results from this study PBEB was detected in low levels in seepage water (average 0.9 ng/l) and in sediment (average 0.04 ng/g) from a car demolishing site in Hokksund area, see Table 35. The levels are much lower than BDE 209 which was in the range 1000 ng/l in seepage water and 10 ng/g d.w in sediment from the same locality. Sediment samples were taken at the discharge point for seepage water in the receiving water (Loselva). PBEB was not detected in any other sample in this screening study.

Table 35 Positive detections of PBEB. Location Water ng/l Sediment ng/g d.w Hellik Teigen 0.6 – 1.3 0.004 – 0.1 LOQ 0.12 0.01 LOD 0.04 0.004

Comparison to previous studies PBEB has previously been detected in quite high concentrations in outdoor air. A study by Hoh et al. reported a relatively high abundance of PBEB in the atmosphere of Chicago (summer of 2003). PBEB was detected in both gas and particle phases (520 pg/m 3 gas phase and 29 pg/m 3 in particle phase) Screening of air samples in three locations in UK and Ireland (reference site) reported a mean concentration of PBEB of 30 pg/m 3 in the southwest Oxford (Lee et al., 2002). In this study no PBEB was detected in indoor air samples from Elköp, Drammen or in outdoor air. The detection limits in outdoor and indoor air were 1 and 30 pg/m 3, respectively. It’s worth mentioning that only the particle phase was analysed in this study which might influence on the results considering that a relatively high proportion of this compound can be found in the gas phase. Levels of sum PBDE were for outdoor air 0.02 pg/m 3 for comparison and in indoor air PBDE was not detected.

PBEB was analyzed over time in lake trout in Lake Ontario, Canada. Concentrations ranged from 17 ± 3 to 320 ± 156 ng /g lipid (Ismail et al., 2009). In a new investigation in 2009 where “new” brominated flame retardants were analysed in artctic animals (sea birds,

60 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) capelin, ice bear, ring seal and polar fox) PBEB was not detected in any samples (KLIF unpublished data). This is the same results as in this study where levels of PBEB was below 0.1 ng/g w.w in investigated fish liver, crab, and mussel, indicating low levels in biological material.

Toxicological effects No studies were found on ecotoxicological effects. Compared to an adapted PNEC water of 530 ng/l and PNEC sed of 1.55 mg/kg d.w (se Table 31) the levels found in this study are much lower.

Concluding remark It is worth noting that PBEB was known to be produced in the 1970’s and 1980’s so the fact that the compound is only detected in relation to a metal recycling facility might be connected to the treatment of “old” waste, including electronical waste, produced in the seventies and eighties.

The compound is detected in seepage water from Helik Teigen and in sediment in the river (Loselva) receiving seepage water from Hellik Teigen. The compound is not detected in biological material. PBEB is only detected at one locality and therefore probably not widely distributed. This together with low concentrations compared to adapted PNEC values PBEB is assessed to be of little to moderate environmental concern.

Pentabromotoluene - PBT Results from this study PBT was detected in relatively low levels in seepage water and in sediment from Hokksund area (Hellik Teigen) and Lindum (Drammen area). In seepage water from Hellik Teigen, Hokksund (car demolishing site) the level was about 5 ng/l and in seepage water from Lindum PBT was detected in one of three samples (~0.4 ng/l). The concentrations in sediment were 0.03 and 0.2 ng/g from Hokksund and Lindum, respectively. PBT was also detected in a few other sediment and waste water samples, but in levels below LOQ.

Table 36 Positive detections of PBT. Location Water ng/l Sediment ng/g d.w Lillehammer RA o n.d. Solumstrand RA o n.d Sandnessundet (Tromsø) n.a. o Lindum Ressurs og Gjenvinning o 0.1 – 0.3 Hellik Teigen 4.4 – 7.5 0.02 – 0.04 LOQ 0.22 0.02 LOD 0.06 0.01 n.a: not analysed, n.d: not detected, o: detected in single replicate and/or very close to detection limit

Comparison to previous studies PBT has been identified in sewage sludge from Swedish STPs (Mattson et al., 1975), in eggs of Glaucous gulls in the Norwegian arctic (

61 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

In the present study PBT was not detected in sewage sludge (<0.09 ng/g d.w) or in biotic samples (<0.1 ng/g w.w). In sediment it was detected in concentrations ranging from 0.02 – 0.04 ng/g d.w in river sediment (Hellik Teigen) which is much lower than the reported concentrations of PBT in Elbe river and its tributaries ranging from <1 – 25 ng/g d.w (Schwarzbauer et al., 2001).

Toxicological effects In a study by Simonesen et al. (2000) the LC 50 for fish was > 5 mg/l (48 hours). Compared to an adapted PNEC water of 530 ng/l and PNEC sed of 1.55 mg/kg d.w (see Table 31) the levels found in this study are much lower.

Concluding remark PBT can be considered to be widely distributed due to the fact that it is detected in northern Norway and in south eastern Norway at several sites. PBT is also detectected in marine receiving waters in Sandnessundet. However, the concentrations seem to be low compared to adapted PNEC values and other studies. The compound is not detected in biological material in this investigation. Overall it is assessed that PBT is of little to moderate environmental concern.

PBT is an environmental transformation product of TBBPA (Arbeli et al., 2006) which again is a elimination product of TBBPA - DBPE. It is likely also a degradation product of many other BFRs, such as DBDPE (in which just the ethyl-ethyl bond would have to be cleaved). Interestingily DPDPE is detected in most of the places where PBT is detected and in general DPDPE is detected in many samples in this investigation.

Hexabromobenzene - HBB Results from this study HBB was detecteted in the highest number of samples in this investigation (except for BDE 209). The compound was detected in seepage water and in sediment from Hokksund (15 ng/l, 0.08 ng/g d.w) and Lindum (1.8 ng/l, 1.3 ng/g d.w), in in- and outgoing water from sewage treatment plants, in sewage sludge, in outdoor air and in moss- and needle samples, see Table 37 for details.

Table 37 Positive detections of HBB. Location Water ng/l Sediment Sludge ng/g Air ng/m3 Moss ng/g Pine ng/g d.w d.w d.w ng/g w.w Sandnessundet n.a o n.a n.a n.a n.a (Tromsø) Mjøsa n.a o n.a n.a n.a n.a Hellik Teigen 11.4 – 19.1 0.1 n.a n.a n.a n.a Lindum 0.6 – 2.8 0.4 – 1.8 n.a n.a n.a n.a Langenes RA 0.1 – 2.8 n.a 0.27 – 0.42 n.a n.a n.a Solumstrand RA 0.7 – 1.9 n.a 0.22 – 0.6 n.a n.a n.a Lillehammer RA 0.2 – 1.1 n.a 0.12 – 0.14 n.a n.a n.a Drammen outdoor n.a n.a n.a o n.a n.a Hurum Energigj. n.a n.a n.a n.a o o LOQ 0.12 0.02 0.4 0.002 0.2 0.04 LOD 0.03 0.01 0.1 0.001 0.06 0.01 n.a: not analysed, o: detected in single replicate and/or very close to detection limit

Comparison to previous studies

62 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

There are many findings of HBB in different environmental compartments in the literature.

HBB has been detected in sediments collected at Osaka and in low concentrations (0.003 – 0.004) ug/g d.w in sediment collected at Tokyo. Sediment concentrations in this study are in general lower than this.

HBB has been found in eggs and plasma of glaucous gulls in the Norwegian Arctic (Verreault et al., 2007). HBB was analyzed in pooled Herring gull egg samples from the Great Lakes of North America in 2004. Although present at much lower levels than the PBDEs (0.24 – 0.53 ng/g wet wt), HBB was generally the most abundant of the non-PBDE BFRs. (Gauthier et al., 2007). An update on the study of BFRs in the Great Lakes found HBB levels ranging in herring Gull eggs from 0.27 – 0.66 ng/g w.w (Gauthier et al., 2009). In a recent study HBB was not detected in sea birds, seal, fish, ice bear an arctic fox in arctic (KLIF unpublished data). Limit of detection in this study varied from 1.6 pg/g w.w to 3.8 pg/g w.w. The study by KLIF is in line with this investigation where HBB was not detected in fish (nor in crab or mussel). HBB is detected in moss and pine but there are no reported investigations in the cited literature where moss and pine are included.

In a study that measured total air levels (particle + gas) of HBB in Egbert, Ontario (a suburb of Toronto) were 0.02 – 0.09 pg / m 3 (Gouteux et al., 2008). These numbers are actually lower than the findings in this study where HBB was measured in concentrations ranging from 1.3 – 10.2 pg/m 3 (particle phase).

HBB has even been detected in human serum samples in concentrations ranging from 0.11 – 1.5 ng/g lipid, with a median of 0.27 ng/g (Zhu et al., 2009).

Toxicological effects Most studies on toxicology in the cited literature are experiments with mice and are not considered very relevant here. In a six days mortality test on the copepod Nitocra spinipes NOEC for HBB was determined to 33.4 mg/l (Breitholtz et al. 2008), much higher than the water concentrations quantified in this study..

Concluding remarks HBB is relatively widely distributed in this study, meaning it is detected in many samples and different environmental compartments. The highest concentrations are found in a car demolishing facility and waste disposal site which probably reflect handling of waste and scrap. In a screening for halogenated compounds in samples from an aluminum recycling plant, handling waste from electronics and electronics plastics and a car shredder, HBB was observed in all scrap samples (Sinkkonen et al., 2004).

Model prediction concludes that HBB mainly binds to sediment and soil and in a very little degree to air and water (Kawamoto and Kuramochi, 2007, Tittlemier et al., 2002). In this study HBB is detected in water in several samples from different locations and in air which is a little bit surprising based on model predictions. Also, the air concentrations in this study are higher than that reported for Ontario, a much more densely populated area than Drammen (Gouteux et al., 2008). This may indicate relatively high local emission from local sources. Compared to a NOEC for HBB of 33.4 mg/l (Breitholtz et al. 2008) or to an adapted PNEC water of 530 ng/l (see Table 31) the water levels found in this study are much lower.

63 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

In this study HBB is detected in sediments from receiving fresh (Mjøsa) and marine (water) indicating that HBB persist to some degree in the environment. This may be interpreted in line with model predictions that HBB binds to a high degree to sediments. Compared to an adapted PNEC sed of 1.55 mg/kg d.w the concentrations in this study are low.

Bioaccumulation studies shows varies results from a high bioaccumulation to none (Sijm et al., 1993;Sijm et al., 1995; Oliver and Niimi, 1985; Zitko and Hutzinger, 1976; Zitko, 1977, Nyholm et al., 2009). In this study HBB was not detected in fish, mussel or crab and in a recent study by KLIF (unpublished data) HBB was not detected in marine birds, mammals and fish, suggesting low bioaccumulation of HBB.

HBB is detected in many samples, in different environmental compartments including receiving waters and in moss and pine needles. HBB is not detected in fish, mussel or crab which may had been more worrying than the detections in moss and pine neddles which are more considered to act as passive samplers. Overall HBB is assessed of being of moderate environmental concern.

1,2 bis(2,4,6-tribromophenoxy)ethane - BTBPE Results from this study BTBPE was detected in sediment (average 4 ng/g and 1.3 ng/g from Lindum and Hokksund (Hellik teigen), respectively and in one sample from Drammensfjorden 2.7 ng/g).

BTBPE was also detected seepage water (average 3 ng/l, 76 ng/l and 0.85 ng/l from Lindum, Hokksund and Lilehammer RA, respectively), and in sludge (2.1 ng/g in one sample from Solumstrand, Drammen and detected but

Table 38 Positive detections of BTBPE. Location Water ng/l Sediment ng/g d.w Sludge ng/g d.w Hellik teigen 59.1 – 107 0.6 – 1.7 n.a Lindum 2.1 – 4.2 3.1 - 4.5 n.a Drammensfjorden n.a o n.a Langenes RA n.d n.a o Solumstrand RA n.d n.a o Lillehammer RA o n.a o LOQ 1.30 0.32 1.6 LOD 0.38 0.1 0.5 n.d: not detected, n.a: not analysed, o: detected in single replicate and/or very close to detection limit.

Comparison to previous studies In Lake Ontario sediment cores was investigated and BTBPE was found in the surface sediment, with average concentration of 6.7 ng/g d.w (Qiu et al., 2007). In diverse samples from the Pearl River Delta, levels in sediment ranged from 0.3 – 22 ng/g dw, in sewage sludge from 0.3 – 1.7 ng/g d.w, in farmland soil from 0.02 – 0.11 ng/g d.w (Shi et al., 2009). These numbers are comparable to the findings in sediment and sludge in this study.

In water a median concentration of 1.96 pg/l (dissolved phase) was reported by Law et al in an study of trophic levels (Law et al., 2006; Law et al., 2007) in Lake Winnipeg (Canada), This is much lower than LOD and LOQ in this study and much lower than the findings in

64 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) water. This is not surprising considering that the Canadian study is reporting the dissolved phase.

There are many findings of BTBPE in biota. Previous studies have detected BTBPE in biotic samples, although often at low concentrations. In a Canadian study BTBPE was found in mussels (mean concentration of 1.29 ng/g lipid wt), and in northern fulmar eggs from the Faroe Islands (North Atlantic) the mean concentration was 0.11 ng/g lipid wt. BTBPE have also been found in eggs and plasma from glaucous gulls in the Norwegian arctic in low concentrations (max 0.96 ng/g lipid wt) were found in egg yolk and in only one plasma sample. (Verreault et al., 2007). For a 1979-2004 time series of lake trout in the Great Lakes, the concentration increased from 0.6 ± 0.3 ng/g lipid to 1979 to 2.6 ± 0.6 ng/g lipid in 1993 (with a doubling time of six years), and since then levels declined about 40% (Ismail et al., 2009), these levels were on average greater than in Lake Winnipeg mentioned above. In a recent study by KLIF (unpublished data) BTBPE was found in some samples of guillemot (Uria lomvia ) in concentration of 7.29 and 11.25 ng/g w.w. The compound was not detected in any others birds investigated, nor in fish (capelin), seal, arctic fox or ice bear.

BTBPE was not detected in fish, mussel or crab samples in the present study (<0.5 ng/g ww or <4.8 ng/g lipid wt), however, the detection limit were above several of the previously measured levels. BTBPE have also been reported in the particulate phase in air near the Great Lakes with median concentrations from 0.5 to 1.2 pg/m 3 (Venier and Hites, 2008). No BTBPE in air in the present study but also in this case the detection limit was higher (6 pg/m 3) than previously reported concentrations.

Toxicological effects As discussed previously there are only data on a few toxicological studies available for BTBPE in the literature. From existing studies (most of them with rat) it is not possible to determine a PNEC for BTBPE. The PNEC for pentaBDE will be used for comparison of observed environmental concentrations to predicted “safe” environmental concentrations. The PNEC water and PNEC sediment for pentaBDE are 0.53 µg/l and 1.33 µg/g, respectively. The highest measured concentration in seepage water (107 ng/l) is approximately 5 times lower than the adapted PNEC water .

Concluding remarks BTBPE was detected in sewage sludge, seepage water and sediment in the present study. In many other studies BTBPE has been detected in biota such as mussel, fish and bird eggs, and BTBPE have shown high biomagnification potential in juvenile rainbow trout (Tomy et al. 2007).

BTBPE was not detected in biota (<0.5 ng/g w.w) in the present study, and the levels are low compared to reported levels for sum of PBDEs (3 – 226 ng/g w.w) in cod liver (KLIF 2008). However, if the usage of BTBPE is increasing as a result of the ban of PBDEs, increased levels in biota can be expected.

In the present study BTBPE was detected in sediment in the river Loselva receiving seepage water from Hellik Teigen. It was also detected in sludge and other water samples. It is detected in biological samples (KLIF 2009 unpublished data among others) and the water concentration (107 ng/l) found in this study is relatively high. Overall BTBPE is therefore assessed to be of moderate environmental concern to environmental concern.

65 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Decabromodiphenylethane - DBDPE

Results from this study DBDPE was detected in seepage water from Hokksund (average 80 ng/l) and in all three sediment samples from Hokksund but only above the LOQ in one sample (1.8 ng/g d.w). DBDPE was also detected in sewage sludge (8 of 9 samples), and above LOQ in 5 samples (>3.3 ng/g d.w) and found in two incoming waster samples from Solumstrand, Drammen, and in two needle samples (0.1 ng/g w.w) from Hurum.

Table 39 Positive detections of DBDPE. Location Water ng/l Sediment ng/g d.w Sludge ng/g d.w Pine needles ng/g ww Hellik teigen 15.3 – 186 o n.a n.a Tromso RA n.d n.a o n.a Solumstrand RA o n.a 3.6 – 8.7 n.a Lillehammer RA n.d n.a 1.3 – 5.0 n.a Hurum n.a n.a n.a o LOQ 4.9 1.79 3.3 0.3 LOD 1.4 0.54 1.0 0.1 n.d: not detected, n.a: not analysed, o: detected in single replicate and/or very close to detection limit.

Comparison to previous studies Ricklund et al., 2008 performed a worldwide survey of sludge from waste water treatment plants and found levels ranging from < dl to 160 ng/g d.w. These findings are comparable to the results of this study.

DBDPE have also been found primarily in the particulate phase in air near the Great Lakes (U.S.) at median concentrations from 1 to 22 pg/m 3(Venier and Hites, 2008) and have been found in tree bark from the Northeastern U.S. ranging from

DBDPE has previously been detected in fish in concentrations about 1 ng/g lipid wt (Lake Winnipeg, Canada, Law et al., 2006; Law et al., 2007). Another study on water birds from e- waste region in the Pearl River Delta reported median DBDPE concentrations of 10 – 176 ng / g in lipids, with the total range from n.d. – 900 ng/g (Luo et al., 2009). In Northern China, in the Yellow River Delta (an important industrial development area, as well as an important breeding ground for birds), median concentrations in various bird egg samples range from n.d. – 1.7 ng/g lipid, and totally from n.d. – 2.2 ng / g overall (Gao et al., 2009). In the present study, no DBDPE was detected in fish, mussel or crab, however, the detection limit (0.66 ng/g w.w or 6.6 ng/g lipid wt) were above some of the previously reported levels. In a recent study by KLIF (unpublished data) DBDPE was found in one sample of guillemot (Uria lomvia ) in concentration of 5.81 ng/g w.w. The compound was not detected in any others birds investigated, nor in fish (capelin), seal (ring seal), arctic fox or ice bear.

Toxicological effects Most studies on toxicology in the cited literature are experiments with mice and are not considered very relevant here. There is one study showing that levels of 19 µg/l were acutely

66 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) toxic to water fleas, with a 48 hour EC50 response (Nakari and Huhtala, 2009). Applying a factor of 100, this will give a PNEC comparable to the PNEC for decaBDE (> 0.2 µg/l). The highest measured concentration in seepage water (186 ng/l) is close to the PNEC water (200 ng/l).

Concluding remarks DBDPE was detected in sediment in the river Loselva receiving seepage water from Hellik Teigen, and in pine needles from Hurum. It is also found in relatively high concentrations in seepage water compared to a PNEC water (200 ng/l). Overall DBDPE is assessed to be of moderate environmental concern to environmental concern.

2,4,6 tribromophenol - TBP Results from this study TBP was found in highest concentration in crab from Tromsö (42.1-131 ng/g w.w) and in one fish liver sample from Sandnessundet (Tromsø, 55.8 ng/g w.w). The compound was also detected in crab and mussel from Drammensfjorden. TBP was also detected in two sediment samples from Tromsø and Mjøsa.

Table 40 Positive detections of TBP. Location Sediment ng/g Fish liver ng/g Blue mussel ng/g Crab ng/g d.w w.w. w.w. w.w. Sandnessundet (Tromsø) o o n.d 42.1 - 131 Mjøsa o n.d n.a n.a Drammensfjorden n.d n.d o 3.0 – 8.2 LOQ 0.24 1.3 1.3 1.3 LOD 0.07 0.4 0.4 0.4 n.d: not detected, n.a: not analysed, o: detected in single replicate and/or very close to detection limit.

Comparison to previous studies TBP has been detected throughout the environment, e.g. in fish, sediment, water, sewage sludge and indoor air. Bromophenols are used as industrial chemicals and they are also naturally produced by marine organisms. One suggested function of naturally produced bromophenols is chemical protection against predators (Hassenkloever et al., 2006).

Previously sediment investigations have reported from no detections of the compound (German Bight, Reineke et al., 2006) to 3690 ng/g d.w (Rohne estuary, Tolosa et al., 1991). The reported values thus vary a lot and in general the detections of TBP in sediments from this investigation (< 0.24 ng/g d.w) seem low.

Bromophenols are, as mentioned, naturally produced by marine organisms, and it has been detected in many marine organisms such as brown algae (up to 5780 ng/g), bryozoa, sponges, hydroids and prawns (Chung et al. 2003, Whitfield et al., 1992). Also, ten different species of fish, collected in August 1992 from the eastern coast of Australia, contained TBP at concentrations of <0.05 to 3.4 ng/g for the carcass and <0.05 to 170 ng/g for the whole gut (analysis of single fish from each species) (Whitfield et al., 1995). Compared to this study the concentration of 55.8 found in fish liver from Sandnessundet (Tromsø) is higher in that found for the whole carcass but lower than that found whole gut in the study by (Whitfield et al., 1995). However, in this study the liver was measured which may contain higher levels than other organs in the fish. Bromophenols is also found in crustaceans and mollusks from the Pacific Ocean (Boyle et al., 1992).

67 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Toxicological effects In six days toxicity test on larvae development of the copepod Nitocra spinipes the NOEC for TBP was determined to 300 µg/l (Breitholtz et al. 2008). For tribromophenol the LC 50 (in fish) was 6.5-6.8 mg/l (96 hours, fathead minnow) and 1.1 mg/l (96 hours, fathead minnow, flow through bioassay) (Simonsen et al., 2000).

Concluding remarks Bromophenol is produced naturally by many marine organisms and as such it is difficult to assess the risk posed by the compound, and what the source of the TBP is. TBP can be a degradation product of other BFRs (TBBP-A), and is also produced naturally by many marine organisms (e.g. algae, sponges) and can enter the food chain in this manner. Interestingly no TBP was found in samples with high concentration of other BFRs (e.g. seepage water from Hellik Teigen) which may indicate natural origin of TBP in the detected samples. The compound was not detected in water and only in concentration well below comparable PNEC values in sediment. But, studies suggest that the potential for bioaccumulation is moderate to high (Simonesen et al 2000) and because of the fact that the compound is detected in several biological samples in this study, it is assessed as being of environmental concern.

Tetrabromobisphenol A dialyll ether - TBBPA-AE Results from this study TBBPA-AE was detected in sediment from Hellik Teigen (Hokksund), Lindum, and Drammensfjorden and in water samples from Hellik Teigen, Langenes RA (Tromsø) and Solumstrand RA (Drammen area).

Table 41 Positive detections of TBBPA-AE. Location Water ng/l Sediment ng/g d.w Hellik Teigen o 0.3 – 0.5 Lindum n.d 2.4 Drammensfjorden n.a 0.5 – 1.3 Solumstrand RA o n.a Tromso, RA o n.a LOQ 0.27 0.04 LOD 0.08 0.01 n.d: not detected, n.a: not analysed, o: detected in single replicate and/or very close to detection limit.

Comparison to previous studies To our best knowledge, no environmental levels of TBBPA-AE are published in the peer reviewed literature.

Toxicological effects As discussed previously the available information on toxicity of TBBPA-AE is scarce. From existing toxicity studies it is not possible to determine a PNEC for TBBPA-AE. As discussed above, the PNEC for TBBPA will be used for comparison of observed environmental concentrations to predicted “safe” environmental concentrations. The PNEC water for TBBPA is 0.25 µg/l and the PNEC sediment is 54 ng/g ww. The highest measured concentration in sediment (1.3 ng/g) is lower than the PNEC sediment (54 ng/g w.w).

68 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Environmental concern TBBPA-AE was detected in sediment in the river Loselva receiving seepage water from Hellik Teigen (Hokksund) and in sediments from Drammensfjorden. TBBPA-AE is therefore assessed to be of moderate environmental concern .

Ethylene bis(tetrabromophtalimide) - BTBPI Results from this study BTBPI was detected in one replica of seepage water from Hellik Teigen (35 ng/l). Note that since the recovery of BTBPI in spiked water (as can be seen in Figure 12) were very low there are great uncertainties related to the quantified concentrations in water in the present study. Analytical problems were experienced analysing BTBPI in several matrices because of interfering peaks and high background. In contrast to other BFRs analysed with HPLC- MS/MS no mass transition were found and the same fragments was monitored as parent and daughter ions (see Table 30), causing these interferences. To verify the detection of BTBPI in seepage water an LTQ OrbiTrap XL was used. This instrument has higher resolution which improves the signal-to-noise. Analysing all samples on a high resolution instrument would be preferable, however, these analyses have not yet been performed.

Comparison to previous studies No environmental levels have to our knowledge been reported for BTBPI.

Toxicological effects Most toxicological studies in the cited literature are studies on rat or rabbit and considered little relevant for the findings here. Compared to an adapted PNEC water of > 200 ng/l (see Table 31) the level here is lower.

Concluding remarks BTBPI is only detected in one sample. The compound is, however, found in receiving water (Loselva) so overall the compound is assessed of being of little to moderate environmental concern.

Decabromodiphenyl ether - BDE 209 BDE 209 was detected in all analysed samples except from indoor air samples, and a few mussel, crab, and fish samples. Positive identifications of BDE 209 can be seen in Table 45. In almost all samples the concentration of BDE 209 was highest among the analysed BFRs, with the exemption of mussel, crab and some fish samples where levels of TBP were highest. Especially sewage sludge samples were heavily dominated by BDE 209; in average the levels were 70 times higher than for DBDPE, 200 times higher than for BTBPE, and three orders of magnitude higher than HBB.

The levels of BDE 209 in the present study are in line with previously reported concentrations in Norway (KLIF 2008). Despite the ban of BDE 209 declining environmental levels cannot yet be observed. BDE 209 is still present in consumer products and a fast decline of this persistent BFR should not be expected.

69 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 42 Positive detections of BDE 209. Location Water Sediment Sludge Air Moss ng/g Pine Fish, ng/l ng/g d.w ng/g d.w pg/m 3 d.w ng/g w.w crab, mussel ng/g w.w Sandnessundet n.a. 0.1 – 0.2 n.a n.a n.a n.a 0.1 - 15 (Tromsø) Mjøsa n.a. 2.2 – 8.9 n.a n.a n.a n.a 0.1 – 2.4 Drammensfjorden n.a. 0.5 - 64 n.a. n.a. n.a. n.a. 0.1 - 43 Bergen n.a. n.a. n.a. n.a. n.a. n.a. 0.2 – 1.7 Losna n.a. 0.1 – 0.3 n.a. n.a. n.a. n.a. n.a. Hellik Teigen 710 - 7.1 – 9.5 n.a n.a n.a n.a n.a. 1840 Lindum 19 - 46 28 - 149 n.a n.a n.a n.a n.a. Langenes RA 3.5 - 118 n.a 47 - 97 n.a n.a n.a n.a. Solumstrand RA 5.7 - 106 n.a 250 - 346 n.a n.a n.a n.a. Lillehammer RA 0.4 - 43 n.a 189 - 260 n.a n.a n.a n.a. Drammen outdoor n.a n.a n.a 8 – 37 n.a n.a n.a. Hurum Energigj. n.a n.a n.a n.a 1.1 – 14 0.2 – 0.5 n.a. LOQ 0.24 0.03 0.3 4.0 2.0 0.1 0.3 LOD 0.07 0.01 0.1 1.0 0.6 0.03 0.1 n.d: not detected, n.a: not analysed, o: detected in single replicate and/or very close to detection limit.

Overall discussion BFRs The phase-out of PBDEs has put attention to other “new” BFRs that are or can be used as replacements. Although called new some of these BFRs have already been in use for several decades, but as a result of the ban of PBDEs there is reason to assume that the usage of these BFRs will increase. In this study 14 BFRs were screened in the Norwegian environment and 9 of these were detected in one or more samples.

Most detections of new BFRs were made in seepage water (dissolved and particle phase combined) from a metal recycling and car demolishing facility in Hokksund. The new BFRs detected were: TBBPA-DBPE, PBT, TBT, PBEB, HBB, BTBPE, DBDPE, TBBPA-AE and BTBPI. It is well-known that so called e-waste contains BFRs, and in the seepage water also high concentrations of the newly banned BFR BDE 209 were observed (~ 1 ug/l). The concentration of BDE 209 was approximately ten times higher than the highest concentration of a new BFR. This is probably because this facility handles end-of-life products and thus the pattern represents past and not present production of BFRs. It can be speculated that the concentration of new BFRs will increase in the future in environmental samples taken close to e-waste deposits when products containing new BFRs is delivered and handlet at these deposits.

Several new BFRs were detected in sediment from the river Loselva receiving seepage water from metal recycling and car demolishing facility in Hokksund (PBT, PBEB, HBB, BTBPE, DBDPE and TBBPA-AE). TBBPA-DBPE was not found in sediment, possibly due to its lower persistence. TBBPA-DBPE (and TBPA) has been shown to be susceptible to hydrolysis. BFRs generally have low water solubility and higher levels in sediment and sludge as compared to water can be expected. Contradictory, most detections were made in water samples. There are several possible explanations for this: the water samples were taken closer to a source of BFRs, both dissolved and particle phase were analysed, and large

70 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) volumes of water were analysed (e.g. 7 litres of waste water compared to 20 g d.w of sediment or 5 g d.w of sewage sludge).

The only BFR (besides from BDE 209) that was detected in mussel, fish, or crab was TBP. TBP was on the other side not detected in the most polluted samples (seepage water, waste water, sewage sludge). This can probably not be explained by a potentially high bioaccumulation potential as compounds with a high bioaccumulation potential are often measured at highest concentration at high trophy level in the food chain. For example, the concentrations of sum of PBDEs are generally higher in carnivorous fish than in mussel and crab. However, for TBP the opposite pattern was observed; the concentrations were generally highest in crab and mussel, and lower in fish liver. The reason for this is unclear but it may indicate that the TBT is of natural origin, as it is naturally produced by several algae and other marine organisms and can enter the food chain in this manner. Alternatively, the TBP may be metabolization products of TBBP-A and derivatives (Arbeli, 2006).

HBB was the new BFR found in most matrices from many localities and thus assumed to be widespread in the environment. DBDPE was also found in many matrices, and as the detection limit was higher for DBDPE, it is not possible to state that this compound is less widespread than HBB. DBDPE was also found in higher concentration than HBB in several matrices: e.g. sewage sludge, seepage water and pine needles. BTBPE was detected in sewage sludge, seepage water and in sediment. As for DBDPE the detection limit for BTBPE is higher than for HBB and it is not possible to state that BTBPE is less spread. Additionally, BTBPE has, in contrast to HBB and DBDPE, shown high biomagnification potential in fish which makes findings of BTBPE in the environment a concern.

TBPH, EHTBB, ATE, DPTE, and TBPA were not detected in any sample in the present study, and these compounds are therefore assessed to be of little environmental concern. However, more studies on these compounds are needed to confirm the low environmental levels in Norway. Some of the compounds considered in the present study were analysed in environmental samples for the first time and the uncertainties related to these results are larger than for compounds that has been analysed before, participated in inter-laboratory calibration, or for which labeled internal standards are available. As work progress and more studies on these compounds are performed the uncertainties in the analysis will be smaller, i.e. the results from a first screening study should not uncritically be used to study time trends.

In general, the toxicological information available of the different new BFRs is scarce, and consequently it is difficult to draw any conclusion of the environmental concern of these BFRs. Most of the available toxicological information concerns the effects of single PBDEs, and as shown in e.g. Breitholtz et al. (2008) mixtures of BFRs at their individual NOEC can cause ecotoxicological effects. In the present assessment there were no measured levels exceeding the adapted PNECs for new BFRs (derived for PBDEs and TBBPA). However, there are large uncertainties in the PNECs, and in a few cases also in measured water concentrations. The presence of new BFRs in the environment warrants new ecotoxicological studies of these compounds, individually and in mixtures, for proper environmental risk assessments of them.

Despite the ban of BDE 209 declining environmental levels cannot yet be observed. BDE 209 is still present in consumer products and a fast decline of this persistent BFR should not be

71 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) expected. In almost all samples the concentration of BDE 209 was highest among the analysed BFRs, with the exemption for mussel, crab and some fish samples where levels of TBP were highest. Especially sewage sludge samples were heavily dominated by BDE 209; in average the levels were 70 times higher than for DBDPE, 200 times higher than for BTBPE, and three orders of magnitude higher than HBB.

Based on the results no clear differences in samples taken from the different waste water treatment plants (using different cleaning technologies) can be identified.

72 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

6.2 Polyfluorinated compounds

In Table 43, an overview of positive detections of the studied PFCs is presented. It demonstrates that most substances were found in soil and sediment samples, whereas PFTrDA was the only substance quantified in sea mussel. Below, the results for each substance included in the report are discussed more thoroughly.

Table 43 Overview of positive detections of the investigated PFCs. +: detected, O: detected in single replicate and/or very close to detection limit, -: not detected. Compound Soil Sediment Water Blue mussel Crab Fish liver 6:2 FTS + + + NQ NQ NQ PFBS + o + - - - PFHxS + + + - + + PFOS + + + - + + PFDS + + - - + + PFPeA + + + - - - PFHxA + + + - - o PFHpA + + + - - - PFOA + + + - o o PFNA o o + - o + PFDA NQ NQ + - - + PFUnDA + + + - + + PFDoDA + + - - o + PFTrDA + + - + + + PFTeDA - - - - - o NQ=Not quantified

Perfluorobutane sulfonate - PFBS Results from this study In soil from the fire fighting training ground at Flesland airport, PFBS was found at distances up to 50 m from the center (fire fighting training ground) with concentrations ranging between 0.28-1.8 ng/g d.w, with the highest concentration 20 m from the training ground. Water samples from Flesland airport had relatively high concentrations average 97 ng/l, but only one sediment sample from Langavatn near Flesland contained PFBS at a concentration near the detection limit. Water from Res-Q had average concentration of 25 ng/l and sediment from Res-Q-5 contained average 0.58 ng/g. PFBS was not found above the detection limit in any of the biota samples.

Table 44 Positive detections of PFBS. Location Water ng/l Sediment ng/g d.w Soil ng/g d.w Flesland airport 68 - 148 n.a. 0.28 – 1.8 Langavatn n.a 0.17 n.a Res-Q (Haugesund) 11 - 35 0.54 – 0.64 n.a n.a: not analysed, o: detected in single replicate and/or very close to detection limit.

73 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Comparison to previous studies In general, low concentrations of PFBS have been discovered in earlier screening projects (e.g. Bakke, 2007; Fjeld, 2004; Woldegiorgis et al, 2006), and the concentrations measured in this study are in the same range as those reported earlier from around Scandinavia. The water concentration from both Res-Q and Flesland airport was higher than those reported from sea water in other Nordic countries (Kallenborn, 2004). In Spain PFBS was detected in 11 municipal drinking water samples with the highest concentration at 69 ng/l (Ericson et al., 2009). The concentrations measured at Flesland airport were in the same range as one replicate of ground water sampled at Gardemoen airport, which is also affected by AAAFs (Amundsen, 2009).

Toxicological effects There has not been found PNEC/NOEC values for water or sediment for PFBS. In acute dietary studies on juvenile birds (mallards and northern bobwhite quail) the no observed adverse effect concentrations (NOAEC) were 5620 and 3160 mg PFBS/kg w.w feed (Newsted et al., 2008). The same study reported a NOAEC of 900 mg/kg w.w. feed for quail reproduction. It has also been demonstrated that the bioaccumulation of PFBS is lower than for longer chained homologues (Renner, 2006).

Concluding remarks In agreement with the low bioaccumulation, PFBS was not detected in the biological samples. PFBS is found in the receiving environment and is assessed as being of moderate environmental concern . But, regarding its effect on the local environment, little information is available, and because the compound is found in water samples at the same levels and higher than that reported elsewhere, and it is found in other investigations around Scandinavia PFBS should maybe be investigated further.

Perfluorohexane sulfonate - PFHxS Results from this study In soil from Flesland fire fighting practice ground, PFHxS was found in all samples, ranging from 0.12-21 ng/g (d.w), with the highest concentration at 20 m from the platform. The water concentration near Flesland airport averaged 401 ng/l and PFHxS was quantified in all sediment samples taken in Langavatn, average concentrations ranging from 0.7-2.6 ng/g d.w. Trace levels were found in crab near the airport while elevated levels were found in fish liver Langavatn, the receiving water of seepage water from the fire fighting training ground (average concentration 160 ng/g).

Water samples from Res-Q also contained PFHxS but at lower concentrations compared to Flesland, average 42 ng/l. Sediment concentrations from sampling station Res-Q-5 averaged 6.0 ng/g d.w. Sediment from Res-Q-4 was close to the limit of detection. PFHxS was not found above the detection limit in biota taken near Res-Q or in the sediments in the receiving water.

74 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 45 Positive detections of PFHxS Location Water ng/l Sediment ng/g Soil ng/g Fish liver Crab ng/g d.w d.w ng/g w.w w.w Flesland airport 319 – 471 n.a 0.12 - 21 n.a n.a Langavatn n.a 0.7 - 2.6 n.a 118 – 268 n.a Flesland brygge n.a n.a n.a n.a 0.12 – 0.46 Res-Q (Haugesund) 33 – 48 0.2 – 6.7 n.a n.d n.d n.d: not detected, n.a: not analysed.

Comparison to previous studies The water concentrations measured at both Res-Q and Flesland airport was in general higher than those reported from other screening studies around the Nordic countries (Kallenborn et al, 2004). In comparison with other AAAF-sites, the concentration was for example lower than reported from Gardermoen groundwater (705-5952 ng/l; Amundsen, 2009).

Concentrations of PFHxS measured in sediment in earlier studies are generally lower, ranging between

Toxicological effects No relevant information regarding toxological effect or no effect concentrations (PNEC, NOEC) has been found.

Concluding remarks Little information is available on the environmental effect of PFHxS. However, relatively high concentrations were found in sediment and water samples, and also in fish liver samples from Langavatn, receiving water of seepage water from the fire fighting training ground at Flesland airport. Also in samples of crabs in the marine receiving waters (Flesland brygge) was PFHxS detected. PFHxS has been detected in sediments from the training ground at Res- Q but not in the receiving waters. Overall PFHxS is assessed as being of environmental concern.

Tetrahydroperfluorooctane sulfonate - 6:2 FTS (THPFOS) Results from this study The concentration of 6:2 FTS measured in the soil from the Flesland airport fire fighting training ground can be seen in Figure 15. The concentration ranged between 0.84-2101 ng/g (d.w), with the highest concentration appearing at 10 m distance from the training platform. Elevated levels of both 6:2 FTS and PFOS and possible other interfering compounds in soil samples 0-30 m resulted in higher uncertainty in the results due to suppression of the internal standard. The water concentration near Flesland was elevated, ranging between 5110-6693 ng/l and 6:2 FTS was quantified in all sediment samples taken in Langavatn, giving average concentrations of 9.1, 7.0 and 6.7 ng/g (d.w) at the three sampling locations, respectively. 6:2 FTS was not detected in mussel or crab in Bleivika but was found in fish liver in Langavatn near Flesland airport. The concentrations found (37-138 ng/g) are however not quality assured since signal enhancement of 6:2 FTS, but not for the internal standard PFOS, occurred in biota samples.

75 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Water samples from Res-Q were difficult to analyze due to unknown component(s) (possibly detergents as the water was “foaming”). The water concentrations in two samples were 607 and 7847 ng/l. The 6:2 FTS concentration in sediment samples Res-Q-5 were higher compared to Flesland sediments, with average concentration of 345 ng/g d.w. Res-Q-4 sediment averaged 17 ng/g d.w while in Bleivika only trace levels of 6:2 FTS were found. No 6:2 FTS was detected in the biological samples taken near Res-Q.

Table 46 Positive detections of 6:2 FTS. Location Water ng/l Sediment ng/g Soil ng/g Fish liver ng/g Crab ng/g d.w d.w w.w w.w Flesland airport 5110 - 6693 n.a 0.84 - 2101 n.a n.a Langavatn n.a 1.5 – 11.8 n.a 37 – 138* n.a Flesland brygge n.a n.a n.a n.d n.d Res-Q (Haugesund) 607 – 7847 * 6.5 - 379 n.a n.a n.a Bleivika (Haugesund) n.a o n.a n.d n.d n.d: not detected, n.a: not analysed, o: detected in single replicate and/or very close to detection limit. * Higher degree of uncertainty due to interference.

0 10 20 30 40 50 6:2 FTS 75 ∑PFSAs 100 150 Distance from training ground (m) 200

0 500 1000 1500 2000 2500 ng/g soil d.w.

Figure 15 Concentration of 6:2 FTS and ∑PFSAs in the soil at the fire fighting practice ground at Fleslands Airport (ng/g d.w).

Comparison to previous studies Concentrations of 6:2 FTS in soil from the present sites are in the same range as those found at both Gardermoen airport (

76 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) study. The water concentrations from both Res-Q and Flesland airport were in the same range as the water concentrations reported from Gardermoen groundwater (611-7194 ng/l). 6:2 FTS could only be detected in fish liver samples from Fleslands airport, but the quantification was not quality assured.

Toxicological effects In general, not much toxicological information is available on 6:2 FTS, however one study describing the effect of elevated concentrations of 6:2 FTS on earthworm ( Eisenia fetida ) reproducibility has been performed by Stubberud (2006). The results from this study demonstrated that concentrations above 21 µg/g (6:2 FTS) were necessary to imply any decrease in reproducibility. This is about ten times higher than maximum concentration found in soil at Flesland airport.

Concluding remarks The detected levels of 6:2 FTS in comparison with the results from Stubberud (2006) could indicate that the concentrations at Flesland do not pose a problem to the local environment. However, since it was detected (but not quantified) in fish liver samples, it could indicate that 6:2 could spread in the environment and thus potentially have adverse effects to living organisms and is thus assessed as being of environmental concern.

Perfluorooctane sulfonate - PFOS

Results from this study Soil was only sampled at the Flesland airport (PFOS 1.6-1905 ng/g d.w), and the highest concentration was found at 10 m from the fire fighting training area (Figure 16). Concentrations of PFOS in sediments from Langavatn had average concentrations of 63, 69 and 53 ng/g (d.w) at the three sampling locations and the average water concentration was 1695 ng/l. PFOS was not detected in sea mussel sampled at Flesland brygge but was detected in crab (average concentration 2.3 ng/g w.w), whereas high concentrations were found in trout livers (2281 ng/g w.w).

At the Res-Q site, higher sediment concentrations were found at Res-Q-5 (average 416 ng PFOS/g d.w) compared to Flesland, whereas the concentration at Res-Q-4 (17 ng/g d.w) were lower. PFOS was not detected in Bleivika. The estimated water concentration was 156 g/l, but the analysis was interfered by unknown component(s) (possibly detergents as the water was “foaming”). PFOS was not detected in sea mussel, but the concentrations quantified in crab and fish liver had similar concentrations, both reaching average values of 2.1 ng/g w.w.

Table 47 Positive detections of PFOS. Location Water ng/l Sediment ng/g Soil ng/g Fish liver ng/g Crab ng/g d.w d.w w.w w.w Flesland airport 1427 - 2078 n.a 1.6 - 1905 n.a n.a Langavatn n.a 35 – 87.6 n.a 2082 - 2532 n.a Flesland brygge n.a n.a n.a n.a 0.8 – 4.9 Res-Q (Haugesund) 131 – 181* 5.3 - 493 n.a n.a n.a Bleivika (Haugesund) n.a o n.a 1.1 – 2.9* 0.7 – 4.1 n.a: not analysed, o: detected in single replicate and/or very close to detection limit. * uncertain values due to interferences

77 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

0 10 20 30 40 50 PFOS 75 ∑PFSAs 100 150 Distance from training ground (m) 200

0 500 1000 1500 2000 2500 ng/g soil d.w.

Figure 16 Concentration of PFOS and ∑PFSAs in the soil at the fire fighting practice ground at Fleslands Airport (ng/g d.w).

Comparison to previous studies When comparing PFOS concentrations it is important to emphasize the difference in quantification of linear or sum PFOS constitutional isomers. The present study only quantified the linear isomer while others, and perhaps more common in older studies, did not resolve different isomers and did a total quantification using the response of the linear isomer.

The concentration of PFOS in soils has rarely been recognized during screening investigations throughout Norway and the other Nordic countries, perhaps due to PFOS low partitioning in soil compared to water. The concentrations in the present study are however in the same range as data from other fire fighting training facilities in Norway (Amundsen, 2009). Compared to the threshold values set by the Norwegian Pollution Control Authority (100 ng/g), the PFOS-concentration is exceeded up to ca 50 m from the fire fighting ground at Flesland.

The concentration of PFOS found in sediments in the vicinity of Fleslands airport (Langavatn) are in general higher than sediment concentrations measured in other studies in Norway and other Nordic countries (

78 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Seepage water was analyzed from Res-Q and the concentrations were in the range 131-181 ng/l, which is considerable higher than concentrations reported from fresh waters around Scandinavia (Kallenborn et al., 2004; Woldegiorigis et al, 2006). The concentrations in seepage water from Flesland airport were even higher, ranging from 1427-2078 ng/l and are in the lower range of the concentrations measured in groundwater at Gardermoen Airport (2394-40116 ng/l; Amundsen et al., 2009).

In a screening project from 2007 (Green et al, 2008), one sea mussel sample from the recipient of Res-Q was analyzed, displaying higher concentration (1.89 ng/g w.w.) than in this study (<0.15 ng/g w.w.). Concentrations of PFOS in fish liver sampled in the sea outside Res-Q were low and in the lower range of the concentrations measured in a Nordic screening project (Kallenborn et al, 2004). The fish liver concentrations from the Langavatn near Flesland airport, on the other hand, displayed very high concentrations, above 2000 ng/g in all samples, which is in the same range as the concentrations measured in fish liver after an accidental spill of fire fighting foam close to Toronto international airport (Moody et al., 2002).

Toxicological effects Due to the persistence and toxicity of PFOS, substantial amounts of ecotoxicological data can be found in the literature (e.g. 3M company, 2003, ATSDR, 2009).

It has been established that soil concentrations above 10 µg/g could lead to reduced reproducibility for earthworms (Stubberud, 2006). At the fire fighting training ground at Flesland airport, the most polluted sample (10 m from the centre) had PFOS-concentration five times lower than this this value.

The toxicity of PFOS to water living organisms is most commonly referred to as concentration of PFOS in the water, and not a maximum tolerable liver concentration. For fresh water, 25 µg/l have been reported as PNEC (Predicted No Effect Concentration; Brook, 2004), whilst the same value for sea water is 2.5 µg/l. 25 µg/l, i.e. the same value which is applied in the Norwegian guideline for fjords and coastal waters (SFT, 2007). This indicate that the concentrations found at Fleslands airport are at least 10 times lower than the PNEC value and waters from Res-Q even lower and would not imply any effects to water living organisms.

PNEC values for sediment living organisms have been established to be 67µg/kg for fresh water sediments and 6.7 µg/kg for marine sediments. Again, the measured concentrations are below these values and would consequently not imply any risk to the sediment living organisms.

Concluding remarks Despite concentrations below some toxicological guidelines limits, it cannot be ruled out that the present contamination at both Flesland and Res-Q could pose a risk to the environment. Especially the high accumulation of PFOS in the fish liver samples from Flesland airport indicates bioaccumulation, and thus implies an environmental concern.

79 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Perfluorodecane sulfonate - PFDS Results from this study Soil concentrations of PFDS at Flesland airport ranged between <0.10-54 ng/g (d.w), with the highest concentration at 10 m distance from the training platform center. PFDS was also quantified in sediment samples from Langevatn near Flesland airport at low concentrations (<0.13-1.7 ng/g). PFDS was not found in the water samples or the sea mussels from Flesland airport, but elevated levels were found in fish liver (average 52 ng/g f.w.) and lower levels in crab (<0.05-0.56 ng/g f.w.)

PFDS was not found at or close to Res-Q in water, sea mussels or fish liver samples, and only in one sample of crab close to the detection limit (0.12 ng/g w.w). It was found in all sediment samples from the sedimentation basins, ranging between 0.16-0.21 and 3.3-5.7 ng/g (d.w) at sampling sites Res-Q-4 and 5, respectively. It was probably detected (high degree of uncertainty) in one sample from Bleivika.

Tabell 48 Positive detections of PFDS. Location Water ng/l Sediment ng/g Soil ng/g Fish liver ng/g Crab ng/g d.w d.w w.w w.w Flesland airport n.d n.a 0.13 - 54 n.a n.a Langavatn n.a 0.22 – 1.7 n.a 37 - 65 n.a Flesland brygge n.a n.a n.a n.a 0.13 – 0.56 Res-Q (Haugesund) n.d 0.16 – 5.7 n.a n.a n.a Bleivika (Haugesund) n.a o n.a n.d 0.12 n.d: not detected, n.a: not analysed, o: detected in single replicate and/or very close to detection limit.

PFDS

0 10 20 30 40 50 PFDS 75 100 150

Distance from training ground (m) 200

0 10 20 30 40 50 60 ng/g soil d.w.

Figure 17 Concentration of PFDS in the soil at the fire fighting practice ground at Fleslands Airport (ng/g d.w).

Comparison to previous studies The soil concentrations were in the same range as those measured at Mongstad airport (

80 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) different screening projects around Norway (

Toxicological effects No toxicological information found.

Concluding remarks Due to the scarce amount of information of PFDS, it is difficult to draw any thorough conclusions, but since it was detected in both sediment and biologic samples at Flesland, it could indicate spreading of the substance and thus potentially imply a problem to the environment it is assessed of being of environmental concern .

Perfluorohexanoic acid - PFHxA Results from this study Concentrations of PFHxA in soil samples from Flesland airport ranged between 0.18-18.5 ng/g (d.w), with the highest concentration at 30 m distance from the fire fighting training platform. The sediment concentrations measured outside Flesland airport varied between

PFHxA was detected in sediment samples from Res-Q-5 with an average concentration of 17 ng/g (d.w). It was not detected in Bleiviken, and only in one sample from Res-Q-4 (0.5 ng/g d.w). Water concentrations were lower compared to Flesland with an average value of 78 ng/l. PFHxA was not found in any biological samples from Res-Q.

Tabell 49 Positive detections of PFHxA. Location Water ng/l Sediment Soil ng/g Fish liver ng/g Crab ng/g ng/g d.w d.w w.w w.w Flesland airport 384.6 – 553.5 n.a 0.18 – 18.5 n.a n.a Langavatn n.a 0.45 – 1.6 n.a 0.24 n.a Res-Q (Haugesund) 64 - 89 0.5 - 19 n.a n.a n.a n.a: not analysed

81 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

PFHxA

0 10 20 30 40 50 PFHxA 75 100

Distance from ground (m) training 150 200

0 5 10 15 20 ng/g soil d.w.

Figure 18 Concentration of PFHxA in the soil at the fire fighting practice ground at Fleslands Airport (ng/g d.w).

Comparison to previous studies Soil concentrations were in the same range as those reported from Gardemoen airport (

Toxicological effects The toxicology of PFHxA has recently been demonstrated through two different studies, where rats have been fed PFHxA in different doses (Chengelis et al, 2009 and Loveless et al, 2009). The lowest reported NOAEL from the two studies was 20 mg/kg/day (Loveless et al, 2009). It is, however, difficult to compare this value with the findings in this investigation.

Concluding remarks Studies on rat indicate a quite high tolerable dose. In addition, PFHxA was only found in one biological sample, which point toward a low bioaccumulation. Regarding the potential risk of PFHxA in the environment, little information is available. However, quite high concentrations of PFHxA were found at the sites, especially in the water, and environmental problems related to this cannot be ruled out. Overall PFHxA is assessed of being of moderat environmental concern to environmental concern .

82 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Other PFCAs Results from this study The amount of PFCAs in the different samples can be seen in Appendix I – Analytical results. In agreement with the results for the PFSAs, the concentrations and number of samples >dl were in general higher in samples from Flesland airport than those from Res-Q. The highest soil concentrations were observed for the compounds with short chain length (0.5-28 ng/g d.w; PFPeA), with decreasing concentrations with longer chain length. PFDA and PFTeA were not found in any soil sample.

Decreasing concentrations of PFCAs with increasing chain length was also found in the sediments, both from Res-Q and Flesland. In the sediments, an interesting phenomenon was observed: compounds with an odd number carbon chain length (PFUnDA and PFTrDA) were present in higher amount than those with even number chains (PFDoDA and PFTeDA). This pattern has been observed in marine environments in earlier studies (e.g. Holmström, 2008), and could be attributed to estimated emission quantities of odd chain homologues compared to even chain compounds (Prevedouros et al, 2006).

Detectable concentrations of PFPeA-PFOS (Res-Q) and PFPeA-PFUnDA (Flesland airport) were found in the water, and the concentrations were generally higher in samples from Flesland than Res-Q. PFOA concentrations ranged between 31-44 ng/l (Res-Q) and 130-191 ng/l (Flesland).

No PFCAs were found in sea mussels from Res-Q and the only PFCA present in sea mussels from Flesland was PFTrDA (average 0.41 ng/g d.w). From Res-Q, PFOA and PFUnDA were detected in both crab and fish liver, whereas PFDA only in fish liver and PFTrDA only in crab. Long chained PFCAs were detected at elevated concentrations in fish liver from Flesland with highest concentration for odd number compounds (PFUnDA 101 ng/g, PFTrDA 44 ng/g).

Comparison to previous studies In comparison with the soil concentration at other fire fighting training facilities, the present concentrations are generally similar or lower than those measured in previous studies (Amundsen, 2009). For example PFOA is lower than previously reported from both Gardermoen (

In agreement with the results from the PFSAs in the sediment samples, the concentration of short chain PFCAs was lower than those reported from Rygge old airport (Amundsen et al, 2009). Furthermore, the concentrations of PFOA from Res-Q-5 were elevated in comparison with different screening projects (Bakke et al, 2007; Kallenborn et al, 2004), whilst the other sediment samples were not, also in agreement with the PFSA-results. Regarding the long chain PFCAs, the results from the present study indicates higher concentrations than previous studies, especially regarding those substances with odd numbered chain lengths. Concentrations of PFUnDA from Langavatn 3 ranged between 21-25 ng/g (d.w), which is higher than concentrations reported from Rygge old airport (

83 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Also the concentration of long chained PFCAs (>C 10 ) in water was higher than previously reported from Gardemoen groundwater (Amundsen et al, 2009).

The low concentrations of PFOA found in sea mussels (<0.10 ng/g w.w) were in agreement with most screening studies (e.g. Kallenborn et al, 2004), and PFOA and PFUnDA-levels in fish liver samples from Res-Q were similar to those measured in other fish liver samples around Norway (Bakke et al, 2007). In contrast, the high concentrations of PFUnDA, PFDoDA, and PFTrDA in fish liver from Flesland airport, where higher than previously reported.

Toxicological effects The most well studied PFCA is PFOA, mainly since it has been produced in such high amounts and thus spread in the environment to a relatively high extent (Prevedouros et al, 2006). Stubberud (2006) measured the effect of soil concentration on the reproducibility of earth worms, and concluded that a concentration of 16 mg PFOA/kg may have consequences for the earth worms. This is approximately 1000 times higher than the highest concentration measured in the present soil (12 ng/g).

Despite PFOA being the most studied, a recent study by Liu et al (2008), demonstrated that both PFDoDA and PFTeDA were more toxic to algae than PFOA. However, quite high water concentrations were necessary (20 mg/L for PFDoDA and PFTeDA, respectively), and neither of the compounds were detected in any water sample during the present study. The shorter chain homologue PFBA, on the other hand, has been demonstrated to be less toxic than PFOA (Das et al, 2008).

The relatively high concentrations of long chained PFCAs present in the biota samples, especially from Flesland airport, could to one extent be explained by differences in the bioconcentration factor (BCF) between the long chained and short chained homologues. Martin et al (2003) demonstrated that the BCF for long chained PFCAs increased by a factor of eight with each additional carbon between 8 and 12 carbons, and could also explain the pattern in the present study.

Concluding remarks It is clear that elevated levels of different PFCAs are present at both Res-Q and Flesland airport. Despite regulatory guidelines and toxicity safe-levels are not available, it cannot be ruled out that the present concentrations may imply an environmental concern. In addition, the higher bioaccumulation of long chained homologues is evident. In the soil and sediment samples, the long chained homologues are present in lower concentrations than the short chained PFCAs, whereas the opposite is true in the biological samples.

84 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Overall discussion PFCs All of the investigated PFCs were detected in one or more of the investigated environmental compartments (matrices). All analysed compounds were found in soil and sediments except PFDA and PFTeDA. In the water samples all compounds were detected except PFDS, PFDoDA, PFTrDA and PFTeDA. Only one PFTrDA was detected in blue mussel and PFBS, PFPeA and PFHpA were the only compounds not detected in biological material.

The PFC-results from Res-Q and Flesland fire fighting training facilities indicate that local contamination is present at both sites, with Flesland generally displaying higher concentrations. In addition, differences between matrices and sampling locations can be seen. The highest concentrations measured (approx. 2000 parts per billion) in this study were for PFOS in soil and fish liver, and also for 6:2 FTS in soil, all samples taken near Flesland airport. PFOS and 6:2 FTS were also found in elevated concentrations in water from Flesland.

In the water samples, the concentrations found in Flesland were higher than those found at Res-Q for many compounds (e.g. PFBS, PFHxS, PFOS, PFHxA, PFHpA, PFOA, PFNA and PFDA). This may be due to the cleaning of the water performed at Res-Q. It should however be noted that despite the cleaning process, the concentration of for example PFOS and PFOA at Res-Q were higher than those reported from screening projects in the Nordic countries (Kallenborn et al, 2004), indicating a local contamination.

Despite the generally higher water concentration at Flesland compared to Res-Q, the highest concentration of each of the compounds in the sediment was found at one sampling location near Res-Q (Res-Q-5). The average concentrations of PFSAs measured at Res-Q-5 were in the range 2.5- 30 times higher than the average concentrations measured in the other sediment samples. The same pattern was also observed for some of the PFCAs, e.g. PFHxA, PFHpA, PFOA. The Res-Q-5 sediment is sampled in a bore hole draining the two sedimentation basins, and the present results demonstrate that this location is highly contaminated by PFCs. But also the sediments in Langavatn, near Flesland, had concentrations of PFHxS, 6:2 FTS, PFOS, PFHxA, PFHpA, PFUnDA and PFDoDA above those measured in different screening projects (Bakke et al, 2007, 2008 etc.), and should thus also be considered contaminated and imply a potential concern to the surrounding environment. No PFC’s were detected in sediments from Bleivika the, marine receiving water of seepage water from Res-Q. It is worth mentioning that the sediment samples consisted of relatively coarse material which may have affected the results because many of the investigated compounds have a higher affinity for finer material (finer grains).

It is notable that contamination of PFCAs also takes place at both Flesland and Res-Q. Carboxylates were detected in water from both locations but to a larger extent at Flesland, ranging from carbon chain length C5 to C11 with PFPeA at the third highest concentration (average 560 ng/l) after PFOS and 6:2 FTS. Fluorotelomer alcohols are used as agents in AFFF (Prevedouros et al., 2006) which might contain traces of PFCAs or biotransform into PFCAs.

Soil was only sampled at the fire fighting facility at Flesland airport, and the results clearly demonstrates contamination of PFC, most likely due to the utilization of AFFFs. A decrease in concentration is observed with distance from the training platform, with exception of the sample taken at the training ground (0 m) which was lower than the sample taken 10 m from

85 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) the center. The reason for this is unknown, it should however be noted that the results from the sampling points 0-30 m from the training ground have a higher degree of uncertainty in comparison with the others due to interfering components.

It is interesting to note that the 6:2 FTS concentrations in soil and water from Flesland are similar or even higher than the PFOS-concentrations. However, the corresponding concentrations of the two compounds in the sediment samples from Langavatn outside Flesland are the opposite, with PFOS-concentrations exceeding the 6:2 FTS-concentrations by approximately a factor 10. This might be an indication that PFOS, which should no longer be utilized at the site, sediment levels are a result from historical use and the relatively high 6:2 FTS levels in soil is a result of ongoing use. The relatively low levels of 6:2 FTS in sediments can also indicate environmental degradation.

The replacement of PFOS, PFBuS, was found in soil from Flesland airport, sediment from Res-Q-5, and water from both locations.

In the biota samples, the highest concentrations (and also the most homologues) were found in fish liver while only PFTrDA was found in low concentrations in sea mussels. This could most likely be explained by differences in their trophic levels, with fish being higher on the food chain compared to mussels. In addition, PFCs accumulate in the liver consequently resulting in concentrated levels for the fish, should whole fish be analyzed lower levels would be found. It is however evident that the concentrations of PFOS, PFHxS, PFDS, PFUnDA and PFTrDA in fish liver from Flesland are elevated. For example the concentration of PFOS was more than thousand times higher at Flesland compared to Res-Q. Different species were taken at the two locations but the difference is most likely due to the sampling sites. Fish sampled near Flesland originate from a small fresh water lake, receiving the seepage water from the fire fighting training area, whereas the fish from Res-Q were taken in the adjacent sea, where the concentration of contaminants could be expected to be more diluted. Moreover, there is a clear pattern between odd and even chain homologues of PFCAs in fish liver from Flesland. In the liver, levels of PFDA (11 ng/g) and PFDoDA(7.5 ng/g) are lower compared to PFUnDA (101 ng/g) and PFTrDA (44 ng/g). This pattern is in accordance with the estimated global emission of PFCAs in year 2000 (Prevedouros et al., 2006).

From the present results, differences between short and long chained PFCs can be observed. The short chained homologues analyzed (e.g. PFBS, PFPeA, PFHxA and PFHpA) are found in the “receiving matrices” (soil, sediment and water), but are not found or only found in small amounts in the biota samples. This could be explained by the low bioaccumulation ability of these homologues, and they are consequently not concentrated in the food chain. The long chained homologues, on the other hand, display an opposite tendency. For example the concentration of PFDS is similar or higher in fish liver samples (from Flesland) compared to the corresponding sediment and soil samples, which indicates bioconcentration. The same pattern is also observed for PFNA, PFUnDA, PFDoDA and PFTrDA. This highlights the importance of measuring contaminants in diverse environmental matrices, since the bioaccumulating homologues may not be present in such high concentrations at the hot spots, but can still reach high concentration in the surrounding biota.

In general, the toxicological information available of the different PFCs is scarce, with the exception of PFOS and PFOA. Consequently, it is difficult to draw any conclusion of the environmental concern of the different homologues. In addition, the toxicological information available concerns the effect of single PFCs on a certain organism or response

86 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) factor, however at the present sites, a mixture of different compounds are present. This may result in other toxicological responses in comparison with those which could be expected from single compound exposure. Analysis of the toxicity of the present matrices (soil, sediment, water) on test organisms would thus give interesting additional information of the actual toxicity present at the sites.

87 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

7. Conclusions 7.1 Brominated flame retardants (BFRs)

Figure 19 presents the highest determined concentrations in the different environmental matrices investigated in this study along with the examples of high concentration from the same matrices in the peer reviewed literature. Also shown in the figure is an indication of whether the compound is assessed of being of no or little environmental concern, moderate environmental concern or of environmental concern, according to the criteria presented in Chapter 6.

10 100 1000

TBPH EHTBB ATE TBPA DPTE TBBPA-DBPE BTBPI ? TBBPA-AE TBP DBDPE BTBPE HBB PBEB PBT

Figure 19 A summary of the highest sediment ( n/g d.w), highest sludge ( ng/g d.w), highest water 3 ( ng/l), highest air ( pg/m ), highest biota ( ng/g w.w) of “new” BFRs in this investigation compared to examples of high sediment ( ng/g d.w), high sludge ( ng/g d.w), high water ( ng/l), high air ( pg/m3) and high biota ( ng/g w.w) in the peer reviewed literature. Green shading indicates a compound assessed to be of no or little environmental concern, yellow shading indicates compounds assessed to be of moderate environmental concern and red shading indicates compounds assessed to be of environmental concern. Symbols in yellow frame and a question mark indicate that no data is found in the peer reviewed literature.

BFRs not detected in any samples TBPH, EHTBB, ATE, DPTE, and TBPA were not detected in any sample in the present study, and these compounds are therefore assessed to be of little environmental concern. However, more studies on these compounds are needed to confirm the low environmental levels in Norway.

88 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

BFRs detected in one or more samples The “new”BFRs detected were: TBBPA-DBPE, PBT, TBT, PBEB, HBB, BTBPE, DBDPE, TBBPA-AE and BTBPI. Of these TBP, DBDPE and BTBPE are assessed of being of environmental concern.

TBP is assessed of being of environmental concern because it is the only compound found in fish, crab and mussel. In addition it is detected in sediments. Bromophenols are naturally produced by many marine organisms so it may be that the findings of TBP is of natural origin but here it’s chosen to classify it as being of environmental concern.

DBDPE is assessed as being of moderate environmental concern to being of environmental concern. The compound is not detected in crab, mussel or fish so it could be assessed as being of moderate environmental concern only. It is detected in pine needles but these are considered more as passive samplers. The reason it is assessed of being of environment concern is the relatively high concentration in seepage water (186 ng/l) compared to the applied PNEC (Predicted No Effect Concentration) of 200 ng/l.

BTBPE is also assessed of being moderate environmental concern to being of environmental concern. Like DBDPE it’s not detected in crab, mussel or fish samples. The reason it’s assessed as being of environmental concern is that it’s detected in biota samples in the arctic (KLIF unpublished data) and that the concentration in seepage water of 107 ng/l is relatively high.

TBBPA-DBPE, PBT, PBEB, HBB, TBBPA-AE and BTBPI are all assessed of being of moderate environmental concern due to the fact that they are detected in the receiving environment.

In general, the toxicological information available of the different new BFRs is scarce, and consequently it is difficult to draw any conclusion of the environmental concern of these BFRs based on such information.

Some of the compounds considered in the present study were analysed in environmental samples for the first time and the uncertainties related to these results are larger than for compounds that has been analysed before, participated in inter-laboratory calibration, or for which labeled internal standards are available. As work progress and more studies on these compounds are performed the uncertainties in the analysis will be smaller, i.e. the results from a first screening study should not uncritically be used to study time trends.

7.2 Perfluorinated compounds (PFCs)

Figure 20 presents the highest determined concentrations in the different environmental matrices investigated in this study along with the examples of high concentration from the same matrices in the peer reviewed literature. Note that only the 5 compounds originally included in the analyzing program are shown. Also shown in the figure is an indication of whether the compound is assessed of being of moderate environmental concern or of environmental concern, according to the criteria presented in Chapter 6.

89 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

10 100 1000 10000

PFHxS

PFDS 47 000 ng/g

PFOS ?

6:2 FTS

PFBS 16 000 ng/l

Figure 20 A summary of the highest sediment ( n/g d.w), highest water ( ng/l), highest soil ( ng/g d.w), highest biota ( ng/g w.w), of PFCs (5 compounds within original scope of work) in this investigation compared to examples of high sediment ( ng/g d.w), high water ( ng/l), high soil ( ng/g d.w), high biota ( ng/g w.w) in the peer reviewed literature. Yellow shading indicates compounds assessed to be of moderate environmental concern and red shading indicates compounds assessed to be of environmental concern. Symbols in yellow frame and a question mark indicate that no data is found in the peer reviewed literature.

All of the investigated PFCs were detected in one or more of the investigated environmental compartments (matrices). All analyzed compounds were found in soil and sediments except PFDA and PFTeDA. In the water samples all compounds were detected except PFDS, PFDoDA, PFTrDA and PFTeDA. Only one PFTrDA was detected in blue mussel.

Overall PFBS, PFPeA and PFHpA were the only compounds not detected in biological material. They are all detected in the receiving environment and are therefore here assessed of being of moderate environmental concern. All others analyzed PFC are detected in biota and therefore assessed of being of environmental concern.

In general, the toxicological information available of the different PFCs is scarce, with the exception of PFOS and PFOA. Consequently, it is difficult to draw any conclusion of the environmental concern of the different homologues. The registration requirements of chemicals under the EU regulation on chemicals and their safe use (REACH) may cause that more toxicological information will be available in the future. In addition, the toxicological information available concerns the effect of single PFCs on a certain organism or response factor, however at the present sites, a mixture of different compounds are present. This may result in other toxicological responses in comparison with those which could be expected from single compound exposure. Analysis of the toxicity of the present matrices (soil, sediment, water) on test organisms would thus give interesting additional information of the actual toxicity present at the sites.

90 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

8. References

3M company, 2003. Environmental and health assessment of perfluoroocane sulfonic acid and its salts.

Ahn, Y.B., Rhee, S.K., Fennell, D.E., Kerkhof, L. J., Hentschel, U., Haeggblom, M.M. (2003) Reductive dehalogenation of brominated phenolic compounds by microorganisms associated with the marine sponge Aplysina aerophoba. Appl.Environ.Microbiol. 69 , 4159- 4166.

Amundsen, C.E., Forfang, I., Aasen, R., Eggen, T., Sørheim, R., Hartnik, T, Næs, K. 2009. Screening of polyfluorinated organic compounds at four fire training facilities in Norway. Norwegian State Pollution Control Agency, report TA-2444.

Arbeli, Z., Ronen, Z. and Diaz-Baez, M. C. (2006). Reductive dehalogenation of tetrabromobisphenol-A by sediment from a contaminated ephemeral streambed and an enrichment culture. Chemosphere . 64 , 1472-1478.

Armitage, J. M., MacLeod, M. and Cousins, I. T. (2009). Response to Comment on "Comparative Assessment of the Global Fate and Transport Pathways of Long-Chain Perfluorocarboxylic Acids (PFCAs) and Perfluorocarboxylates (PFCs) Emitted from Direct Sources". Environmental Science & Technology . 43 , 7153-7154.

Arp, H. P. H. and Goss, K. U. (2009). Gas/Particle Partitioning Behavior of Perfluorocarboxylic Acids with Terrestrial Aerosols. Environmental Science & Technology . 43 , 8542-8547.

Arp, H. P. H., Niederer, C. and Goss, K. U. (2006). Predicting the partitioning behavior of various highly fluorinated compounds. Environmental Science & Technology . 40 , 7298- 7304.

ATSDR. 2009. Draft Toxicological profile for perfluoroalkyls. U.S. Department of Health and Human Services. Agency for Toxic Substances and Disease Registry.

Bakke, T., Fjeld, E., Skaare, B.B., Berge, J.A., Green, N., Ruus, A., Schlabach, M., Botnen, H. 2007. Kartlegging av metaller og utvalgte nye organiske miljøgifter 2006. Krom, arsen, perfluoralkylstoffer, dikloretan, klorbenzener, pentaklorfenol, HCBD og DEHP (in Norwegian). Norwegian State Pollution Control Agency, report TA-2284.

Bakke, T., Boitsov, S., Brevik, E.M., Gabrielsen, G.W., Green, N., Helgasson, L.B., Klungsøyr, J., Leknes, H., Miljeteig, C., Måge, A., Rolfsnes B.E., Savinova, T., Schlabach, M., Skaare, B.B., Valdersnes, S. 2008. Mapping selected organic contaminants in the Barents Sea 2007. Norwegian State Pollution Control Agency, report TA-2400.

Belfroid, A., Meiling, J., Drenth, H. J., Hermens, J., Seinen, W. and Vangestel, K. (1995). DIETARY UPTAKE OF SUPERLIPOPHILIC COMPOUNDS BY EARTHWORMS (EISENIA-ANDREI). Ecotoxicology and Environmental Safety . 31 , 185-191.

91 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Bindra & Narang, 1995. Combustion of flame retardants. Chemosphere, Volume 31, Issues 11-12. pp. 4413-4425.

Bossi, R., Riget, F. F. and Dietz, R. (2005). Temporal and spatial trends of perfluorinated compounds in ringed seal (Phoca hispida) from Greenland. Environmental Science & Technology . 39 , 7416-7422.

Boyle, J.L, Lindsay, R.C, Stuiber, D.A (1992). Bromophenol Distribution in Salmon and Selected Seafoods of Fresh- and Saltwater Origin. Journal of Food Science, Volume 57, Issue 4, pp. 918 – 922.

Breitholtz M, Nyholm JR, Karlsson J, Andersson PL. Are individual NOEC levels safe for mixtures? A study on mixture toxicity of brominated flame-retardants in the copepod Nitocra spinipes . Chemosphere 2008, 72: 1242-1249.

Breivik, K., Wania, F., Muir, D. C. G., Alaee, M., Backus, S. and Pacepavicius, G. (2006). Empirical and modeling evidence of the long-range atmospheric transport of decabromodiphenyl ether. Environmental Science & Technology . 40 , 4612-4618.

Brooke, D., Foottitt, A., Nwaogu, T.A. 2004. Environmental risk evaluation report: perfluorooctane sulfonate (PFOS). UK Environmental Agency.

Brown, T. N. and Wania, F. (2008) Screening chemicals for the potential to be persistent organic pollutants: a case study of arctic contaminants. Environ. Sci. Technol., 42, 5202- 5209.

Butt, C. M., Muir, D. C. G., Stirling, I., Kwan, M. and Mabury, S. A. (2007). Rapid response of arctic ringed seals to changes in perfluoroalkyl production. Environmental Science & Technology . 41 , 42-49.

Campbell, T. Y., Vecitis, C. D., Mader, B. T. and Hoffmann, M. R. (2009). Perfluorinated Surfactant Chain-Length Effects on Sonochemical Kinetics. Journal of Physical Chemistry A. 113 , 9834-9842.

Canada, G. o. (2008). Perfluorooctane Sulfonate and its Salts and Certain Other Compounds Regulations (P.C. 2008-974). Canada Gazette Part II . 142 , 1322-1347.

Chatonnet, P., Bonnet, S., Boutou, S., Labadie, M.-D. 2004. Identification and responsibility 2,4,6 –tribromoanisole in musty corked odors wine. J. Agric. Food Chem. 51, 6752-6760.

Chemtura (2008) Chemtura, Firemaster 550. Great Lakes Chemical Co. URL: http://www.e1.greatlakes.com/freb/common/pdf/Page2_F550vsPenta.pdf

92 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Chen, S. J., Ma, Y. J., Wang, J., Chen, D., Luo, X. J. and Mai, B. X. (2009). Brominated Flame Retardants in Children's Toys: Concentration, Composition, and Children's Exposure and Risk Assessment. Environmental Science & Technology . 43 , 4200-4206.

Cheng, J., Psillakis, E., Hoffmann, M. R. and Colussi, A. J. (2009). Acid Dissociation versus Molecular Association of Perfluoroalkyl Oxoacids: Environmental Implications. Journal of Physical Chemistry A . 113 , 8152-8156.

Chengelis, C.B., Kirkpatrick, J.B., Radovsky, A., Shinohara, M. 2009. A 90-day repeated dose oral (gavage) toxicity study of perfluorohexanoic acid (PFHxA) in rats (with functional observational battery and motor activity determinations). Reproductive Toxicology, 27:342- 351

Covaci, A., Gerecke, A. C., Law, R. J., Voorspoels, S., Kohler, M., Heeb, N. V., Leslie, H., Allchin, C. R. and de Boer, J. (2006). Hexabromocyclododecanes (HBCDs) in the environment and humans: A review. Environmental Science & Technology . 40 , 3679-3688.

Covaci, A., Voorspoels, S., Abdallah, M. A. E., Geens, T., Harrad, S. and Law, R. J. (2009). Analytical and environmental aspects of the flame retardant tetrabromobisphenol-A and its derivatives. Journal of Chromatography A . 1216 , 346-363.

Chung, H.Y., Ma, W.C.J., Kim, J.S. 2003 a. Seasonal distribution of bromophenols in selected Hong Kong seafood. J. Agric. Food Chem. 51 6752-6760.

Chung, H.Y., Ma, W.C.J., Kim, J.S., Chen, F. 2003 b. Seasonal variations of bromophenols in brown algae (Padina arborescens, Sargassum siliquastratum and Lobophora variegate) collected in Hong Kong. J. Agric. Food Chem. 51 2619-2624.

Darnerud, P. O. (2003). Toxic effects of brominated flame retardants in man and in wildlife. Environment International . 29 , 841-853.

Das, K.P., Grey, B.E., Zehr, R.D., Wood, C.R., Butenhoff, J.L., Chang, S.-C., Ehresman, D.J., Tan, Y.-M., Lau, C. 2008. Effects of perfluorobutyrate exposure during pregnancy in the mouse. Toxicological Sciences, 105:173-181.

Davis, E. F. and Stapleton, H. M. (2009). Photodegradation Pathways of Nonabrominated Diphenyl Ethers, 2-Ethylhexyltetrabromobenzoate and Di(2-ethylhexyl)tetrabromophthalate: Identifying Potential Markers of Photodegradation. Environmental Science & Technology . 43 , 5739-5746.

Davis, M. F., Gracz, H., Vendeix, F. A. P., de Serrano, V., Somasundaram, A., Decatur, S. M. and Franzen, S. (2009). Different Modes of Binding of Mono-, Di-, and Trihalogenated Phenols to the Hemoglobin Dehaloperoxidase from Amphitrite ornata. Biochemistry . 48 , 2164-2172. de Wit, C. A. (2002). An overview of brominated flame retardants in the environment. Chemosphere . 46 , 583-624.

93 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Dietz, R., Bossi, R., Riget, F. F., Sonne, C. and Born, E. W. (2008). Increasing perfluoroalkyl contaminants in east Greenland polar bears (Ursus maritimus): A new toxic threat to the Arctic bears. Environmental Science & Technology . 42 , 2701-2707.

Dungey, S., and Akintoye, L. 2007. Environmental risk evaluation report: 1,1`(ethane-1,2 diyl)bis(pentabromobenzene) CAS: 84852—53-9. Product code SCHO0507BMOR-E-P. Environmental Agency for England and Wales. Available at www.environment- agency.gov.uk.

EFSA (2008). Opinion of the Scientific Panel on Contaminants in the Food chain on Perfluoroctane sulfonate (PFOS), Perfluorooctanoic acid (PFOA) and their salts. The EFSA Journal . 653 , 1-131.

EPA (2006). 2010/15 PFOA Stewardship Program. Prepared by Report (http://www.epa.gov/oppt/pfoa/pubs/stewardship/pfoaguidance.pdf )

EPA (2009). 2009 Edition of the DrinkingWater Standards and HealthAdvisories. Prepared by Report ( http://www.epa.gov/waterscience/criteria/drinking/dwstandards2009.pdf )

Ericson, I., Domingo, J.L., Nadal, M., Bigas, E., Llebaria, X., van Bavel, B., Lindstöm, G. 2009. Levels of Perfluorinated chemicals in municipal drinking water from Catalonia, Spain: Public health implications. Archives of environmental contamination and toxicology 57:631- 38.

EU (2003). The ban of marketing the Penta- and Octa-BDE mixtures. Directive 2003/11/EC of the European parliament and of the council of 6 february 2003 amending for the 24th time Council Directive 76/769/EEC relating to restrictions on the marketing and use of certain dangerous substances and preparations (pentabromodiphenyl ether, octabromodiphenyl ether). Off. J. Eur. Union, L 42, 15.2.2003, 45-46.

EU (2006). DIRECTIVE 2006/122/ECOF THE EUROPEAN PARLIAMENT AND OF THE COUNCIL of 12 December 2006. Official Journal of the European Union . L 372 , 33-34.

European Commission “Directive concerning the performance of analytical methods and the interpretation of results”. Council Directive 96/23/EC, notified under C(2002) 3044

European Union 2001. European Union Risk Assessment Report. Diphenyl ether, pentabromo derivative.

European Union 2003. European Union Risk Assessment Report. Diphenyl ether, octabromo derivative.

European Union 2002. European Union Risk Assessment Report. Bis(pentabromophenyl)ether.

European Union 2007. European Union Risk Assessment Report. Hexabromocyclododecane.

Fink, U., Hajduk, F. and Ishikawa, Y. (2005). Flame Retardants. Prepared by Report (http://www.sriconsulting.com/ )

94 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Fjeld, E., Schlabach, M., Berge, J.A., Green, N., Eggen, T., Snilsberg, P., Vogelsang, C., Rognerud, S., Kjellberg, G., Enge, E.K., Dye, C., A., Gundersen, H. 2005. Kartlegging av utvalgte nye organiske miljøgifter 2004. Bromerte flammhemmere, perfluoralkylstoffer, irgarol, diuron, BHT og dicofol (in Norwegian). Norwegian State Pollution Control Agency, report TA-2096.

Flodin, C. and Whitfield, F.B.1999. 4-hydroxybenzoic acid: a likely precursor of 2,4,6- tribromophenol in Ulva lactuca . Phytochemistry, volume 51, Issue 2. pp. 249-255

Gao, F., Luo, X. J., Yang, Z. F., Wang, X. M. and Mai, B. X. (2009). Brominated Flame Retardants, Polychlorinated Biphenyls, and Organochlorine Pesticides in Bird Eggs from the Yellow River Delta, North China. Environmental Science & Technology . 43 , 6956-6962.

Gauthier, L.T., Hebert, C.E., Weseloh, D.V.C., Letcher, R.J. (2007) Current-use flame retardants in the eggs of herring gulls (Larus argentatus) from the Laurentian Great lakes. Environ. Sci. Technol., 41, 4561-4567.

Gauthier, L. T., Potter, D., Hebert, C. E. and Letcher, R. J. (2009). Temporal Trends and Spatial Distribution of Non-polybrominated Diphenyl Ether Flame Retardants in the Eggs of Colonial Populations of Great Lakes Herring Gulls. Environmental Science & Technology . 43 , 312-317.

Gente, G., La Mesa, C., Muzzalupo, R. and Ranieri, G. A. (2000). Micelle formation and phase equilibria in a water-trifluoroethanol-fluorocarbon surfactant system. Langmuir . 16 , 7914-7919.

Goss, K. U. and Arp, H. P. H. (2009). Comment on "Experimental pK(a) Determination for Perfluorooctanoic Acid (PFOA) and the Potential Impact of pK(a) Concentration Dependence on Laboratory-Measured Partitioning Phenomena and Envrionmental Modeling". Environmental Science & Technology . 43 , 5150-5151.

Goss, K. U. (2008). The pK(a) values of PFOA and other highly fluorinated carboxylic acids. Environmental Science & Technology . 42 , 456-458.

Goss, K. U. and Bronner, G. (2006). What is so special about the sorption behavior of highly fluorinated compounds? Journal of Physical Chemistry A . 110 , 9518-9522.

Gouteux, B., Alaee, M., Mabury, S. A., Pacepavicius, G. and Muir, D. C. G. (2008). Polymeric Brominated Flame Retardants: Are They a Relevant Source of Emerging Brominated Aromatic Compounds in the Environment? Environmental Science & Technology . 42 , 9039-9044.

Green, N., Schlabach, M., Bakke, T., Brevik, E.M., Dye, C., Herzke, D., Huber, S., Plosz, B., Remberger, M., Schøyen, M., Uggerud, H.T., Vogelsang, C. 2008. Screening of selected metals and new organic contaminants 2007. Phosphorus flame retardents, polyfluorinated

95 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) organic compounds, nitro-PAHs, silver, platinum and sucralose in air, wastewater treatment facilities, and recipients. Norwegian Pollution Control Agency, report TA-2367.

Hardy, M.L., Margitich, D., Ackerman, L., Smith, R.L. (2002) The subchronic oral toxicity of ethane, 1,2-bis(pentabromophenyl) (Saytex 8010) in rats. Int. J. Toxicol., 21, 165-170.

Hardy, M., Biesemeier, J., Banasik, M. and Stedeford, T. (2008). Comment on "Alternate and New Brominated Flame Retardants Detected in U.S. House Dust". Environmental Science & Technology . 42 , 9453-9454.

Hardy, M. L. and Ranken, P. F. (2008). Brominated Flame Retardants, Polychlorinated Biphenyls, and Organochlorine Pesticides in Captive Giant Panda (Ailuropoda melanoleuca) and Red Panda (Ailurus fulgens) from China. Environmental Science & Technology . 42 , 8172-8172.

Hardy, M. L., Banasik, M. and Stedeford, T. (2009). Toxicology and human health assessment of decabromodiphenyl ether. Critical Reviews in Toxicology . 39 , 1-44.

Harrad, S., Ibarra, C., Abdallah, M. A. E., Boon, R., Neels, H. and Covaci, A. (2008). Concentrations of brominated flame retardants in dust from United Kingdom cars, homes, and offices: Causes of variability and implications for human exposure. Environment International . 34 , 1170-1175.

Holmström, Katrin. 2008. Tissue distribution of perfluorinated surfactants in common guillemot ( Uria aalge ) from the Baltic Sea. Environmental Science & Technology, 42:5879- 5884.

Hu, G. C., Dai, J. Y., Xu, M. Q., Luo, X. J. and Mai, B. X. (2008a). Response to Comment on "Brominated Flame Retardants, Polychlorinated Biphenyls, and Organochlorine Pesticides in Captive Giant Panda (Ailuropoda melanoleuca) and Red Panda (Ailurus fulgens) from China". Environmental Science & Technology . 42 , 8173-8174.

Hu, G. C., Luo, X. J., Dai, J. Y., Zhang, X. L., Wu, H., Zhang, C. L., Guo, W., Xu, M. Q., Mai, B. X. and Wei, F. W. (2008b). Brominated flame retardants, polychlorinated biphenyls, and organochlorine pesticides in captive giant panda (Ailuropoda melanoleuca) and red panda (Ailurus fulgens) from China. Environmental Science & Technology . 42 , 4704-4709.

Goss, K. U. (2008). The pK(a) values of PFOA and other highly fluorinated carboxylic acids. Environmental Science & Technology . 42 , 456-458.

Goss, K. U. and Arp, H. P. H. (2009). Comment on "Experimental pK(a) Determination for Perfluorooctanoic Acid (PFOA) and the Potential Impact of pK(a) Concentration Dependence on Laboratory-Measured Partitioning Phenomena and Envrionmental Modeling". Environmental Science & Technology . 43 , 5150-5151.

96 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Hart, K., Gill, V. A. and Kannan, K. (2009). Temporal Trends (1992-2007) of Perfluorinated Chemicals in Northern Sea Otters (Enhydra lutris kenyoni) from South-Central Alaska. Archives of Environmental Contamination and Toxicology . 56 , 607-614.

Hassenkloever et al., 2006 mangler I reflista

Higgins, C. P. and Luthy, R. G. (2006). Sorption of perfluorinated surfactants on sediments. Environmental Science & Technology . 40 , 7251-7256.

Higgins, C. P. and Luthy, R. G. (2007). Modeling sorption of anionic surfactants onto sediment materials: An a priori approach for perfluoroalkyl surfactants and linear alkylbenzene sulfonates (vol 110, pg 14054, 2006). Environmental Science & Technology . 41 , 6316-6316.

Hites, R. A. (2004). Polybrominated diphenyl ethers in the environment and in people: A meta-analysis of concentrations. Environmental Science & Technology . 38 , 945-956.

Holmstrom, K. E., Jarnberg, U. and Bignert, A. (2005). Temporal trends of PFOS and PFOA in guillemot eggs from the Baltic Sea, 1968-2003. Environmental Science & Technology . 39 , 80-84.

International Organization for Standardization, ISO/DIS 25101 ( www.iso.org )

Ismail, N., Gewurtz, S. B., Pleskach, K., Whittle, D. M., Helm, P. A., Marvin, C. H. and Tomy, G. T. (2009). BROMINATED AND CHLORINATED FLAME RETARDANTS IN LAKE ONTARIO, CANADA, LAKE TROUT (SALVELINUS NAMAYCUSH) BETWEEN 1979 AND 2004 AND POSSIBLE INFLUENCES OF FOOD-WEB CHANGES. Environmental Toxicology and Chemistry . 28 , 910-920.

Jonsson, S. and Horsing, M. (2009). Investigation of sorption phenomena by solid phase extraction and liquid chromatography for the determination of some ether derivatives of tetrabromobisphenol A. Journal of Physical Organic Chemistry . 22 , 1120-1126.

Julander, A., Westberg, H., Engwall, M. and van Bavel, B. (2005). Distribution of brominated flame retardants in different dust fractions in air from an electronics recycling facility. Science of the Total Environment . 350 , 151-160.

Karlsson, M., Julander, A., van Bavel, B. and Hardell, L. (2007). Levels of brominated flame retardants in blood in relation to levels in household air and dust. Environment International . 33 , 62-69.

Hoh, E., Zhu, L. and Hites, R.A. (2005) Novel flame retardants, 1,2-Bis(2,4,6- tribromophenoxy)ethane and 2,3,4,5,6-pentabromoethylbenzene, in United States' environmental samples. Environ. Sci. Technol., 39, 2472-2477.

Hu G.C., Luo X.J., Dai J.Y., Zhang X.L., Wu H., Zhang C.L., Guo W., Xu M.Q., Mai B.X., Wei F.W. (2008) Brominated Flame Retardants, Polychlorinated Biphenyls, and Organochlorine Pesticides in Captive Giant Panda (Ailuropoda melanoleuca) and Red Panda (Ailurus fulgens) from China. Environ. Sci. Technol.; 42 , 4704-4709.

97 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Kallenborn, R., Berger, U., Järnberg, U. et al. 2004. Perfluorinated alkylated substances (PFAS) in the Nordic environment. TemaNord 2004:552.

Karlsson, M., Ericson, I., van Bavel, B., Jensen, J.K., Dam, M. (2006) Levels of brominated flame retardants in Northern Fulmar (Fulmarus glacialis) eggs from the Faroe Islands. Sci. Total Environ., 367, 840-846.

Karlsson, M., Julander, A., van Bavel, B., Hardell, L. (2007) Levels of brominated flame retardants in blood in relation to levels in household air and dust. Environ. Int., 33, 62-69.

Kawamoto, K. and Kuramochi, H. (2007) Fate assessment of brominated organic compounds including flame retardants based on physico-chemical parameters experimentally measured. Prepr. Ext. Abstr. ACS Natl. Meet., Am. Chem. Soc., Div. Environ. Chem., 47, 440-444. 2007.

Kemikalieinspektionen/Swedish Chemical Inspectorate. 2006. Hexabromocyklododekan (HBCDD) och tetrabrombisfenol-A (TBBPA). Rapport nr 3/06.

Kierkegaard, A., Bjoerklund, J. and Friden, U. (2004) Identification of the flame retardant decabromodiphenyl ethane in the environment. Environ. Sci. Technol., 38, 3247-3253.

KLIF (2004). Kartlegging av utvalgte nye organiske miljøgifter -bromerte flammehemmere, klorerte parafiner, bisfenol A og triclosan. Prepared by Fjeld, E., Schlabach, M., Berge, J. A., Eggen, T., Snilsberg, P., Källberg, G., Rognerud, S., Enge, E. K., Borgen, A.and Gundersen, H. Report TA-2006/2004, Oslo. (http://www.klif.no/publikasjoner/overvaking/2096/ta2096.pdf )

KLIF (2005a). Kartlegging av utvalgte nye organiske miljøgifter 2004. Bromerte flammehemmere, perfluoralkylstoffer, irgarol, diuron, BHT og dicofol. Prepared by Fjeld, E., Schlabach, M., Berge, J. A., Green, N. W., Eggen, T., Snilsberg, P., Vogelsang, C., Rognerud, S., Kjellberg, G., Enge, E. K., Dye, C.and Gundersen, H. Report TA-2096/2005, (http://www.klif.no/publikasjoner/overvaking/2096/ta2096.pdf )

KLIF (2005b). PerFluorAlkylStoffer (PFAS) og PerFluorOktanylSulfonat (PFOS) – relaterte forbindelser. Prepared by Report (http://www.klif.no/arbeidsomr/kjemikalier/pfos/pfas_handlingsplan.pdf )

KLIF (2007a). Kartlegging av metaller og utvalgte nye organiske miljøgifter 2006. Prepared by Bakke, T., Fjeld, E., Skaare, B. B., Berge, J. A., Green, N. W., Ruus, A., Schlabach, M.and Botnen, H. Report TA-2284/2007, ( www.klif.no/publikasjoner/2284/ta2284.pdf )

KLIF (2007b). PFOA in Norway. Prepared by Report (http://www.klif.no/publikasjoner/2354/ta2354.pdf )

KLIF (2007c). REVIDERING AV KLASSIFISERING AV METALLER OG ORGANISKE MILJØGIFTER I VANN OG SEDIMENTER. Prepared by KLIF. Report TA-2229/2007, Oslo. ( http://www.klif.no/publikasjoner/2229/ta2229.pdf )

98 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

KLIF (2008a). Ban on Deca-BDE. Prepared by Report TA-2401/2008, (www.klif.no/publikasjoner/2401/ta2401.pdf )

KLIF (2008b). Nye miljøgifter i sedimentprøver indre Oslofjord. Prepared by Breedveld, G. D.and Arp, H. P. H. Report TA-2479, ( http://www.klif.no/publikasjoner/2479/ta2479.pdf )

KLIF (2008c). Polybrominated diphenyl ethers and perfluorinated compounds in the Norwegian environment. Prepared by Økland, T. E.and Skoog, K. Report TA-2450/2008, Oslo. (www.klif.no/publikasjoner/2450/ta2450.pdf )

KLIF (2008d). SCREENING OF POLYFLUORINATED ORGANIC COMPOUNDS AT FOUR FIRE TRAINING FACILITIES IN NORWAY. Prepared by Report TA- 2444/2008, (http://www.klif.no/publikasjoner/2444/ta2444.pdf )

KLIF (2008e). SFTS ARBEID MED PERFLUORERTE FORBINDELSER 2008-2009. Prepared by Report TA-2395/2008, ( http://www.klif.no/publikasjoner/2395/ta2395.pdf )

KLIF (2009a). Current State of Knowledge and Monitroing requirements. Emerging “new” brominated flame retardants in flame retarded products and the environment. Prepared by Harju, M., Heimstad, E. S., Herzke, D., Sandanger, T., Posner, S.and Wania, F. Report TA- 2462/2009, ( www.klif.no/publikasjoner/2462/ta2462.pdf )

KLIF (2009b). Handlingsplan for reduksjon av utslipp av bromerte flammehemmere. Oppdatert november 2009. Prepared by Report (www.klif.no/upload/.../bromerteflammehemmere_handlingsplan2009.pdf )

KLIF (2009c). Tilstandsklasser for forurenset grunn. Prepared by KLIF. Report TA- 2553/2009, ( www.klif.no/publikasjoner/2553/ta2553.pdf )

Klosterhaus, S., Konstantinov, A. and Stapleton, H. M. (2008) Characterization of the brominated chemicals in a pentaBDE replacement mixture and their detection in biosolids collected from two San Francisco bay area waste water treatment plants. Oakland, San Fransisco Estuary Institute. URL: http://www.sfei.org/rmp/posters/08BFR_Poster_klosterhaus_shrunk.pdf

Knudsen, G.A., Jacobs, L.M., Kuester, R.K. and Sipes, I.G. (2007) Absorption, distribution, metabolism and excretion of intravenously and orally administered tetrabromobisphenol A [2,3-dibromopropyl ether] in male Fischer-344 rats. Toxicology, 237, 158-167.

Koss, G., Doring, H., Stremme, C., Seubert, S., Seubert, A., Arndt, B., Seidel, J., Koransky, W. and Ippen, H. (1986) Elimination of porphyrins and their precursors in rats pretreated with hexabromobenzene. Food Chem. Toxicol., 24 , 325-328.

La Guardia, M. J., Hale, R. C. and Harvey, E. (2007). Evidence of debromination of decabromodiphenyl ether (BDE-209) in biota from a wastewater receiving stream. Environmental Science & Technology . 41 , 6663-6670.

99 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Lam, J. C. W., Lau, R. K. F., Murphy, M. B. and Lam, P. K. S. (2009). Temporal Trends of Hexabromocyclododecanes (HBCDs) and Polybrominated Diphenyl Ethers (PBDEs) and Detection of Two Novel Flame Retardants in Marine Mammals from Hong Kong, South China. Environmental Science & Technology . 43 , 6944-6949.

Lau, C., Anitole, K., Hodes, C., Lai, D., Pfahles-Hutchens, A. and Seed, J. (2007). Perfluoroalkyl acids: A review of monitoring and toxicological findings. Toxicological Sciences . 99 , 366-394.

Lau, C., Butenhoff, J. L. and Rogers, J. M. (2004). The developmental toxicity of perfluoroalkyl acids and their derivatives. Toxicology and Applied Pharmacology . 198 , 231- 241.

Law, R. J., Herzke, D., Harrad, S., Morris, S., Bersuder, P. and Allchin, C. R. (2008). Levels and trends of HBCD and BDEs in the European and Asian environments, with some information for other BFRs. Chemosphere . 73 , 223-241.

Law, K., Halldorson, T., Danell, R., Stern, G., Gewurtz, S., Alaee, M., Marvin, C., Whittle, M. and Tomy, G. (2006) Bioaccumulation and trophic transfer of some brominated flame retardants in a Lake Winnipeg (Canada) food web. Environ. Toxicol. Chem., 25, 2177-2186.

Law, K., Halldorson, T., Danell, R., Stern, G., Gewurtz, S., Alaee, M., Marvin, C., Whittle, M. and Tomy, G. (2007) Bioaccumulation and trophic transfer of some brominated flame retardants in a Lake Winnipeg (Canada) food web. [Erratum to document cited in CA145:277560]. Environ. Toxicol. Chem., 25, 190.

Lee, R.G.M., Thomas, G.O., and Jones, K.C. (2002) Atmospheric concentrations of PBDEs in western Europe. Organohal. Comp., 58, 193-196.

Leung, A. O. W., Luksemburg, W. J., Wong, A. S. and Wong, M. H. (2007). Spatial distribution of polybrominated diphenyl ethers and polychlorinated dibenzo-p-dioxins and dibenzofurans in soil and combusted residue at Guiyu, an electronic waste recycling site in southeast China. Environmental Science & Technology . 41 , 2730-2737.

Li, X., Yang, L., Liu, E. and Xu, W. (2004) Research on risk characteristics of decabrominated diphenyl ethane. Prog. Saf. Sci. Technol., Proc. Int. Symp., 4, pt. B, 2164- 2166.

Loveless, S.E., Slezak, B., Serex, T., Lewis, J., Mukerji, P., O’Connor, J.C., Donner, E.M., Frame, S.R., Korzeniowski, S.H., Buck, R.C. 2009. Toxicological evaluation of sodium perfluorohexanoate. Toxicology, 264:32-44.

Lui, W., Chen, S., Quan, X., Jin, Y.-H. 2008. Toxic effect of serial perfluorosulfonic and perfluorocarboxylic acids on the membrane system of a fresh water algae measured by flow cytometry. Environmental toxicology and chemistry. 27:1597-1604.

Luo, X. J., Zhang, X. L., Liu, J., Wu, J. P., Luo, Y., Chen, S. J., Mai, B. X. and Yang, Z. Y. (2009). Persistent Halogenated Compounds in Waterbirds from an e-Waste Recycling Region in South China. Environmental Science & Technology . 43 , 306-311.

100 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Martin, J.W., Mabury, S.A., Solomon, K.R., Muir, D.C.G. 2003. Bioconcentration and tissue distribution of perfluorinated acids in rainbow trout ( Oncorhynchus mykiss ). Environmental Toxicology & Chemistry, 22:196-204.

Melcher, J., Schlabach, M., Andersen, M. S. and Vetter, W. (2008). Contrasting the seasonal variability of halogenated natural products and anthropogenic hexachlorocyclohexanes in the southern Norwegian atmosphere. Archives of Environmental Contamination and Toxicology . 55 , 547-557.

Mendoza, C.E., Collins, B.T., Shields, J.B., Laver, G.W. (1978) Effects of hexachlorobenzene or hexabromobenzene on body and organ weights of preweanling rats after a reciprocal transfer between treated and control dams. J. Agric. Food Chem., 26, 941-945.

Mendoza, C.E., Shields, J.B., Laver, G.W. (1979) Comparison of the porphyrinogenic activity of hexabromobenzene and hexachlorobenzene in primiparous wistar rats. Bull. Environ. Contam. Toxicol., 21, 358-364.

Minh, N. H., Isobe, T., Ueno, D., Matsumoto, K., Mine, M., Kajiwara, N., Takahashi, S. and Tanabe, S. (2007). Spatial distribution and vertical profile of polybrominated diphenyl ethers and hexabromocyclododecanes in sediment core from Tokyo Bay, Japan. Environmental Pollution . 148 , 409-417.

Moody, C.A., Martin, J.W., Kwan, W.C., Muir, D.C.G., Mabury, S.A. 2002. Monitoring perfluorinated surfactants in biota and surface water samples following and accidental release of fire-fighting foam into Etobicoke Creek. Environmental Science & Technology (36) 545- 551.

Nakari, T. and Huhtala, S. (2009). In Vivo and In Vitro Toxicity of Decabromodiphenyl Ethane, a Flame Retardant. Environmental Toxicology . DOI 10.1002/tox.20499 .

Newsted, J.L., Beach, S.A., Gallagher, S.P., Giesy, J.P. 2008. Acute and chronic effects of perfluorobutane sulfonates (PFBS) on the mallard and northern bobwhite quail. Archives of environmental contamination and toxicology 54:535-545.

Niimi, A. J. and Oliver, B. G. (1988) Influence of molecular weight and molecular volume on dietary absorption efficiency of chemicals by fishes. Can. J. Fish. Aquat. Sci., 45, 222-227.

Nyholm JR, Lundberg C, Andersson PL. Biodegradation kinetics of selected brominated flame retardants in aerobic and anaerobic soil.Environmental Pollution in press .

Nyholm, J. R., Norman, A., Norrgren, L., Haglund, P. and Andersson, P. L. (2009). UPTAKE AND BIOTRANSFORMATION OF STRUCTURALLY DIVERSE BROMINATED FLAME RETARDANTS IN ZEBRAFISH (DANIO RERIO) AFTER DIETARY EXPOSURE. Environmental Toxicology and Chemistry . 28 , 1035-1042.

101 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Oeberg, T., Warman, K. and Bergstroem, J. (1987) Brominated aromatics from combustion. Chemosphere, 16, 2451-2465.

Oliver, B.G. and Niimi, A.J. (1985) Bioconcentration factors of some halogenated organics for rainbow trout: limitations in their use for prediction of environmental residues. Environ. Sci. Technol., 19, 842-849.

Oliveira, A. S., Silva, V. M., Veloso, M. C. C., Santos, G. V. and De Andrade, J. B. (2009). Bromophenol concentrations in fish from Salvador, BA, Brazil. Anais Da Academia Brasileira De Ciencias . 81 , 165-172.

Pakalin, S., Cole, T., Steinkellner, J., Nicolas, R., Tissier, C., Munn, C., and Eichenreich, S. (2007). Review on production processes of decabromodiphenyl ether (DecaBDE) used in polymeric applications in electrical and electronic equipment, and assessment of the availability of potential alternatives to DecaBDE). European Chemicals Bureau, Institute of Health and Consumer Protection, Joint Research Centre, European Commission. (EUR 22693 EN). Pettersson-Julander, A., Van Bavel, B., Engwall, M., and Westberg, H. (2004) Personal air sampling and analysis of polybrominated diphenyl ethers and other bromine containing compounds at an electronic recycling facility in Sweden. J. Environ. Monit., 6, 874-880.

Pan, G., Jia, C. X., Zhao, D. Y., You, C., Chen, H. and Jiang, G. B. (2009). Effect of cationic and anionic surfactants on the sorption and desorption of perfluorooctane sulfonate (PFOS) on natural sediments. Environmental Pollution . 157 , 325-330.

Paul, A. G., Jones, K. C. and Sweetman, A. J. (2009). A First Global Production, Emission, And Environmental Inventory For Perfluorooctane Sulfonate. Environmental Science & Technology . 43 , 386-392.

Phillips, M. M., Dinglasan-Panlilio, M. J. A., Mabury, S. A., Solomon, K. R. and Sibley, P. K. (2007). Fluorotelomer acids are more toxic than perfluorinated acids. Environmental Science & Technology . 41 , 7159-7163.

Prevedouros, K., Cousins, I. T., Buck, R. C. and Korzeniowski, S. H. (2006). Sources, fate and transport of perfluorocarboxylates. Environmental Science & Technology . 40 , 32-44.

Potrzebowksi and Chance, 2004), mangler

Prevedouros, K., Cousins, I. T., Buck, R. C. and Korzeniowski, S. H. (2006). Sources, fate and transport of perfluorocarboxylates. Environmental Science & Technology . 40 , 32-44.

Psillakis, E., Cheng, J., Hoffmann, M. R. and Colussi, A. J. (2009). Enrichment Factors of Perfluoroalkyl Oxoanions at the Air/Water Interface. Journal of Physical Chemistry A . 113 , 8826-8829.

Rayne, S. and Forest, K. (2009a). Comment on "Comparative Assessment of the Global Fate and Transport Pathways of Long-Chain Perfluorocarboxylic Acids (PFCAs) and

102 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Perfluorocarboxylates (PFCs) Emitted from Direct Sources". Environmental Science & Technology . 43 , 7155-7156.

Rayne, S. and Forest, K. (2009b). A comparative assessment of octanol-water partitioning and distribution constant estimation methods for perfluoroalkyl carboxylates and sulfonates. Nature Precedings . doi:10.1038/npre.2009.3282.2 .

Rayne, S. and Forest, K. (2009c). Perfluoroalkyl sulfonic and carboxylic acids: A critical review of physicochemical properties, levels and patterns in waters and wastewaters, and treatment methods. Journal of Environmental Science and Health Part a-Toxic/Hazardous Substances & Environmental Engineering . 44 , 1145-1199.

Rayne, S., Forest, K. and Friesen, K. J. (2009). Extending the semi-empirical PM6 method for carbon oxyacid pKa prediction to sulfonic acids: Application towards congener-specific estimates for the environmentally and toxicologically relevant C1 through C8 perfluoroalkyl derivatives. Nature Precedings . hdl:10101/npre.2009.2922.1 .

Renner, R. 2006. The long and the short of perfluorinated replacement. Environmental Science & Technology, 40:12-13.

Qiu, X., Marvin, C.H. and Hites, R.A. (2007) Dechlorane plus and other flame retardants in a sediment core from Lake Ontario. Environ. Sci. Technol., 41, 6014-6019.

Qiu, X. and Hites, R.A. (2008) Dechlorane plus and other flame retardants in tree bark from the northeastern United States. Environ. Sci. Technol., 42, 31-36.

Rahm, S., Green, N., Norrgran, J. and Bergman, A. (2005) Hydrolysis of environmental contaminants as an experimental tool for indication of their persistency. Environ. Sci. Technol., 39, 3128-3133.

Renner, R. (2006). The long and the short of perfluorinated replacements. Environmental Science & Technology . 40 , 12-13.

Ricklund, N., Kierkegaard, A. and McLachlan, M. S. (2008). An international survey of decabromodiphenyl ethane (deBDethane) and decabromodiphenyl ether (decaBDE) in sewage sludge samples. Chemosphere . 73 , 1799-1804.

Rosen, M. B., Lau, C. and Corton, J. C. (2009). Does Exposure to Perfluoroalkyl Acids Present a Risk to Human Health? Toxicological Sciences . 111 , 1-3.

RTECS (2008) Registry of Toxic Effects of Chemical Substances. National Institute for Occupational Safety and Health, USA. URL: http://www.cdc.gov/niosh/rtecs/

Sawal, G., Feibicke, M., Meinecke, S., Warmbrunn-Suckrow, E. and Lepom, P. (2008) Identification of a novel brominated flame retardant in dust. Organohal. Comp., 70, 2029- 2032

103 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Schenker, U., Scheringer, M., Macleod, M., Martin, J. W., Cousins, I. T. and Hungerbuhlert, K. (2008a). Contribution of volatile precursor substances to the flux of perfluorooctanoate to the arctic. Environmental Science & Technology . 42 , 3710-3716.

Schenker, U., Soltermann, F., Scheringer, M. and Hungerbuhler, K. (2008b). Modeling the Environmental Fate of Polybrominated Diphenyl Ethers (PBDEs): The Importance of Photolysis for the Formation of Lighter PBDEs. Environmental Science & Technology . 42 , 9244-9249.

SFT, 2007. A litterature survey on selected chemical substances. TA-2238/2007. 112 pp.

SFT. 2007. Guidelines on classification of environmental quality in fjords and coastal waters – A revision of the classification of water and sediments with respect to metals and organic contaminants (in Norwegian, English summary). Norwegian State Pollution Control Agency, report TA-2229.

SFT, 2005. Mapping of PFOS in fire fighting foam. TA-2139/2005. 20 pp.

Shi, T., Chen, S. J., Luo, X. J., Zhang, X. L., Tang, C. M., Luo, Y., Ma, Y. J., Wu, J. P., Peng, X. Z. and Mai, B. X. (2009). Occurrence of brominated flame retardants other than polybrominated diphenyl ethers in environmental and biota samples from southern China. Chemosphere . 74 , 910-916.

Sijm, D.T.H.M., Part, P. and Opperhuizen, A. (1993) The influence of temperature on the uptake rate constants of hydrophobic compounds determined by the isolated perfused gills of rainbow-trout (Oncorhynchus-Mykiss). Aquat. Toxicol., 25, 1-14.

Sijm, D.T.H.M., Verberne, M.E., Dejonge, W.J., Par, P. and Opperhuizen, A. (1995) allometry in the uptake of hydrophobic chemicals determined in-vivo and in isolated- perfused gills. Toxicol. Appl. Pharmacol., 131, 130-135.

Simonsen, F. A., Stavnsbjerg, M., Moller, L.M., and Madsen, T. (2000) Brominated flame retardants; toxicity and ecotoxicity. København, Miljøstyrelsen (Environmental Project No. 568) URL: http://www2.mst.dk/udgiv/publications/2000/87-7944-288-9/pdf/87-7944-289- 7.pdf

Sinkkonen, S., Paasivirta, J., Lahtipera, M. and Vattulainen, A. (2004) Screening of halogenated aromatic compounds in some raw material lots for an aluminium recy cling plant. Environ. Internat., 30, 363-366.

Sjodin, A., Carlsson, H., Thuresson, K., Sjoelin, S., Bergman, Aa. and Oestman, C. (2001) Flame retardants in indoor air at an electronics recycling plant and at other work environments. Environ. Sci. Technol., 35, 448-454.

Smith, A.G. and Francis, J.E. (1980) Relative abilities on a molar basis of

104 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) hexafluoro-, hexachloro- and hexabromobenzenes to decrease liver uroporphyrinogen decarboxylase activity and cause porphyria in female rats. Res. Commun. Chem. Pathol. Pharmacol., 28, 377-384.

Stapleton, H. M., Klosterhaus, S., Eagle, S., Fuh, J., Meeker, J. D., Blum, A. and Webster, T. F. (2009). Detection of Organophosphate Flame Retardants in Furniture Foam and U.S. House Dust. Environmental Science & Technology . 43, 7490-7495.

Stapleton, H.M., Allen, J.G., Kelly, S.M., Konstantinov, A., Klosterhaus, S., McClean, M.D. and Webster, T.F. (2008a) Alternate and new brominated fame retardants detected in U.S. house dust. Organohal. Comp., 70, 438-441.

Stapleton, H. M., Kelly, S. M., Allen, J. G., Watkins, D. J., Heiger-Bernays, W. J., McClean, M. D., Webster, T. F., Konstantinov, A. and Klosterhaus, S. (2008). Response to Comment on "Alternate and New Brominated Flame Retardants Detected in U.S. House Dust". Environmental Science & Technology . 42 , 9455-9456.

Stock, N. L., Muir, D. C. G. and Mabury, S. A. (2010). Perfluoroalkyl Compounds. In Persistent Organic Pollutants , ed. Harrad, S., Wiley. Chippenham.25-67.

Stebberud, H. 2006. Økotokikologiske effecter av PFOS, PFOA og 6:2 FTS på meitemark (Eisenia fetida ). (in Norwegian, English summary). Norwegian State Pollution Control Agency, report TA-2212.

Szymanska, J. A., Bruchajzer, E., Sporny, S. and Piotrowski, J. K. (1998) Changes in selected indicator of liver impairment after repeated administration of mono- and polybromobenzenes in mice. Bull. Environ. Contam. Toxicol., 61, 22-30.

Szymanska, J.A. and Piotrowski, J.K. (2000) Hepatotoxicity of monobromobenzene and hexabromobenzene: effects of repeated dosage in rats. Chemosphere, 41, 1689-1696.

Szymanska, J.A., Czerski, J. and Frydrych, B. (2002) Tissue concentrations of porphyrins: effects of repeated dosage of selected polybromobenzenes in rats. Organohal. Comp., 56, 1- 4.

Takigami, H., Suzuki, G., Hirai, Y. and Sakai, S. (2009). Brominated flame retardants and other polyhalogenated compounds in indoor air and dust from two houses in Japan. Chemosphere . 76 , 270-277.

Thomsen, C., Lundanes, E., and Becher, G. (2001) Brominated flame retardants in plasma samples from three different occupational groups in Norway. J. Environ. Monit. 3 , 366-370.

Thomsen, C., Lundanes, E., and Becher, G. (2002) Brominated Flame Retardants in Archived Serum Samples from Norway: A Study on Temporal Trends and the Role of Age. Environ. Sci. & Technol. 36 , 1414-1418.

Thomsen, C., Liane, V. H. and Becher, G. (2007) Automated solid-phase extraction for the determination of polybrominated diphenyl ethers and polychlorinated biphenyls in serum-

105 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) application on archived Norwegian samples from 1977 to 2003. J. Chromatogr., B: Anal. Technol. Biomed. Life Sci. 846 , 252-263.

Tittlemier, S.A., Halldorson, T., Stern, G.A. and Tomy, G.T. (2002) Vapor pressures, aqueous , and Henry's law constants of some brominated flame retardants. Environ. Toxicol. Chem., 21, 1804-1810.

Tolosa, I., Bayona, J. M. and Albaiges, J. (1991) Identification and occurrence of brominated and nitrated phenols in estuarine sediments. Mar. Pollut. Bull., 22, 603-607.

Tomy, G.T., Palace, V.P., Pleskach, K., Ismail, N., Oswald, T., Danell, R.t, Wautier, K., and Evans, B. (2007) Dietary exposure of juvenile rainbow trout (Oncorhynchus mykiss) to 1,2- bis(2,4,6-tribromo- phenoxy)ethane: bioaccumulation parameters, biochemical effects, and metabolism. Environ. Sci. Technol., 41, 4913-4918.

TOXNET (2008) Databases on toxicology, hazardous chemicals, environmental health, and toxic releases. U.S. National Library of Medicine (NLM). URL: http://toxnet.nlm.nih.gov .

UNEP (2009). SC-4/19: Establishing indicative elements of a work programme to facilitate the elimination of listed brominated diphenyl ethers and the restriction or elimination of perfluorooctane sulfonic acid and its salts, perfluorooctane sulfonyl fluoride and other chemicals listed in Annexes A or B of the Convention at the fourth meeting of the Conference of the Parties. Prepared by Report (http://chm.pops.int/Programmes/NewPOPs/Publications/tabid/695/language/en- US/Default.aspx )

US EPA (2004b) HPV data summary and test plan for 1H-Isoindole-1,3-(2H)-dione, 2,2'- (1,2-ethanediyl)bis(4,5,6,7-tetrabromo--). Prepared by Albemarle Corporation, Baton Rouge, LA. (Report 201-15090. 201-15090), pp. 1-11.

Venier, M. and Hites, R.A. (2008) Flame retardants in the atmosphere near the Great Lakes. Environ. Sci. Technol., 42, 4745-4751.

Verreault, J., Gebbink, W.A., Gauthier, L.T., Gabrielsen, G.W., and Letcher, R.J. (2007) Brominated flame retardants in Glaucous Gulls from the Norwegian Arctic: more than just an issue of polybrominated diphenyl ethers. Environ. Sci. Technol., 41, 4925-4931.

Vetter, W., Haase-Aschoff, P., Rosenfelder, N., Komarova, T. and Mueller, J. F. (2009). Determination of Halogenated Natural Products in Passive Samplers Deployed along the Great Barrier Reef, Queensland/Australia. Environmental Science & Technology . 43 , 6131- 6137.

106 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Vetter, W. and Janussen, D. (2005). Halogenated natural products in five species of antarctic sponges: Compounds with POP-like properties? Environmental Science & Technology . 39 , 3889-3895.

Vetter, W., von der Recke, R., Ostrowicz, P. and Rosenfelder, N. (2010). Liquid chromatographic enantioseparation of the brominated flame retardant 2,3-dibromopropyl- 2,4,6-tribromophenyl ether (DPTE) and enantiomer fractions in seal blubber. Chemosphere . 78 , 134-138.

Villeneuve, D.C. and Khera, K.S. (1975) Placental-transfer of halogenated (pentachlorobenzene, pentachloronitrobenzene, and hexabromobenzene). Environ. Physiol. Biochem., 5, 328-331.

Von Recke, R. and Vetter, W. (2007) Synthesis and characterization of 2,3-dibromopropyl- 2,4,6-tribromophenyl ether (DPTE) and structurally related compounds evidenced in seal blubber and brain. Environ. Sci. Technol., 41, 1590-1595.

Vonderheide, A. P., Mueller, K. E., Meija, J. and Welsh, G. L. (2008). Polybrominated diphenyl ethers: Causes for concern and knowledge gaps regarding environmental distribution, fate and toxicity. Science of the Total Environment . 400 , 425-436.

Wallington, T. J., Hurley, M. D., Xia, J., Wuebbles, D. J., Sillman, S., Ito, A., Penner, J. E., Ellis, D. A., Martin, J., Mabury, S. A., et al. (2006). Formation of C7F15COOH (PFOA) and other perfluorocarboxylic acids during the atmospheric oxidation of 8 : 2 fluorotelomer alcohol. Environmental Science & Technology . 40 , 924-930.

Wania, F. (2007). A global mass balance analysis of the source of perfluorocarboxylic acids in the Arctic ocean. Environmental Science & Technology . 41 , 4529-4535.

Watanabe, I. and Sakai, S. (2003). Environmental release and behavior of brominated flame retardants. Environment International . 29 , 665-682.

Watanabe, I., Kashimoto, T. and Tatsukawa, R. (1985) Brominated phenols and anisoles in river and marine sediments in Japan. Bull. Environ. Contam. Toxicol., 35, 272-278.

Watanabe, I., Kashimoto, T. and Tatsukawa, R. (1986) Hexabromobenzene and its debrominated compounds in river and estuary sediments in Japan. Bull. Environ. Contam. Toxicol., 36, 778-784.

Weiss, J. M., Andersson, P. L., Lamoree, M. H., Leonards, P. E. G., van Leeuwen, S. P. J. and Hamers, T. (2009). Competitive Binding of Poly- and Perfluorinated Compounds to the Thyroid Hormone Transport Protein Transthyretin. Toxicological Sciences . 109 , 206-216.

Whitfield, F.B., Helidoniotis, F., Svoronos, D., Shaw, K.J., and Ford, G. L. (1995) The source of bromophenols in some species of Australian ocean fish. Water Sci. Technol., 31, 113-120.

107 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Whitfield, F.B., Shaw, K.J., Walker, D.I. (1992) The source of 2,6-dibromophenol - cause of an iodoform taint in Australian prawns. Water Sci. Technol., 25, 131-138.

Whitfield, F.B., Helidoniotis, F., Shaw, K.J. and Svoronos, D. (1997) Distribution of bromophenols in Australian wild-harvested and cultivated prawns (Shrimp). J. Agric. Food Chem., 45, 4398-4405.

WHO, 1995. Tetrabromobisphenol A and derivatives. Geneve, World Health Organization (Environmental health criteria 172).

WHO (1994). Polybrominated biphenyls. International programme on chemical safety. . Prepared by Report Environmental Health Criteria 152, Geneva. (http://www.inchem.org/documents/ehc/ehc/ehc152.htm )

Woldegiorgis, A., Andersson, J., Remberger, M., Kaj, L., Ekheden, Y., Blom, L., Brorström- Lundén, E. 2006. Results from the Swedish National Screening Programme 2005. IVL Report B1698.

Yamaguchi, Y., Kawano, M. and Tatsukawa, R. (1988a). TISSUE DISTRIBUTION AND EXCRETION OF HEXABROMOBENZENE AND ITS DEBROMINATED METABOLITES IN THE RAT. Archives of Environmental Contamination and Toxicology . 17 , 807-812.

Yamaguchi, Y., Kawano, M., Tatsukawa, R. and Moriwaki, S. (1988b). HEXABROMOBENZENE AND ITS DEBROMINATED COMPOUNDS IN HUMAN ADIPOSE TISSUES OF JAPAN. Chemosphere . 17 , 703-707.

York, R.G. 2003. Oral (gavage) combined repeated dose toxicity study of T-7706 with the reproduction/developmental toxicity screening test. 3M, US EPA AR-226-1523.

Yuan, Q., Ravikrishna, R. and Valsaraj, K. T. (2001). Reusable adsorbents for dilute solution separation. 5. Photodegradation of organic compounds on surfactant-modified titania. Separation and Purification Technology . 24 , 309-318.

Zitko, V. (1977) Uptake and excretion of chlorinated and brominated hydrocarbons by fish. Biological Station, St. Andrews, New Brunswick, Fish Marine Service. (Technical Report, 737).

Zitko, V. and Hutzinger, O. (1976) Uptake of chlorobiphenyls and bromobiphenyls, hexachloromobenzene and hexabromobenzene by fish. Bull. Environ. Contam. Toxicol., 16, 665-673.

Zhu, L. Y. and Hites, R. A. (2006). Brominated flame retardants in tree bark from North America. Environmental Science & Technology . 40 , 3711-3716.

Zhu, L. Y., Ma, B. L. and Hites, R. A. (2009). Brominated Flame Retardants in Serum from the General Population in Northern China. Environmental Science & Technology . 43 , 6963- 6968.

108 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

9. Appendix I – Analytical results

Table X. Measured concentrations of BFRs in sediment samples (ng/g dw) PBT PBEB HBB BTBPE DBDPE DPTE TBPA TBP ATE TBBPA- BTBPI EHTBB TBBPA- TBPH AE DBPE Tromsö, SAS-1b 0.01 n.d. 0.01 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-1c n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-1c n.d. n.d. n.d. n.d. n.d. n.d. n.d. 3.3 n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-4b n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-4b n.d. n.d. 0.02 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-4b n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-6b n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-6b n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö, SAS-6b n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj, Drsol-01 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj, Drsol-01 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj, Drsol-01 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj, Drsol-02 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj, Drsol-02 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj, Drsol-02 n.d. n.d. n.d. 2.7 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj, Drsol-03 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. 0.5 n.d. n.d. n.d. n.d. Drammensfj, Drsol-03 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. 0.6 n.d. n.d. n.d. n.d. Drammensfj, Drsol-03 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. 1.3 n.d. n.d. n.d. n.d. Mjösa, MLI-1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. 2.2 n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-3 n.d. n.d. 0.02 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa, MLI-3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Losna n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Losna n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Losna n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.

109 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Hellik Teigen 0.02 0.004 0.1 1.7 0.5 n.d. n.d. n.d. n.d. 0.5 n.d. n.d. n.d. n.d. Hokksund Hellik Teigen 0.03 0.1 0.1 0.6 1.3 n.d. n.d. n.d. n.d. 0.3 n.d. n.d. n.d. n.d. Hokksund Hellik Teigen 0.04 0.01 0.1 1.7 1.8 n.d. n.d. n.d. n.d. 0.3 n.d. n.d. n.d. n.d. Hokksund Lindum Resurs och 0.3 n.d. 1.8 4.5 n.d. n.d. n.d. n.d. n.d. 2.4 n.d. n.d. n.d. n.d. Gjenvinning Lindum Resurs och 0.1 n.d. 0.4 4.4 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Gjenvinning Lindum Resurs och 0.2 n.d. 1.8 3.1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Gjenvinning

LOD 0.01 0.00 0.29 0.01 0.10 0.54 0.04 0.07 0.07 0.01 1.16 0.16 0.05 0.39 LOQ 0.02 0.01 0.9 7 0.02 0.32 1.79 0.13 0.24 0.23 0.02 3.85 0.52 0.18 1.29

Table X. Measured concentrations of BFRs in water samples (ng/L) PBT PBEB HBB BTBPE DBDPE DPTE TBPA TBP ATE TBBPA- BTBPI EHTBB TBBPA- TBPH AE DBPE Tromsö, incoming n.d. n.d. 2.8 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. wastewater Tromsö, outgoing n.d. n.d. 1.4 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. water Tromsö, incoming n.d. n.d. 2.4 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. 18 n.d. wastewater Tromsö, outgoing n.d. n.d. 1.2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. water Tromsö, incoming n.d. n.d. 0.2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. wastewater Tromsö, outgoing n.d. n.d. 0.1 n.d. n.d. n.d. n.d. n.d. n.d. 0.46 n.d. n.d. n.d. n.d. water Drammen n.d. n.d. 0.5 n.d. 1.17 n.d. n.d. n.d. n.d. 0.06 n.d. n.d. n.d. n.d. Drammen n.d. n.d. 0.5 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen n.d. n.d. 1.2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen 0.17 n.d. 0.7 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.

110 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Drammen n.d. n.d. 1.9 n.d. 9.04 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen n.d. n.d. 0.9 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammer n.d. n.d. 0.2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammer n.d. n.d. 0.4 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammer n.d. n.d. 0.3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammer n.d. n.d. 0.2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammer 0.09 n.d. 0.6 0.85 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammer n.d. n.d. 1.1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Hokksund, Seepage 4.4 0.6 11.4 61.7 15.3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. 15.8 n.d. water Hokksund, Seepage 5.0 0.9 15.7 59.1 38.7 n.d. n.d. n.d. n.d. n.d. n.d. n.d. 159.6 n.d. water Hokksund, Seepage 7.5 1.3 19.1 107.0 185.7 n.d. n.d. n.d. n.d. 2.0 35 n.d. 66.0 n.d. water Lindum, 0.4 n.d. 2.8 4.2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Seepage water Lindum, n.d. n.d. 1.8 2.1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Seepage water Lindum, n.d. n.d. 0.6 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Seepage water

LOD 0.06 0.04 0.03 0.38 1.44 0.16 0.51 1.30 1.41 0.08 0.32 0.34 0.38 0.13 LOQ 0.22 0.12 0.12 1.30 4.90 0.54 1.75 4.42 4.78 0.27 1.08 1.14 1.28 0.46

111 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table X. Measured concentrations of BFRs in sewage sludge (ng/g dw)

PBT PBEB HBB BTBPE DBDPE DPTE TBPA TBP ATE TBBPA BTBPI EHTBB TBBPA TBPH AE DBPE Langnes, n.d. n.d. 0.42 0.97 1.2 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Langnes, n.d. n.d. 0.33 n.d. 2.6 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Langnes, n.d. n.d. 0.27 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Solumstrand, n.d. n.d. 0.22 2.08 6.6 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen Solumstrand n.d. n.d. 0.6 0.5 8.7 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Solumstrand n.d. n.d. 0.38 n.d. 3.6 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammar n.d. n.d. 0.14 1.27 3.3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammar n.d. n.d. 0.14 1.36 5.0 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Lillehammar n.d. n.d. 0.12 1.38 1.3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.

LOD 0.09 0.07 0.1 0.5 1.0 0.2 0.6 1.6 1.7 0.1 0.4 0.4 0.5 0.2 LOQ 0.3 0.2 0.4 1.6 3.3 0.7 2.2 5.5 5.9 0.3 1.3 1.4 1.6 0.6

112 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table X. Measured concentrations of BFRs in fish crab and mussel (ng/g ww)

PBT PBEB HBB BTBPE DBDPE DPTE TBPA TBP ATE TBBPA- BTBPI EHTBB TBBPA- TBPH AE DBPE Drammensfj Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. 55.8 n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa Trout - liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa Trout - liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa Trout - liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Mjösa Trout - liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammensf Blue mussel n.d. n.d. n.d. n.d. n.d. n.d. n.d. 1.1 n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj Blue mussel n.d. n.d. n.d. n.d. n.d. n.d. n.d. 0.3 n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Blue mussel n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Crab n.d. n.d. n.d. n.d. n.d. n.d. n.d. 104.8 n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Crab n.d. n.d. n.d. n.d. n.d. n.d. n.d. 130.5 n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Crab n.d. n.d. n.d. n.d. n.d. n.d. n.d. 42.1 n.d. n.d. n.d. n.d. n.d. n.d. Tromsö Crab n.d. n.d. n.d. n.d. n.d. n.d. n.d. 123.9 n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj Crab n.d. n.d. n.d. n.d. n.d. n.d. n.d. 2.4 n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj Crab n.d. n.d. n.d. n.d. n.d. n.d. n.d. 8.2 n.d. n.d. n.d. n.d. n.d. n.d. Drammensfj Crab n.d. n.d. n.d. n.d. n.d. n.d. n.d. 3.0 n.d. n.d. n.d. n.d. n.d. n.d. Bergen Liver, ling n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Bergen Liver, tusk n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Bergen Liver, pollack n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Bergen Cod liver n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.

LOD 0.1 0.1 0.1 0.5 0.7 0.2 0.6 0.4 0.8 0.1 0.4 0.3 0.3 0.3 LOQ 0.3 0.2 0.2 1.7 2.2 0.7 2.1 1.3 2.9 0.3 1.3 1.1 0.9 1.1

113 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table X. Measured concentrations of BFRs in needles (ng/g ww) PBT PBEB HBB BTBPE DBDPE DPTE TBPA TBP ATE TBBPA- BTBPI EHTBB TBBPA- TBPH AE DBPE Hurum n.d. n.d. 0.02 n.d. 0.1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. energigjenvinning 1 Hurum n.d. n.d. n.d. n.d. 0.1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. energigjenvinning 2 Hurum n.d. n.d. 0.05 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. energigjenvinning 3 Hurum n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. 0.16 n.d. energigjenvinning 4

LOD 0.01 0.01 0.01 0.1 0.1 0.1 0.1 0.1 0.1 0.2 0.1 0.1 0.1 0.2 LOQ 0.04 0.03 0.04 0.4 0.3 0.2 0.2 0.3 0.2 0.7 0.2 0.4 0.4 0.8

Table X. Measured concentrations of BFRs in moss (ng/g dw) PBT PBEB HBB BTBPE DBDPE DPTE TBPA TBP ATE TBBPA- BTBPI EHTBB TBBPA- TBPH AE DBPE Hurum n.d. n.d. 0.1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. energigjenvinning 1 Hurum n.d. n.d. 0.1 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. energigjenvinning 2 Hurum n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. energigjenvinning 3 Hurum n.d. n.d. 0.05 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. energigjenvinning 4

LOD 0.04 0.03 0.06 0.3 7.6 0.7 0.7 1.0 0.7 2.1 0.6 1.4 1.3 2.6 LOQ 0.15 0.09 0.20 1.2 25.9 2.3 2.5 3.2 2.3 7.2 2.1 4.6 4.3 8.7

Table X. Measured concentrations of BFRs in indoor air (ng/m 3)

114 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

PBT PBEB DPTE HBB BTBPE DBDPE TBPA TBP ATE TBBPA- BTBPI EHTBB TBBPA- TBPH AE DBPE Drammen 1A n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen 1B n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen 1C n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen blind n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.

LOD 0.05 0.03 0.08 0.02 0.28 1.04 0.43 0.37 0.74 0.05 0.15 0.22 0.32 0.17 LOQ 0.2 0.1 0.3 0.1 0.9 3.5 1.5 1.3 2.5 0.2 0.5 0.8 1.1 0.6

Table X. Measured concentrations in outdoor air (ng/m3) PBT PBEB DPTE HBB BTBPE DBDPE TBPA TBP ATE TBBPA- BTBPI EHTBB TBBPA- TBPH AE DBPE Drammen 2A n.d. n.d. n.d. 0.0013 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen 2B n.d. n.d. n.d. 0.0014 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen 2C n.d. n.d. n.d. 0.0102 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. Drammen blind n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.

LOD 0.001 0.001 0.002 0.001 0.006 0.022 0.009 0.008 0.016 0.001 0.003 0.005 0.007 0.004 LOQ 0.003 0.002 0.006 0.002 0.020 0.076 0.032 0.027 0.055 0.003 0.011 0.016 0.024 0.012

115 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Concentration (ng/g d.w) of PFCs in soil samples from Flesland airport. Distance from platform Dry weight 6:2 FTS PFBS PFHxS PFOS PFDS PFPeA PFHxA (%) 0 m 92% 612* 0.29 6.2 273* 3.2* 0.68 2.5 10 m 73% 2101* 1.4 17 1905* 54* 7.0 8.1 20 m 67% 751* 1.8 21 705* 8.5* 17 10.2 30 m 78% 329* 1.0 15 331* 4.0* 28 18.5 40 m 64% 38 0.67 5.2 129 2.8 17 11.5 50 m 67% 5.2 0.28 5.3 96 1.1 11 6.8 75 m 71% 7.7 <0.11 1.1 24 0.36 8.8 3.6 100 m 80% 9.2 <0.11 0.54 6.1 0.19 3.7 1.5 150 m 57% 2.5 <0.11 0.64 7.6 0.13 0.88 0.32 200 m 58% 0.84 <0.11 0.12 1.6 <0.10 0.52 0.18

Distance from platform PFHpA PFOA PFNA PFDA PFUnDA PFDoDA PFTrDA PFTeDA 0 m 0.80 1.4 <1.0 NQ 0.94 0.58 <0.1 <0.1 10 m 4.5 7.7 NQ NQ 5.8* 3.77* 0.25* <0.1 20 m 2.8 4.4 NQ NQ 1.6* NQ <0.1 <0.1 30 m 4.6 4.4 1.7* NQ <1.5 NQ 0.14* <0.1 40 m 6.7 12.2 6.4 NQ 3.7 4.1 0.32 <0.3 50 m 2.3 2.7 <1.0 NQ 0.48 0.6 0.32 <0.1 75 m 1.8 1.2 <1.0 NQ <0.1 <0.1 <0.1 <0.11 100 m 0.77 0.54 <1.0 NQ <0.5 0.13 <0.1 <0.1 150 m 0.27 0.36 <0.60 NQ <0.5 0.17 <0.1 <0.1 200 m 0.19 0.23 <0.70 NQ <0.5 <0.1 <0.1 <0.1 NQ Not quantified *Higher uncertainty due to ion suppression (recovery between 10-20%) of internal standard

116 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 2. Concentration (ng/g d.w) of PFCs in sediment samples from Langavatn near Flesland airport. Sample ID Dry weight 6:2 FTS PFBS PFHxS PFOS PFDS PFPeA PFHxA (%) Langavatn 1:1 8.2% 9.1 <0.11 2.2 70 0.89 0.91 1.3 Langavatn 1:2 8.5% 8.4 0.17 2.4 49 0.22 0.52 0.70 Langavatn 1:3 9.2% 9.8 <0.11 2.6 70 0.42 0.69 0.90 Langavatn 2:1 7.0% 12 <0.14 1.4 88 <0.21 0.69 0.76 Langavatn 2:2 8.1% 7.8 <0.14 1.3 85 0.38 <0.57 <0.6 Langavatn 2:3 12% 1.5 <0.10 0.80 35 <0.13 0.42 0.58 Langavatn 3:1 12% 4.1 <0.10 0.65 49 1.7 0.78 1.6 Langavatn 3:2 12% 7.3 <0.10 0.77** 57 1.3 0.57 0.70 Langavatn 3:3 13% 8.7 <0.10 0.70 53 0.40 0.37 0.45

Sample ID PFHpA PFOA PFNA PFDA PFUnDA PFDoDA PFTrDA PFTeDA Langavatn 1:1 0.32 <0.5 <2.5 NQ 11 0.50 3.2 <0.28 Langavatn 1:2 0.20 <0.5 <2.5 NQ 9.8 <0.5 0.81 <0.28 Langavatn 1:3 0.28 <0.5 <2.5 NQ 4.9 <0.5 0.77 <0.28 Langavatn 2:1 <0.1 <0.75 <3.2 NQ NQ NQ <0.1 <0.36 Langevann 2:2 <0.5 <0.75 <2.8 NQ 23 0.30 2.5 <0.31 Langevann 2:3 <0.5 <0.75 <1.9 NQ 0.56 <0.11 <0.1 <0.21 Langevann 3:1 0.15 <0.5 <1.8 NQ 21 0.76 7.9 <0.20 Langevann 3:2 0.17 <0.5 <1.9 NQ 26 0.96 11 <0.21 Langevann 3:3 0.17 <0.5 <1.8 NQ 25 0.18 0.90 <0.20 NQ Not quantified *Higher uncertainty due to ion suppression (recovery between 10-20%) of internal standard ** Difference between qualifier and quantitation ion >50%

117 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 3. Concentration (ng/g d.w) of polyfluorinated compounds in sediment samples taken near Res-Q. Sample ID Dry weight 6:2 FTS PFBS PFHxS PFOS PFDS PFPeA PFHxA (%) Res-Q-4:1 76% 27 <0.10 0.39 28 0.16 0.12 0.50 Res-Q-4:2 80% 6.5 <0.10 0.20 5.3 0.21 0.09 <0.50 Res-Q-5:1 81% 336* 0.56 6.7 369* 3.6* 2.2 16.7 Res-Q-5:2 75% 379* 0.54 5.9 493* 3.3* 2.7 19.0 Res-Q-5:3 72% 319* 0.64 5.5 385* 5.7* 2.2 16.5 Res-Q Bleivika 1 62% <0.14 <0.10 <0.10 <0.10 <0.10 <0.19 <0.10 Res-Q Bleivika 2 63% NQ <0.10 <0.10 NQ NQ <0.36 <0.10 Res-Q Bleivika 3 51% 0.20 <0.10 <0.80 <0.10 <0.10 <0.27 <0.10

Sample ID PFHpA PFOA PFNA PFDA PFUnDA PFDoDA PFTrDA PFTeDA Res-Q-4:1 <0.10 0.37 <0.47 NQ 2.0 0.16 1.0 <0.10 Res-Q-4:2 <0.10 0.21 <0.41 NQ 1.7 <0.10 0.83 <0.10 Res-Q-5:1 1.4 5.5 <1.4 NQ 1.4* <0.50 2.1* <0.10 Res-Q-5:2 1.2 5.7 <1.5 NQ 4.8 <0.50 4.9 <0.10 Res-Q-5:3 1.2 4.7 2.1 NQ 39* <0.50 4.4* <0.10 Res-Q Bleivika 1 <0.10 <0.50 <3.2 NQ <0.10 <0.10 <0.10 <0.10 Res-Q Bleivika 2 <0.10 <0.50 NQ NQ <0.10 <0.10 <0.10 <0.10 Res-Q Bleivika 3 <0.10 <0.50 <4.5 NQ <0.10 <0.10 <0.10 <0.10 NQ Not quantified *Higher uncertainty due to ion suppression (recovery between 10-20%) of internal standard

118 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 4. Concentration (ng/L) of PFCs in filtered water samples taken near Flesland airport and Res-Q. Sample ID 6:2 FTS PFBS PFHxS PFOS PFDS PFPeA PFHxA PFHpA Flesland airport 1 6693 74 471 2078 <0.41 631 554 205 Flesland airport 2 5254 68 413 1581 <0.39 552 439 177 Flesland airport 3 5110 148 319 1427 <0.38 498 385 172 Res-Q 1 NQ 35 48 NQ NQ 29 89 7.1 Res-Q 2 607* 11 33 131* <0.35* 22 64 11 Res-Q 3 7847* 28 44 181* <0.36* 33 81 7.7

Sample ID PFOA PFNA PFDA PFUnDA PFDoDA PFTrDA PFTeDA Flesland airport 1 191 33 4.2 17 <0.62 <0.60 <0.29 Flesland airport 2 143 22 3.4 9.1 <0.58 <0.27 <0.27 Flesland airport 3 130 30 3.7 6.8 <0.57 <0.27 <0.27 Res-Q 1 44 <12 <1.0 NQ NQ NQ NQ Res-Q 2 32 <12 <0.24 NQ NQ NQ NQ Res-Q 3 37 <12 <0.50 NQ NQ NQ NQ NQ Not quantified *Higher uncertainty due to ion suppression (recovery between 10-20%) of internal standard

119 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 5. Concentration (ng/g f.w.) of PFCs in biota samples taken near Flesland airport. Sample ID 6:2 FTS PFBS PFHxS PFOS PFDS PFPeA PFHxA PFHpA Sea mussel ( Mytilus edulis ) 1 nd <0.10 <0.05 <0.15 <0.05 <0.20 <0.30 <0.05 Sea mussel ( Mytilus edulis ) 2 nd <0.10 <0.05 <0.15 <0.05 <0.50 <1.0 <0.05 Sea mussel ( Mytilus edulis ) 3 nd <0.10 <0.05 <0.15 <0.05 <0.50 <1.0 <0.05 Crab ( Cancer pagurus ) 1 nd <0.20 0.46 1.0 <0.05 <0.30 <0.10 <0.05 Crab ( Cancer pagurus ) 2 nd <0.10 0.28 4.9* 0.56 <0.20 <0.29 <0.05 Crab ( Cancer pagurus ) 3 nd <1.3 0.12 2.3 0.17 <0.97 <0.16 <0.10 Crab ( Cancer pagurus ) 4 nd <0.10 <0.10 0.8 0.13 <0.20 <0.30 <0.05 Trout liver ( Salmo trutta ) 1 Detected <0.05 268 2103 65 <0.96 <0.15 <0.10 Trout liver ( Salmo trutta ) 2 Detected <0.05 132 2407 37 <0.20 0.24 <0.05 Trout liver ( Salmo trutta ) 3 Detected <0.30 118 2082 57 <0.51 <0.2 <0.07 Trout liver ( Salmo trutta ) 4 Detected <0.30 123 2532 49 <0.50 <0.15 <0.05

Distance from platform PFOA PFNA PFDA PFUnDA PFDoDA PFTrDA PFTeDA Sea mussel ( Mytilus edulis ) 1 <0.07 <0.10 <0.10 <0.10 <0.10 0.47 NQ Sea mussel ( Mytilus edulis ) 2 <0.08 <0.10 <0.10 <0.10 <0.10 0.38 NQ Sea mussel ( Mytilus edulis ) 3 <0.08 <0.10 <0.10 <0.10 <0.10 0.38 NQ Crab ( Cancer pagurus ) 1 0.82 0.12 <0.20 0.51 0.42 1.1 NQ Crab ( Cancer pagurus ) 2 <0.70 <0.30 <0.20 <1.0 <0.50 1.2* NQ Crab ( Cancer pagurus ) 3 <0.10 <0.23 <0.20 1.2 <0.50 2.0 NQ Crab ( Cancer pagurus ) 4 0.16 <0.20 <0.20 0.72 0.24 0.95 NQ Trout liver ( Salmo trutta ) 1 0.49 10** 14 140 14 83 NQ Trout liver ( Salmo trutta ) 2 <0.40 4.0 10* 81 5.6 39 NQ Trout liver ( Salmo trutta ) 3 <0.60 NQ 11 95 5.6 28 NQ Trout liver ( Salmo trutta ) 4 <0.20 NQ 11 89 4.7 24 NQ NQ Not quantified nd not detected *Higher uncertainty due to ion suppression (recovery between 10-20%) of internal standard ** Difference between qualifier and quantitation ion >50% Detected= compound detected but could not be quantified due to signal enhancement and lack of labeled standard

120 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 6. Concentration (ng/g f.w.) of PFCs in biota samples taken near Res-Q. Sample ID 6:2 FTS PFBS PFHxS PFOS PFDS PFPeA PFHxA PFHpA Sea mussel ( Mytilus edulis ) 1 nd <0.05 <0.05 <0.15 <0.10 <0.20 <0.20 <0.05 Sea mussel ( Mytilus edulis ) 2 nd <0.05 <0.05 <0.15 <0.10 <0.20 <0.20 <0.05 Sea mussel ( Mytilus edulis ) 3 nd <0.05 <0.05 <0.15 <0.10 <0.20 <0.20 <0.05 Crab ( Cancer pagurus ) 1 nd <0.20 <0.10 2.2 <0.05 <0.30 <0.10 <0.05 Crab ( Cancer pagurus ) 2 nd <0.20 <0.10 4.1 <0.05 <0.30 <0.10 <0.05 Crab ( Cancer pagurus ) 3 nd <0.20 <0.10 1.3 <0.05 <0.30 <0.10 <0.05 Crab ( Cancer pagurus ) 4 nd <0.59 <0.10 0.7 0.12 <0.99 <0.16 <0.10 Pollock liver ( pollacius pollachius ) 1 nd <0.20 <0.05 2.2** <0.05 <0.40 <0.30 <0.05 Saithe liver ( pollachius virens ) 2 nd <0.20 <0.05 2.9** <0.05 <0.40 <0.30 <0.05 Saithe and cod liver ( pollachius virens and gadus morhua ) 3 nd <0.10 <0.05 1.1** <0.05 <0.20 <0.31 <0.05

Distance from platform PFOA PFNA PFDA PFUnDA PFDoDA PFTrDA PFTeDA Sea mussel ( Mytilus edulis ) 1 <0.10 <0.10 <0.10 <0.05 <0.10 <0.10 NQ Sea mussel ( Mytilus edulis ) 2 <0.10 <0.10 <0.10 <0.20 <0.20 <0.20 NQ Sea mussel ( Mytilus edulis ) 3 <0.10 <0.10 <0.10 <0.05 <0.10 <0.20 NQ Crab ( Cancer pagurus ) 1 0.24 <0.20 <0.30 0.67 <0.30 <0.30 NQ Crab ( Cancer pagurus ) 2 <0.70 <0.24 <0.50 1.1** <0.20 0.49 NQ Crab ( Cancer pagurus ) 3 <0.40 <0.20 <0.10 <0.50 <0.30 <0.30 NQ Crab ( Cancer pagurus ) 4 <0.05 <0.24 <0.15 <0.60 <0.20 <0.50 NQ Pollock liver ( pollacius pollachius ) 1 <0.50 <0.20 <0.20 0.85 <0.20 <0.30 NQ Saithe liver ( pollachius virens ) 2 0.14 <0.70 0.31 1.3 <0.30 <0.30 NQ Saithe and cod liver ( pollachius virens and gadus morhua ) 3 <0.10 <0.20 0.16 <0.60 <0.20 0.28 NQ NQ Not quantified nd not detected *Higher uncertainty due to ion suppression (recovery between 10-20%) of internal standard ** Difference between qualifier and quantitation ion >50%

121 Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

10. Appendix II – PNEC values for BFRs

Diphenyl ether, pentabromo derivative: (PNECs taken from EU Risk Assessment Report 2001, available at: http://ecb.jrc.ec.europa.eu/esis) PNEC water : 0.53 µg/L PNEC sed : 1.55 mg/kg dw (=µg/g dw) PNEC soil : 0.38 mg/kg dw

Diphenyl ether, octabromo derivative: (PNECs taken from EU Risk Assessment Report 2003) Commercial mixture: PNEC water : ≥0.2 µg/L PNEC sed : ≥ 127 mg/kg dw (=µg/g dw) PNEC soil : ≥ 23.8 mg/kg dw

Hexabromodiphenyl ether: PNEC water : 0.53 µg/L PNEC sed : ≥ 7.0 mg/kg dw (=µg/g dw) PNEC soil : ≥ 1.3 mg/kg dw

DecaBDE: (PNECs taken from EU Risk Assessment Report 2002) PNEC water : ≥0.2 µg/L PNEC sed : ≥ 127 mg/kg dw (=µg/g dw) PNEC soil : ≥ 98 mg/kg dw

HBCD: (PNECs taken from EU Risk Assessment Report 2007) PNEC water : 0.31 µg/L PNEC STP : 0.15 mg/L PNEC sed : 0.86 mg/kg dw (=µg/g dw) PNEC soil : 5.9 mg/kg dw

TBBPA (PNECs taken from Aqua-team Report) PNEC water : 0.052 µg/L PNEC sed : 0.063 mg/kg (=µg/g) PNEC soil : 0.00094 mg/kg

TBBPA (PNECs taken from Swedish Chemical Inspectorate Report) PNEC water : 0.25 µg/L PNEC sed : 0.054 mg/kg ww (=µg/g dw)

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

11. Appendix III – Description of “new” BFRs included in the screening

The information below is a reproduction of the description given on these compounds in an earlier report (KLIF, 2009), with additional updates from the peer reviewed literature ( in italics ). For each chemical, a summary of current knowledge on the following five topics is presented:

● Chemical structure and physical data ● Toxicity data ● Bioaccumulation, degradation and fate ● Environmental levels (published data) ● Emissions and monitoring data from the Nordic countries

Tetrabromobisphenol A bis (2,3 – dibromopropyl-ether)

Br Br Br

BrCH 2 CH CH 2 O Me O CH 2

C Br Br Me Characteristics of the compound: CAS No.: 21850-44-2 CA Index name: , 1,1'-(1- methyl- ethylidene)b is(3,5- dibromo-4- (2,3- dibromopro poxy) -

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04) b) Data from SciFinder data base originating from experimentally determined properties c) Data from the WHO report: “Environmental health criteria 172. Tetrabromobisphenol A and derivatives” experimental data (WHO, 1995)

Toxicity: The acute LD 50 for mice was > 20 g/kg when given in feed and observed for 14 days. The acute dermal LD 50 for mice was > 20 g/kg when applied to closely clipped intact skin for 24 hours, and then observed for 14 days (WHO, 1995). Mice were administered levels of 200 or 2000 mg/kg per day in their diet for 90 days. At the end of the study, no deaths had occurred at either level. No abnormal symptoms were observed in the pathological examination (WHO, 1995). Low sub-chronic NOAEL = 200 mg/kg (Pakalin et al., 2007). Tetrabromo-bisphenol A bis(2,3-dibromopropyl ether) (TBBPA-DBPE) did not show any immunotoxic effect, in vitro, on the splenocytes of C57BL/6 mice (Pullen et al., 2003).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Bioaccumulation, degradation and fate: Biodegradation tests have shown a negative response, and accumulation in carp was judged to be very small showing that TBBPA-DBPE might not be readily biodegradable (WHO, 1995). Experimental studies on the hydrolysis of environmental contaminants showed TBBPA-DBPE to be susceptible to hydrolysis, at the same level as DDT with an experimental half-life of < 0.02 hours at 273K (methanol/DMF, 5/95 ratio) with sodium methoxid as a strong nucleophile. The elimination product, TBBPA bis(bromopropenyl ether), might be the more prevalent compound in sediments in a similar manner as DDE is for DDT (Rahm et al., 2005)

Environmental levels: TBBPA-DBPE at a concentration of 1.3 ng/g dust wt where identified in dust collected near an artificial stream and pond system in Berlin, Germany (Sawal et al., 2008).

Update: In the Pearl River Delta, China (one of China’s most urbanized and industrialized areas) substantial TBBPA-DBPE levels were measured in sediment (<1.5 – 2300 ng / g d.w), in air (241-1240 pg/m3), sewage sludge (238 – 8946 ng/g d.w) and in farmland soil (17 – 60 ng/g d.w) (Shi et al., 2009). They were also measured for in dust, birds (watercock) and various fish near an e-waste processing area in Southern China, and were not measured above quantification levels (circa 1.5 ng/g d.w). No other levels in the environment could be found, and no other findings were reported in a recent review on TBBPA derivatives in the environment (Covaci et al., 2009).

Emissions and monitoring data from the Nordic countries: No data available.

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Tetrabromobisphenol A dialyll - ether

Me Br Br C

Me H2C CH CH 2 O O CH 2 CH CH 2 Br Br

Characteristics of the compound: CAS No.: 25327-89-3 CA Index name: Benzene, 1,1'-(1- methylethylidene)bis[3,5- dibromo-4-(2-propen-1-yloxy)- Molecular Formula: C21 H20 Br4 O2 Molecular Weight (g/mol) 624.00 a Melting Point/range ( °C) 118-120 b, 115-120 cBoiling Point/range ( °C) 525.0±50.0 a Vapour Pressure (Pa (25 °C)) 1.83E-08 a Water Solubility (g/l (25 °C)) 1.60E-05 a a Partition Coefficient (log P ow ) 8.539±0.614 a Partition Coefficient (log K oc ) 6.02

a) Data from the SciFinder originating from calculated properties (ACD/labs Software V9.04) b) Data from the SciFinder data base originating from experimentally determined properties c) Data from the WHO report: “Environmental health criteria 172. Tetrabromobisphenol A and derivatives” experimental data (WHO, 1995)

Toxicity: TBBPA-diallylether showed no dermal or acute oral toxicity of rats using single doses of up to 5.0 g/kg or to 1 g/kg/day in the feed for 28 days. The acute inhalation LC 50 in rats was 13.4 mg/l. TBBPA-diallylether gavage exposure of pregnant rats from gestation days 6 through 15 caused no maternal toxicity and was not embryotoxic, fetotoxic, nor teratogenic (TOXNET, 2008). In a report the WHO concludes that, based on the available data, the acute oral and dermal toxicities of this compound are low (WHO, 1995).

Bioaccumulation, degradation and fate: Experimental studies on the hydrolysis of environmental contaminants showed TBBPA-diallyl ether not to be easily hydrolysed, with an experimental half-life of > 240 hours at 333 K (Methanol/DMF, 0.5/99.5 ratio) with sodium methoxid as a strong nucleophile. This suggests that TBBPA-bis(diallyl ether) might be resistant to environmental degradation (Rahm et al., 2005).

Environmental levels: No environmental levels of TBBPA-diallyl ether is published in the peer reviewed literature.

Emissions and monitoring data from the Nordic countries: No data available.

Update : No new information on this compound could be found in the peer-reviewed literature, in terms of environmental levels or toxicity, though additional issues related to

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) production and analytical determination have been presented (Covaci et al., 2009; Jonsson and Horsing, 2009).

Hexabromobenzene (HBB)

Br Br Br

Br Br Br

Characteristics of the compound: CAS No.: 87-82-1 CA Index name: Benzene, 1,2,3,4,5,6- hexabromo- Molecular Formula: C6 Br6 Molecular Weight (g/mol) 551.49 a Melting Point/range ( °C) 327 b, 326-327 c, 326 d Boiling Point/range ( °C) 417.5 ± 40 a Vapour Pressure (Pa (25 °C)) 1.14E-04a 3.17E-04 (liquid subcooled) d 7.5E-04 ( liquid c sub cooled )

a) Data from the SciFinder originating from calculated properties (ACD/labs Software V9.04) b) Data from the SciFinder data base originating from experimentally determined properties c) Experimental results from Tittlemier et al. (Tittlemier et al., 2002) d) Experimental results from Kawamoto and Kuramochi (2007)

Toxicity: Lowest toxic dose (TDLo) reported for HBB shown for rat for an intraperitoneal route of exposure was 150 mg/kg body wt giving biochemical effects on liver and porfyrin including bile pigments. For a continuous oral exposure to rat the TDLo were 3024 mg/kg/12weeks giving effects on liver, enzyme inhibition, induction, or change in blood or tissue levels and on ester-ases. Another study gave a TDLo for oral exposure on rat of 225 mg/kg/3days giving biochemical effects on liver, such as the activation of the hepatic microsomal mixed oxidases (RTECS, 2008). HBB was administered to mice as a single intraperitoneal dose of 20-90% of the approximate lethal dose or acute intoxication. No histopatological changes where observed, but HBB decreased the liver glutathione (GSH) levels, increased the gamma-glutamyltrasferase activity in serum and increased the malondialdehyde in liver (Szymanska et al., 1998). The lowest toxic dose reported in birds for an oral exposure on quail and chicken was 1.5 g/kg/15 days and 52.5 g/kg/12 weeks, respectively. Toxic effects on quail were on liver, metabolic effects, enzyme in-hibition, induction, or change in blood or tissue levels and porfyrin including bile pigments. Effects on chicken was effects on liver weight, weight loss or lack of weight gain, enzyme inhibition, induction, or change in blood or tissue levels and activation of hepatic mixed oxidase (RTECS, 2008).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Many studies on the effect of HBB on the heme synthesis in rat show HBB as a porfyrinogen (Mendoza et al., 1979;Smith and Francis, 1980;Koss et al., 1986;Szymanska and Piotrowski, 2000;Szymanska et al., 2002). Szym-anska and Piotrowski concluded that based on these and previous results HBB should be classified as a porfyrinogen (Szymanska and Piotrowski, 2000). No teratogenic effects were observed for HBB when orally administered to rats during day 5 to 15 of gestation at the maximum concentration of 200 mg/kg (Khera and Villeneuve, 1975).

Placental transfer of HBB has been observed in rats and HBB is accumulated primarily in the adipose tissue (Villeneuve and Khera, 1975). For prew-eaning rat pups feeded by HBB treated dams, transmission through the milk had effect on pup liver weight but showed no effect on weight on other organs (Mendoza et al. 1978).

Update : In a six days mortality test on the copepod Nitocra spinipes NOEC for HBB was determined to 33.4 mg/l (Breitholtz et al. 2008).

Bioaccumulation, degradation and fate: The bioconcentration and uptake rates, using perfused gills of rainbow trout showed that HBB had similar uptake rate constant as hexachlorobenzene while it did not accumulate in guppy (Sijm et al., 1993;Sijm et al., 1995). In another study the bioconcentration factor was determined for rainbow trout, results was a BCF of 1100 which is a high bioconcentration factor (Oliver and Niimi, 1985) and an absorbtion efficiency of 0.28-0.18 while the whole body halflife was >13-31 days (Niimi and Oliver, 1988). Another laboratory study showed that HBB did not bioconcentrate or accumulate either from water or food by juvenile atlantic salmon (Zitko and Hutzinger, 1976;Zitko, 1977). An assessment of HBB fate has been done using a multimedia mass balance model (Fugacity model level III) and experimentally determined physicochemical parameters. HBB is predicted to primarily distribute to soil (93%) and sediments (6.7%) and not to air and water (below 0.04%) (Kawamoto and Kuramochi, 2007). Tittlemier et al. (Tittlemier et al., 2002) also predicted HBB to be primarily distributed in soils (>98%) and sediments and the release into the environment would result in localized distributions.

Update : A study found that HBB is one of the residues found in polymeric brominated flame retardants that can be released at room temperature (Gouteux et al., 2008). The same study performed a Level III Fugacity model, and concluded that the high levels of HBB observed in the environment cannot be accounted for by release from polymers below, but are indicative that it is being used as an additive itself; however, this conclusion is only preliminary.

Studies to investigate the uptake behavior of HBB in earth worms show only very limited uptakes, with an uptake efficiency from 0.7 – 7.5 %, a biomagnification factor of only 0.7, and elimination rate constants between 0.04 and 0.09 day -1(Belfroid et al., 1995). Studies on zebrafish also indicated low levels of uptake, which were below detection limits, likely due to low uptake and metabolization (Nyholm et al., 2009) Metabolization of HBB in humans and rats was also reported (Yamaguchi et al., 1988a; Yamaguchi et al., 1988b). In a laboratory study on biodegradation kinetics of HBB in soil, the half-lives of HBB in aerobic and anaerobic soil were determined to 22 and 120 days, respectively (Nyholm et al. in press).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Environmental levels: HBB was analyzed in pooled herring gull egg samples from the Great Lakes of North America in 2004. Although present at much lower levels than the PBDEs (0.24 – 0.53 ng/g wet wt), HBB was generally the most abundant of the non-PBDE BFRs. The authors concluded that there are non-PBDE BFRs in the aquatic food web of the Great Lakes (Gauthier et al., 2007). HBB has been detected in air (Gauthier et al., 2007) and sediments (Watanabe et al., 1986).

Update : Measured total air levels (particle + gas) of HBB in Egbert, Ontario (a suburb of Toronto) were 0.02 – 0.09 pg/m 3 (Gouteux et al., 2008). An update on the study of BFRs in the Laurentian Great Lakes found HBB levels ranging in herring Gull eggs from 0.27 – 0.66 ng/g w.w (Gauthier et al., 2009). Human serum samples were collected in Tianjin, China, and about 20% of samples contained measureable levels of HBB, ranging from 0.11 – 1.5 ng/g lipid, with a median of 0.27 ng/g (Zhu et al., 2009). It is worth noting that levels of HBB in humans have been observed before, in the 1980s HBB was commonly used in Japan, and measureable levels of HBB and metabolites could be found in the adipose tissues of Japanese Yamaguchi et al., 1988b.

Emissions and monitoring data from the Nordic countries: In a screening for halogenated compounds in samples from an aluminum recycling plant, handling waste from electronics and electronics plastics and a car shredder, HBB was observed in all scrap samples (Sinkkonen et al., 2004). Further, HBB has been found in eggs and plasma of glaucous gulls in the Norwegian Arctic. Levels in egg yolk samples (0.4-2.6 ng/g wet wt) were comparable to those of the minor PBDEs (28, 116 and 155). Non-PBDE BFRs constituted only a small fraction of the total BFR content in egg yolk samples (Verreault et al., 2007).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Pentabromotoluene (PBT)

Br Br Me

Br Br Br

Characteristics of the compound: CAS No.: 87-83-2 CA Index name: Benzene, 1,2,3,4,5-pentabromo- 6-methyl- Molecular Formula: C7 H3 Br5 Molecular Weight (g/mol) 486.62 a Melting Point/range ( °C) 280-282 b, 288 c, 299 c Boiling Point/range ( °C) 394.4±37.0 a Vapour Pressure (Pa (25 °C)) 1.22E-03a Water Solubility (g/l (25 °C)) 7.80E-04 a a c Partition Coefficient (log P ow ) 5.872±0.615 , 5.43 a Partition Coefficient (log K oc ) 4.57

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04). b) Data from the SciFinder data base originating from experimentally determined properties. c) Data from a report by the Danish EPA (Simonsen et al., 2000).

Toxicity: Sprague-Dawley rats (15 rats/sex/dose level) were exposed to pentabromotoluene (PBT) in the diet 0.05 to 500.0 mg/kg diet (≈ 0.003-40 mg/kg body wt/day) for 91 days. No clinical signs of toxicity were observed, and growth rate and food consumption was not affected. PBT caused no dramatic changes in biochemistry, haematology and gross pathology. Mild dose-dependent histological changes were observed in the thyroid, liver, and kidney of rats fed PBT diets. The no observed adverse effect level (NOAEL) was 5.0 mg/kg diet ( ≈ 0.35 mg/kg body wt/day) (Chu et al., 1987). PBT was administered to rat and the TDLo was 4200 mg/kg/28days and 13.65 g/kg/91days with effects on liver and kidney/ureter/bladder and endocrine effects such as changes in tyroid weight and effect on haematology (normocytic anemia) (RTECS, 2008). The LC 50 for fish was > 5 mg/l (48 hours) (Simonsen et al., 2000). No adverse foetal effects were observed when doses up to 600 mg/kg body wt were given orally to rats during organogenesis (Simonsen et al., 2000).

Bioaccumulation, degradation and fate: Studies of dietary absorption eff-iciency in rainbow trout and BCF showed that PBT had an absorption eff-iciency of 0.18-0.28 and whole-body half-life of 13-23 days and a BCF of 270 (Oliver and Niimi, 1985;Niimi and Oliver, 1988). The bioconcentration factor in fish was determined to 4.5-39 (Simonsen et al., 2000). PBT was found to not be readily biodegradable (7% of BOD, 4weeks, 100 mg/l substance, 30 mg/l sludge) (Simonsen et al., 2000)

Environmental levels: Pentabromo-toluene was analyzed in egg pools of herring gull from the Great Lakes of North America in 2004. Results showed much lower levels than for PBDEs (0.004 – 0.02 ng/g wet wt) and were generally the lowest of the non-PBDE BFRs.

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

The authors concluded that the results suggests that there are non-PBDE BFRs in the aquatic food web of the Great Lakes (Gauthier et al., 2007). In a screening of sediment samples from the Elbe river and its tributaries for new contaminants, pentabromotoluene showed a concentration range of <1-25 ng/g dry wt (Schwarzbauer et al., 2001).

Update : It should be added here that another way for PBT to enter the environment is that it, along with TBP, is an environmental transformation product of TBBPA (Arbeli et al., 2006). It is likely also a degradation product of many other BFRs, such as DBDPE (in which just the methyl-methyl bond would have to be cleaved).

Emissions and monitoring data in the Nordic countries: Sewage sludge samples from Swedish waste-water treatments plants contained a few brominated toluenes such as penta- and two isomers of tetra-bromotoluene (Mattsson et al., 1975). PBT have been found in eggs and plasma from glaucous gulls in the Norwegian arctic. Levels in plasma was in the range of

Pentabromoethylbenzene (PBEB)

Br Br Et

Br Br Br

Characteristics of the compound: CAS No.: 85-22-3 CA Index name: Benzene, 1,2,3,4,5- pentabromo-6-ethyl- Molecular Formula: C8 H5 Br5 Molecular Weight (g/mol) 500.65 a Melting Point/range ( °C) 138 b Boiling Point/range ( °C) 413.3±40.0 a Vapour Pressure (Pa (25 °C)) 3.2E-04 a Water Solubility (g/l (25 °C)) 3.50E-04 a a Partition Coefficient (log P ow ) 6.403±0.615 a Partition Coefficient (log K oc ) 4.86

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04). b) Data from the SciFinder data base originating from experimentally determined properties.

Toxicity: Administration of penta-bromoethylbenzene (PBEB) onto the skin of rabbit’s gave an LD 50 > 8g/kg, no details was reported on effects (RTECS, 2008).

Bioaccumulation, degradation and fate: Studies of dietary absorption efficiency in rainbow trout and BCF showed that PBEB had an absorption efficiency of 0.26 and whole-

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) body half-life of 38 days while the bioconcentration study resulted in a moderate BCF of 330 (Oliver and Niimi, 1985;Niimi and Oliver, 1988).

Environmental levels: PBEB was analyzed in egg pools of herring gull from the Great Lakes of North America in 2004. PBEB levels were found in the range of 0.03 – 1.4 ng/g wet wt, which is 0.7 % compared to sum PBDEs (47, 99, and 100) (Gauthier et al., 2007). A study by Hoh et al. reported a relatively high abundance of PBEB in the atmosphere of Chicago (summer of 2003). PBEB was detected in both gas and particle phases (520 pg/m 3 gas phase and 29 pg/m 3 in particle phase), with peak intensities 100 times higher than for the PBDEs (sum PBDEs tri-hexa of 47 pg/m 3). Other compounds such as tetrabromoethylbenzenes, which the authors believe to be byproducts of PBEB, were detected but not quantified (Hoh et al., 2005). Screening of air samples in three locations in UK and Ireland (reference site) reported a mean concentration of PBEB of 30 pg/m 3 in the southwest Oxford (Lee et al., 2002).

Update : PBEB was analyzed in a time series of lake trout in Lake Ontario, Canada. Concentrations ranged from 17 ± 3 to 320 ± 156 ng/g lipid, and showed no relationship with time between the years 1979 and 2004 (Ismail et al., 2009). It is worth noting that PBEB was known to be produced in the 1970’s and 1980’s, so its constant concentration in this sample might indicate that production levels or usage levels did not change substantially over this time frame. In a new investigation where “new” brominated flame retardants were analysed in higher animals (sea birds, capelin, ice bear, ring seal and and polar fox) in Arctic PBEB was not detected in any samples (KLIF unpublished data).

Emissions and monitoring data in the Nordic countries: Filter dust, cyclone dust and light fluff samples of an aluminium recycling plant in Finland, handling waste from electronics and electronics plastics and a car shredder was screened for halogenated compounds. PBEB was observed in all scrap samples with quite high concentrations (no concentrations was assigned), and was among the most abundant alkylbromobenzenes (Sinkkonen et al., 2004). PBEB have been found in eggs and plasma from glaucous gulls in the Norwegian arctic. PBEB was below the method limit of quantification (MLOQ) values in plasma and was only detected in the range (0.03-0.23 ng/g wet wt) in egg yolk samples. Non-PBDE BFRs constitute only a small fraction of the total BFR content in egg yolk samples (Verreault et al., 2007).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

2,4,6 – tribromophenyl alyll ether (ATE)

Br

O CH 2 CH CH 2

Br Br

Characteristics of the compound: CAS No.: 3278-89-5 CA Index name: Benzene, 1,3,5-tribromo-2- (2-propen-1-yloxy)- Molecular Formula: C9 H7 Br3 O Molecular Weight (g/mol) 370.86 a Melting Point/range ( °C) Not available Boiling Point/range ( °C) 339.5±37.0 a Vapour Pressure (Pa (25 °C)) 4.9E-02 a Water Solubility (g/l (25 °C)) 2.0E-02 a a Partition Coefficient (log P ow ) 4.974±0.564 a Partition Coefficient (log K oc ) 4.08

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04).

Toxicity: No data available .

Bioaccumulation, degradation and fate: 2,4,6-Tribromophenyl allyl ether (ATE) was proposed to be one of 120 high production chemicals which are structurally similar to known arctic contaminants and/or have partitioning properties that suggests they are potential arctic contaminants (Brown and Wania, 2008).

Environmental levels: ATE was found in blubber and brain of hooded and harp seal from the Barents sea at concentrations of 5.4 – 9.1 and 3.1 – 10 ng/g wet wt, respectively (Vetter, 2001;Von Recke and Vetter, 2007). The authors showed that an experimental aerobic degradation using corrinoids reduced 2,3-dibromopropyl-2,4,6-tribromophenyl ether (DPTE), which is the main component of Bromkal 73-5 PE, to ATE and 2-bromoallyl-2,4,6- tribromophenyl ether. Comparing the ratio in blubber and brain in harp seal showed that DPTE is the more pro-minent BFR in these samples (ratio ATE/DPTE of 0.018 in blubber and 0.030 in brain) and the authors conclude that the presence of ATE is probably mainly due to the transformation of DPTE (Von Recke and Vetter, 2007). ATE have also been detected in 15 of 18 municipal sewage sludge samples in Germany from 10 different sewage treatment plants at a range of < 0.005-0.091 mg/kg dry wt. Also DPTE was found in 12 out of 18 samples at a range of < 0.025-0.596 mg/kg dry wt (ratio of the mean ATE/DPTE of 0.17) proving that ATE and DPTE is not degraded in these sludge treatment processes (Weisser, 1992). These types of compounds seems also to accumulate to a higher degree than PBDEs in brain tissues of harp seal and seems to pass the blood-brain barrier (Von Recke and Vetter, 2007).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Update: No new reports of this compound in the environment could be found. It was analysed for in a study in Great Lake Herring Gulls but no concentrations were reported (Gauthier et al., 2009).

Emissions and monitoring data in the Nordic countries: No data available.

2,3 – dibromopropyl – 2,4,6 – tribromophenyl ether (DPTE)

Characteristics of the compound: CAS No.: 35109-60-5 CA Index name: Benzene, 1,3,5-tribromo-2- (2,3-dibromopropoxy)- Molecular Formula: C9 H6 Br6 O Molecular Weight (g/mol) 530.68 Melting Point/range ( °C) Boiling Point/range ( °C) Vapour Pressure (Pa (25 °C)) 1.0E -05 a (subcooled) a a) SPARC online calculator (accesses February, 2010 at http://sparc.chem.uga.edu/sparc/search/searchcas.cfm)

Toxicity : No data available .

Bioaccumulation, degradation and fate: DPTE was selected as it is the main component of bromkal 73-5PE, and apropable reductive precursor of the priority compound; 2,4,6- Tribromophenyl allylether (ATE, CAS 3278-89-5) (KLIF, 2009; von der Recke and Vetter, 2007).

Environmental levels: Quoting from Vetter et al. (Vetter et al., 2010): “ Concentrations of DPTE in the North Pacific were from 0.3 to 5.6 lg / kg lipid weight In sewer slime from German urban residential zones, the amount of DPTE in sewage sludge was found to be up to 1.9 mg / kg dry weight (Sauer et al., vom Wasser, 1997). In addition, DPTE caused the highest peak in a GC/ECD chromatogram of snoek fillet (Thyrsites atun) from the South Atlantic, but its concentration was not calculated (Hackenberg et al., ES&T 2003). Blubber and brain samples of hooded seals (Cystophora cristata) and harp seals (Phoca groenlandica) from the Barents and Greenland Seas contained up to 470 g / kg wet weight DPTE (von der

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Recke and Vetter, 2007). Moreover, the DPTE brain/blubber ratio in harp and hooded seals was 5–30-fold higher compared to PCBs and PBDEs (von der Recke and Vetter, 2007).

Emissions and monitoring data in the Nordic countries: Levels in blubber and brain from harp seals in the Barents Sea were 322 – 470 and 130 – 340 µg / kg wet weight, respectively.

2,4,6 – tribromophenol (TBP)

Br

OH

Br Br

Characteristics of the compound: CAS No.: 118-79-6 CA Index name: Phenol, 2,4,6-tribromo- Molecular Formula: C6 H3 Br3 O Molecular Weight (g/mol) 330.80 a Melting Point/range ( °C) 94-95b, 95-96 b, 95.5 c, 87-89 c Boiling Point/range ( °C) 286.8 b, 244 c, 282-290 c 286 cVapour Pressure (Pa (25 °C)) 0.41 a Water Solubility (g/l) 1.30 b, 7.1E-2 (at 15°C) c 0.995 (at 15°C) c 0.968 (at 25°C) c 0.883 (at 35°C) c a c c Partition Coefficient (log P ow ) 4.326±0.486 , 4.02 , 3.3 a Partition Coefficient (log K oc ) 2.98 (pH 7)

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04) b) Data from SciFinder data base originating from experimentally determined properties c) Data from a report by the Danish EPA (Simonsen et al., 2000)

Toxicity: Oral LD 50 values during an administration of TBP to male and female rats was 1,995 and 1,819 mg/kg body wt, respectively (Simonsen et al., 2000). Another study on male and female rats resulted in oral LD 50 values of 5,012 and 5,012 mg/kg body wt, respectively. Signs of toxicity included decreased motor activity, nasal discharge, lacrimation, tremors, prostration, clonic convulsions and death (Simonsen et al., 2000). 3 Inhalation studies on rat showed an LC 50 >1630 mg/m /4 hours, with effects on sense organs (ptosis on eye) (RTECS, 2008). An inhalation study of dust with TBP gave an LC 50 of >1.63 mg/l/4 hours (65% of the particles were less than 6 microns) and another study with an LC 50 of >200 mg/l/1 hour tested at two concentration levels 2 or 200 mg/l. Signs at both concentrations included nasal discharge, eye squint, increased followed by decreased respiratory rates, prost-ration, salivation, lacrimation, erythema, increased followed by decreased motor activity, and ocular and nasal porphyrin discharge. No details were available

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) about e.g. particle size or distribution. All rats appeared normal from day 7 post exposure, except on day 10 of the 14 day observation period when one rat at low exposure level exhibited nasal porphyrin discharge (Simonsen et al., 2000).

Dermal administration of TBP on rabbit gave an LD 50 of >2000 mg/kg body wt and >8,000 mg/kg body wt (Simonsen et al., 2000). Three groups of rats each consisting of 5 males and 5 females in a subchronic toxicity study, were exposed (whole-body) to atmospheric dust concentrations (analytical) of 0, 0.10 and 0.92 mg/l, respectively, for 6 hours/day, 5days/week, for 3 weeks. The NOAEL in this study appears to be <0.10 mg/l for females and 0.10 mg/l for males. No dermal toxicity to albino rabbits during a 28-day sub acute dermal toxicity study was observed (Simonsen et al., 2000).

TBP inhalation of 100 µg/m 3/24 hours for female rats during 1-21 days after conception gave fetotoxicity (except death, eg. stunted fetus), developmental abnormalities and behavioral changes of the newborn pup. TBP inhalation study on female mouse of 0.15mg/m 3, 122 days pre-mating, showed post implant-ation mortality and fetotoxicity (RTECS, 2008).

In a pilot study, mated Charles River CD female rats were dosed with TBP by gavage at 10 to 3,000 mg/kg/day from gestation day 6 through day 15. All animals died at the highest dose group after one day of treatment. There were slight decreases in body weight gains between days 6 and 12, an increase in post implantation losses, and a slight decrease in the number of viable foetuses at the 1,000 mg/kg/day dose group. The NOAEL appears to have been 300 mg/kg/day for both dams and foetuses (embryotoxicity). In order to investigate the developmental neurotoxicity and immunotoxicity, pregnant Wistar rats were exposed to TBP by inhalation 0.03 - 1.0 mg/m 3, from day 1 to 21 of ges-tation. The results suggested that TBP during this exposure regime may be a developmental neurotoxicant, embryo-toxicant and foetotoxicant but not immunotoxicant. The NOAEL for developmental neurotoxicity could not be established (<0.03 mg/m 3), and the NOAEL for maternal neurotoxicity was 0.3 mg/m 3 (Simonsen et al., 2000).

Pregnant wistar rats were orally administered Aroclor 1254, hydroxylated PCBs, BDE-47 and TBP (25mg/kg/day) during gestational day 10-16. They monitored endocrine effects, developmental landmarks, sexual and neurobehavioural development and transplacental transfer. Results indicated that the hydroxylated PCB metabolites and BFRs are capable of placental transfer while no effects was observed (at these concentrations) on the developmental landmarks (Buitenhuis et al., 2004).

For tribromophenol the LC 50 (in fish) was 6.5-6.8 mg/l (96 hours, fathead minnow) and 1.1 mg/l (96 hours, fathead minnow, flow through bioassay) (Simonsen et al., 2000).

Update : The binding of TBP to Hemoglobin Dehaloperoxidase from Amphitrite ornate has been characterized (Davis et al., 2009). In a six days toxicity test on larvae development of the copepode Nitocra spinipes the NOEC for TBP was determined to 300 µg/l (Breitholtz et al. 2008).

Bioaccumulation, degradation and fate: Aerobic biodegradation was tested using TBP at 100 mg/l in an activated sludge inoculum in the Japanese MITI test. It reached 49% of its theoretical BOD in 28 days. TBP was not degraded over 14 days in a marine sediment slurry.

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

No ring degradation was reported for TBP at 100 mg/l, over a 5-day period using a soil inoculum. Water samples collected from two treatment ponds were unable to degrade TBP over 32 days (TOXNET, 2008). An anaerobic degradation test showed that TBP was rapidly dehalogenated with >90% degradation in 2 days in a marine sediment slurry. A first-order rate constant of 0.19 day -1 was reported for TBP in anoxic sediment from Loosdrechtse Plassen (TOXNET, 2008). BCF values of 513 and 83 were measured in zebrafish and fathead minnow, respectively, for TBP. These BCF values suggest that the potential for bioconcentration in aquatic organisms is moderate to high (Simonsen et al., 2000). Flodin et al. studied the biosynthesis pathway of bromophenols from the green marine algae Ulva lactuca showing that certain precursors are converted to bromophenols by the bromoperoxidase (Flodin and Whitfield, 1999). As the ecological function of bromophenols are not yet clear, researchers suggest that they may play a role as a natural chemical defense and deterrence (Hassenkloever et al., 2006).

TBP was proposed to be one of 120 high production chemicals which are structurally similar to known arctic contaminants and/or have partitioning properties that suggests they are potential arctic contaminants (Brown and Wania, 2008).

Update : In a biodegradation study in soil relatively fast degradation of TBP was observed. The half-life of TBP in aerobic and anaerobic soil was determined to 10 and 7 days, respectively (Nyholm et al. in press). In an uptake study of TBP in zebrafish, relatively low levels were measured in zebrafish as compared to the feed suggesting low biomagnification potential. The uptake efficiencies of TBP were estimated to 45 %, and the half life of TBP in zebrafish to 1.3 days (Nyholm et al., 2009).

Environmental levels: Endeavour prawns from Exmouth Gulf, Shark Bay, and Groote Elylandt, Australia, contained TBP at concentrations of 41 to 97, 7.8, and 8.5 ug/kg, respectively (Whitfield et al., 1992). Ten different species of fish, collected in August 1992 from the eastern coast of Australia, contained TBP at concentrations of <0.05 to 3.4 ng/g for the carcass and <0.05 to 170 ng/g for the whole gut (analysis of single fish from each species) (Whitfield et al., 1995). Ocean fish were separated by species into pelagic carnivores, benthic carnivores, diverse omnivores and restricted omnivores; concentrations in the flesh ranged from <0.01 to 0.9 ng/g, <0.01 to 12 ng/g, <0.01 to 4.3 ng/g, and 0.1 to 1.4 ng/g, respectively, while concentrations in the gut ranged from <0.01 to 11 ng/g, <0.01 to 230 ng/g, 0.04 to 55 ng/g, and 7 to 45 ng/g, respectively (Whitfield et al., 1998). Marine fish (salmon), crustaceans and mollusks from the Pacific Ocean contained bromophenols while only low levels was found in freshwater salmon from the Great Lakes (Boyle et al., 1992). Thirty samples of 9 species of prawns, collected from the eastern coast of Australia from 1993 to 1996, contained TBP at concentrations of <0.01 to 170 ng/g while TBP concentrations in cultivated prawns ranged from <0.01 to 0.53 ng/g (Whitfield et al., 1997). Concentrations of TBP were measured in brown algae (14 to 38 ug/kg wet weight), red algae (4.5 to 68 ug/kg), bryozoa (24 and 27 ug/kg), a hydroid (29 ug/kg), and sponges (0.22 to 240 ug/kg) collected from Exmouth Gulf, Australia, in October 1990 (Whitfield et al., 1992). An analysis of selected Hong Kong seafood (eg. rabbitfish, clam and shrimp) as well as brown algae found concentrations of the total bromophenol content to be 40-7000 ng/g dry wt, which varied with the season with crab having the highest seasonal concentration of mono- di- and tribromophenols (Chung et al., 2003a;Chung et al., 2003b). Ahn et al. found sponges to contain brominated organic compounds up to 12% of the dry wt of eg. bromoindoles,

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) bromophenols (mono-, di- and tribromophenol), and bromopyrroles and showed that these aerophobic sponges harbor bacteria that are capable of an reductive dehalogen-ation processes (Ahn et al., 2003). Upper river and marine sediment layers in Osaka Prefecture, Japan, collected in 1981 through 1983 at 12 different locations, contained TBP at concentrations ranging from <0.2 ppb to 35 ug/kg dry wt (Watanabe et al., 1985). Surficial sediments from 5 sampling sites in the Rhone estuary, collected in 1987/1988, contained TBP at concentrations of 26 to 3690 ng/g dry wt (Tolosa et al., 1991). TBP have been found in wine as contaminants from old winerys originating from structural elements of the winery or the wooden containers (Chatonnet et al., 2004). TBP might be found as a product in the combustion of tribromoaniline and SB 2O3 flame retarded materials such as textiles and plastics (Bindra and Narang, 1995). Indoor dust in Japan was studied for potential thyroid disrupting compounds suggesting that TBP and pentachlorophenol are potential thyroid disrupting compounds in homes and work environments of Japan and other counties while indoor dust also is an important exposure route to children (Suzuki et al., 2008).

Update : TBP was taken up by SPMD samplers in the Great Barier Reef, Australia, but due to lack of calibration these could not be converted into aqueous concentrations (Vetter et al., 2009). They were also found in isolated Antarctic sponges(Vetter and Janussen, 2005). TBP was analysed in indoor air and outdoor air in two family homes in Hokkaido, Japan (Takigami et al., 2009), where in the vapor phase TBP in outdoor air was 49 and 73 pg/m 3, and indoor concentrations were elevated by an order of magnitude from 220 – 690 pg/m 3 (much smaller amounts were reported sorbed to dust, at 15 and 30 pg/m 3). This indicates TBP is being generated indoor. In muscles and fish from Salvador, Brazil, levels ranged from 6 – 171 ng / g in muscle and <3 – 104 ng/g in fish stomach.

It should be additinally noted here that another way for TBP to enter the environment is that it, along with PBT and other compounds, is an environmental transformation product of TBBPA (Arbeli et al., 2006).

Emissions and monitoring data in the Nordic countries: The raw flue gas from a Swedish hazardous waste incinerator, located at Norrtorp, and fed chlorinated (mainly solvents) and brominated waste (tetrabutylammonium bromide) contained TBP at <14, 380, and 260 ng/m 3 over three tests; bromides were present initially at 32, 110, and 530 mg/m 3. Flue gas from this incinerator, fed municipal waste, contained TBP at 4-5 ng/m 3. Peat combustion released TBP at concentrations of <5 to 60 ng/m 3 (Oeberg et al., 1987). TBP was analysed among other FRs (TBBP-A and PBDEs) in 22 municipal waste water treatment plants. TBP was in the range n.d. to 0.9 ng/g wet wt (Oberg et al., 2002). During an investigation of sediments and water in the North and Baltic sea, TBP was found in water samples from the the German Bight in the range of n.d. to 6 ng/l but was not found in any of the investigated sediment samples (Reineke et al., 2006). A study on the occupational exposure to BFRs in Norwegian workers at an electronics dismantling plant, plasma samples was collected and analysed for PBDEs, halogenated phenols (including TBP) and tetrachloro- and bromobisphenol A. TBP was generally the most abundant BFR at 0.17 to 81 ng/g lipid wt (Thomsen et al., 2001). In a study on temporal trends (1977-2003) and the role of age, pooled serum samples from the Norwegian population was analysed for BFRs. TBP was found not to follow a similar trend as PBDEs. PBDEs increased from 0.5 ng/g lipid wt in 1977 to 4.8 ng/g lipid wt in 1988 and the levels stabilized from 1989 to 2003 while TBP showed no relations to trends or age

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) which might be due to short half-lives of TBP in humans (Thomsen et al., 2002;Thomsen et al., 2007)

Update : Vapour concentrations of TBP in Lista, Norway were monitored weakly for one year (Melcher et al., 2008), concentrations appeared seasonal, highest between August and November, ranging from below detection limits to 5.9 pg/m3.

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Decabromodiphenylethane (DBDPE)

Br Br Br Br CH 2 CH 2

Br Br Br Br Br Br

Characteristics of the compound: CAS No.: 84852-53-9 CA Index name: Benzene, 1,1'-(1,2- ethanediyl)bis[2,3,4,5,6- pentabromo- Molecular Formula: C14 H4 Br10 Molecular Weight (g/mol) 971.22 a Melting Point/range ( °C) 334-337 b, 344-349 c 348-353 d, 351-355 d Boiling Point/range ( °C) 676+/-50 a Vapour Pressure (Pa (25 °C)) 6.0E-15 a, <1E-04 (20 °C) d Water Solubility (g/l (25 °C)) 2.10E-07 a, 7.2E-04 d a Partition Coefficient (log P ow ) 11.1 Partition Coefficient (log K ) 7.0 a oc a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04) b) Data from the SciFinder data base originating from experimentally determined properties c) Experimental data from Li et al. (Li et al., 2004) d) Experimental data from the UK EPA (Dungey and Akintoye, 2007)

Toxicity: Single dose and long term (90 days) oral administration of DBDPE in rats resulted in high LD 50 values (> 5000 mg/kg body wt) and LDLo (90 g/kg/90 days) where the highest doses gave changes in liver weight and slight histomorphological effects. The dermal acute toxicity (LD 50 ) in rabbits was > 2000 mg/kg body wt, the NOAEL in rat was estimated to ≥1000 mg/kg/day (Hardy et al., 2002;RTECS, 2008) while the inhalation route is not suspected by the Canadian centre for occupation health and safety to be acutely toxic (RTECS, 2008). Authors conclude that the lack of toxicity for DBDPE is likely due to poor bioavailablity due to its high molecular weight and poor water solubility (Hardy et al., 2002). Acute toxicity data on ingestion, inhalation, dermal contact in rat at a maximum concentration of oral - 2000mg/kg body wt (single dose), dermal – 2000 mg/kg body wt (24 hours) and inhalation – 50 mg/l (1 hour). Results proved that DBDPE exhibit a low acute oral, dermal and inhalation toxicity (Li et al., 2004).

Update: A review of the animal and human toxicity of DBDPE (Hardy et al., 2009), and concluded that this was relatively low in toxicity (compared to other BFRs), with a reference dose of 4 mg/kg body weight/day being safe of sensitive subpopulations. Regarding environmental benchmarks, one study show that levels of 19 µg/l were acutely toxic to water fleas, with a 48 hour EC50 response (Nakari and Huhtala, 2009).

Bioaccumulation, degradation and fate: DBDPE was considered to be not readily biodegradable in an activated sludge inocculum, tested in compliance with the OECD test

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) guidelines 301C (modified MITI test). 100 mg/l of DBDPE with a purity of 96.6 % was incubated with 30 mg/l activated sludge at 25 °C over 28 days. No degradation was observed as measured by the BOD. The risk evaluation report by UK EPA suggested that DBDPE is unlikely to be rapidly degraded, based on biodegradation results above, calculations on the atmospheric (OH radical) and aquatic (hydrolysis) degradation rates. They stated that other degradation mechanisms cannot be excluded, such as anaerobic degradation as may be found in WWTP (Dungey and Akintoye, 2007). DBDPE did not bioconcentrate (BCF of <2.5-<25) in Japanese carp during an 8 week exposure while compounds with a molecular weight below 700 Da did bioconcentrate (Hardy, 2004). Biomagnification was observed in between the trophic levels of Lake Winnipeg (Canada) food web resulting in an BMF of 0.2-9.2 for DBDPE, and was highest between the top predator walleye and bottom-feeding white suckers with an BMF of 9.2 (Law et al., 2006;Law et al., 2007).

Update : An exposure study of children to DBDPE in childrens toys estimated that, within the sample study, expsosures were 1323 to 15 085 pg/kg bw-day (Chen et al., 2009).

Environmental levels: A food web study of Lake Winnipeg (Canada) observed this compound for the first time in fish (walleye) with the highest mean concentration of DBDPE of 1.0 ± 0.5 ng/g lipid wt while it was not detected in zooplankton, mussels, and whitefish. The concentration in sediments was below the method detection limit (MDL) and the concentration in lake water was difficult to measure due to the high lipophilicity of DBDPE (Law et al., 2006;Law et al., 2007). Several non-PBDE BFRs was detected in a study of egg pools of herring gulls (Larus argentatus ) from seven colonies in the Great Lakes (collected in 1982 to 2006). The concentrations of DBDPE in eggs from 2005 and 2006 of three of the seven colonies were 1.3 to 288 ng/g wet wt and surpassed decaBDE. The authors concluded that there is an indication that there have been a continual exposure and bioaccumulation of several BFRs in the Great Lakes (Gauthier et al., 2008). DBDPE was recently detected in two species of captive panda in China in 87 and 71 % of the giant and red panda samples, at concentrations up to 863 ng/g lipid wt, respectively. DBDPE and decaBDE dominated the samples and the authors suggested that these levels might relate to significant production, use or disposal of BFRs in China (Hu et al., 2008). DBDPE have been found in house dust in the U.S. ranging from <10 to 11070 ng/g dust wt with a median value of 201 ng/g dust wt (Stapleton et al., 2008), which is ten times higher than the levels found in Sweden (Karlsson et al., 2007). DBDPE have also been found primarily in the particulate phase in air near the Great Lakes (U.S.) at median concentrations from 1 to 22 pg/m 3(Venier and Hites, 2008) and have been found in tree bark from the Northeastern U.S. ranging from ND to 0.73 ng/g bark wt (Qiu and Hites, 2008).

Update: In the Pearl River Delta, China (a heavily industrialized area), concentrations from 39 – 364 ng/g d.w were measured in sediment, 402 – 3578 pg/m3 in air (vapor and particle), 266 – 1464 ng/g dw in sewage sludge, and 18 – 36 ng/g d.w in farmland soil (Shi et al., 2009). In an e-waste processing area in Southern China, DBDPE levels ranged from <2.5 – 96 ng/g d.w in dust, 10 – 50 ng/g w.w in diverse organs from watercock (bird) samples, and were < 3.8 ng/g lipid in diverse fish samples. In another study of sediments from this region, DBDPE concentrations ranged from 19 – 430 ng/g d.w. Another study on waterbirds from e- waste region in the Pearl River Delta reported median DBDPE concentrations of 10 – 176 ng/g in lipids, with the total range from n.d. – 900 ng/g (Luo et al., 2009). In Northern China, in the Yellow River Delta (an important industrial development area, as well as an important

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) breeding ground for birds), median concentrations in various bird egg samples range from n.d. – 1.7 ng/g lipid, and totally from n.d. – 2.2 ng/g overall (Gao et al., 2009). In herring gull eggs from the Laurentian Great Lakes, concentrations of DBDPE ranged from n.d. – 44 ng/g w.w (Gauthier et al., 2009). Elsewhere in China, levels of DBDPE were found in various organs of Giant and Red Panda from n.d. to 41 ng/g lipid (Hardy and Ranken, 2008; Hu et al., 2008a; Hu et al., 2008b). Levels were also spotted in occasional tree bark from Arizona, thought to be due to nearby production sources (Zhu and Hites, 2006). A study in different indoor environments in Birmingham, UK (cars, offices, and houses) found dust concentrations of < dl to 3400 ng/g (with average in cars, offices and houses being 400, 170 and 270 ng/g, respectively) (Harrad et al., 2008). A survey of worldwide sludge from waste water treatment plants found levels ranging from < dl to 160 ng/g d.w (Ricklund et al., 2008).

Emissions and monitoring data in the Nordic countries: BFRs were determined in air, sedimentary dust and plasma from five households in Sweden. DBDPE was not detected in plasma but in one of the five air samples at a concentration of 0.013 ng/m 3, while in sedimentary dust the DBDPE was among the most abundant BFRs with an average concentration of 47ng/g dust wt, 1/10 of the concentration of BDE-209. Due to the limited data no firm conclusions could be drawn on the relationship of the plasma and air or dust levels while sumBDE concentrations correlated between plasma and dust levels. The author suggests that as BTBPE and DBDPE were found in household dust at similar concentrations as many of the PBDEs, show that humans and especially toddlers are exposed to these compounds in their homes via the dust (Karlsson et al., 2007). DBDPE was monitored in wastewater, sludge, sediment and indoor air in Sweden. DBDPE was observed in 25 of 50 Swedish waste water treatment facilities with an estimated concentration of 100 ng/g dry wt, an air sample from an electronic dismantling plant showed a concentration of 0.6 ng/m 3 and DBDPE was also found in water piping insulation (Kierkegaard et al., 2004). To evaluate exposure to BFR in an electronic recycling facility, personal air monitoring was done for 2 years. A total of 22 polybrominated di-Ph ethers (PBDE) and 2 other Br- containing organic compounds were analyzed and evaluated in 17 personal air samples (Pettersson-Julander et al., 2004). One of the compounds was identified as as 1,2-bis(2,4,6- tribromophenoxy)ethane (BTBPE) based on full scan spectra and previous identifications (Sjodin et al., 2001), the other compound was tentatively identified as DBDPE based on fullscan spectra (Pettersson-Julander et al., 2004).

Update : Later investigations of BFRs in Swedish electronic recycling plants reported DBDPE levels in dust from < 0.02 – 0.79 ng/m3 (Julander et al., 2005).

1,2 – bis(2,4,6 – tribromophenoxy)ethane (BTBPE) Note : This is also referred abbreviated as TBE in the literature.

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Br Br

O CH 2 CH 2 O

Br Br Br Br

Characteristics of the compound: CAS No.: 37853-59-1 CA Index name: Benzene, 1,1'-[1,2- ethanediylbis(oxy)]bis[2,4,6- tribromo- Molecular Formula: C14 H8 Br6 O2 Molecular Weight (g/mol) 687.64 a Melting Point/range ( °C) Not available Boiling Point/range ( °C) 566.4±50.0 a Vapour Pressure (Pa (25 °C)) 3.88E-10a Water Solubility (g/l (25 °C)) 1.90E-05 a a Partition Coefficient (log P ow ) 7.880±0.863 a Partition Coefficient (log K oc ) 5.66

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04).

Toxicity: Oral exposure of 1,2-bis(2,4,6-tribromophenoxy)ethane (BTBPE) to rat and dog showed weak acute toxicity (LD > 10g/kg body wt), no obvious effect was seen for rats exposed to BTBPE in the diet for 14 days (Nomeir et al., 1993;RTECS, 2008). Inhalation exposure of BTBPE on rat showed an LC of > 36.68 g/m 3/4 hours, with effects such as behavioral and gastrointestinal changes and dermatitis. Inhalation exposure during three weeks gave a TCLo of 20 g/m 3/4hours/3weeks with effect on lungs, thorax, or respiration (RTECS, 2008). Dermal administration of BTBPE to rabbit showed an LD > 10 g/kg body wt with nutritional and gross metabolic changes (RTECS, 2008).

Update: Hardy et al. report a reference dose of 0.243 mg/kg/day for children based on U.S. EPA recommended composite and a lowest reported no-observed-adverse-effect level NOAFL (Hardy et al., 2008).

Bioaccumulation, degradation and fate: A study of juvenile rainbow trout exposed through the diet to BTBPE for 49 days followed by a 154 days depuration showed a linear uptake rate and elimination phase with a depuration half-life of 54.1 days. The determined biomagnifications factor of 2.3 suggested that this chemical have a high potential for biomagnification in the aquatic food web. No metabolites were detected. Biochemical results indicate that BTBPE is not a potent thyroid axis disruptor (Tomy et al., 2007). Biomagnification between trophic levels of the Lake Winnipeg (Canada) food web resulted in a BMF of 0.1-2.5 for BTBPE (Law et al., 2006;Law et al., 2007).

Update : An exposure study of children to BTBPE in childrens toys estimated that, within the sample study, exposures were 24.2 to 822 pg/kg bw-day (Chen et al., 2009).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Environmental levels: A study of trophic levels of Lake Winnipeg (Canada) food web found BTBPE with the highest concentration observed in mussels (mean concentration of 1.29 ng/g lipid wt). BTBPE was also detected in walleye, whitefish and zooplankton. In water the median concentration was 1.96 pg/l (dissolved phase) while the concentration in sediments was below the method detection limit (Law et al., 2006; Law et al., 2007). BTBPE have been found in dust collected from 19 homes in Boston (U.S.) and ranging from 1.6-789 ng/g (Stapleton et al., 2008). BTBPE was also found in herring gull eggs around the Great Lakes in 2004, suggesting that new types of BFRs are present in the aquatic food web of the Great Lakes (Verreault et al., 2007; Gauthier et al. 2008). Lake Ontario sediment cores was investigated and BTBPE was found in the surficial sediment, with average concentration of 6.7 ng/g dry wt (Qiu et al., 2007). BTBPE was also detected in tree bark from the same region in the Northeastern U.S. ranging from ND to 0.62 ng/g bark (Qiu and Hites, 2008). BTBPE have also been reported in the particulate phase in air near the Great Lakes with median concentrations from 0.5 to 1.2 pg/m 3 (Venier and Hites, 2008). BTBPE was detected in air around the Great Lakes in 2002-2003, the authors concluded that these findings correlates with sources of known manufacturing of flame retardants in southern Arkansas. In these regions higher levels of DecaBDE, HBCDD and BTBPE have been observed in air (Hoh and Hites, 2005).As BTBPE is starting to replace the PentaBDE mixture (Potrzebowksi and Chance, 2004), the importance of monitoring environmental levels is quite significant. The ban of marketing the Penta- and Octa-BDE mixtures (European Union, 2003) may increase the use of BTBPE as well as DBDPE. This may increase the levels found of these compounds in the domestic environment and biota.

Update : Many studies that measure BTBPE usually also measure DBDPE, and the two appear to usually be co-occurring in the environment, though levels are not necessarily similar. In many studies, DBPDE levels dominate by 1 – 2 orders of magnitude; however, other studies have reported situations where BTBPE levels dominate. Thus, their appearance in the environment does not seem to be coupled, though they are more commonly analysed for in environmental studies of BFR concentrations. Regarding recently reported environmental concentrations, several new reports on BTBPE presence has been reported. For a 1979-2004 time series of lake trout in the Great Lakes, the concentration increased from 0.6 ± 0.3 ng/g lipid to 1979 to 2.6 ± 0.6 ng/g lipid in 1993 (with a doubling time of six years), and since then levels declined about 40% (Ismail et al., 2009), these levels were on average greater than in Lake Winnipeg mentioned above. In other samples from Great Lake biota, BTBPE levels in herring gull eggs ranged steadily across samples from 0.27 – 0.66 ng/g w.w. In diverse samples from the Pearl River Delta, levels in sediment ranged from 0.3 – 22 ng/g d.w, in sewage sludge from 0.3 – 1.7 ng/g d.w, in farmland soil from 0.02 – 0.11 ng/g d.w, and in air from 0.3 - 1.66 pg/m3 (Shi et al., 2009). From an e-waste processing area in this region, samples ranged from 15 – 110 ng/g d.w in dust, 0.07 – 6.2 in soil, 0.07 – 2-4 in diverse bird organs, and from < 0.012 – 0.15 in diverse marine life samples (Shi et al., 2009). An earlier tree bark study, not mentioned in the previous report found that BTBPE in parc samples could range from not detected (<0.5 ng/g lipid) to 24 ± 5 ng/g lipid (Zhu and Hites, 2006)). A study in different indoor environments in Birmingham, UK (cars, offices, and houses) found dust concentrations of < dl to 1900 ng/g (with average in cars, offices and houses being 7.7, 7.2 and 120 ng/g, respectively) (Harrad et al., 2008) (thus, compared to the same studies report on DBDPE, BTBPE levels are much lower in cars and offices, but similar in houses).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Emissions and monitoring data in the Nordic countries: BFRs were determined in air, sedimentary dust and plasma from five households in Sweden. BTBPE was not detected in plasma or air but in sedimentary dust. This is similar to DBDPE, and at the same concentrations as several of the most abundant BFRs (Karlsson et al., 2007). Measurments of outdoor air was below LOQ (Sjodin et al., 2001). To evaluate exposure to BFR in an electronic recycling facility, personal air monitoring was done for 2 years. A total of 22 PBDEs and 2 other Br-containing organic compounds were analyzed and evaluated in 17 personal air samples (Pettersson-Julander et al., 2004). One of the compounds was identified as BTBPE based on full scan spectra and previous identifications (Sjodin et al., 2001), the other compound was tentatively identified as DBDPE. BTBPE was the second most abundant compound of the BFRs of all samples and was semiquantitatively determined to <0.6-39 ng/m 3 (Pettersson-Julander et al., 2004).

BTBPE was found in small amounts in northern fulmar eggs from the Faroe Islands (North Atlantic) with a mean concentration of 0.11 ng/g lipid wt. This concentration is 150 times lower than the sumBDEs (Karlsson et al., 2006). BTBPE have also been found in eggs and plasma from glaucous gulls in the Norwegian arctic. Only low concentrations of BTBPE (max 0.96 ng/g lipid wt) were found in egg yolk and in only one plasma sample. Results suggested that BTBPE and other non-BDE BFRs may undergo long-range atmospheric transport to arctic regions, bioaccumulate (at low concentrations) and are maternally transferred to eggs (Verreault et al., 2007).

Update : Another study on air particles in Swedish electronic-waste recycling plants reported elevated levels of BTBPE compared to DBDPE, with BTPE levels ranging from <0.71 – 12.15 ng/m 3 (DBDPE was < 0.02 – 0.79 – see above) (Julander et al., 2005) Regarding the study on levelsof BTBPE reported in household air and dust, and in relation to blood samples (Karlsson et al., 2007), BTBPE was not found at the elevated levels reported earlier. Rather, it was not dectected in the vapour phase, nor in plasma samples, but could range in dust levels from 2.7 – 8.2 ng/g dust d.w

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Ethylene bis(tetrabromophtalimide) (EBTPI)

Br Br O O Br Br N CH 2 CH 2 N Br

Br O O Br Br

Characteristics of the compound: CAS No.: 32588-76-4 CA Index name: 1H-Isoindole-1,3(2H)-dione, 2,2'-(1,2-ethanediyl) bis[4,5,6,7-tetrabromo- Molecular Formula: C18 H4 Br8 N2 O4 Molecular Weight (g/mol) 951.47 a Melting Point/range ( °C) Not available Boiling Point/range ( °C) 827.9±65.0 a Vapour Pressure (Pa (25 °C)) 1.97E-25a Water Solubility (g/l (25 °C)) 2.60E-05 a, 3.0E-12 b a b Partition Coefficient (log P ow ) 7.561±0.855 , 9.80 a Partition Coefficient (log K oc ) 5.49

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04) b) Estimated in EPIWin (Hardy et al., 2002)

Toxicity: Oral administration of EBTPI to rat gave an LD50 of >7500 mg/kg, effects was not reported (RTECS, 2008). EBTPI has a NOEL in rat and rabbit of 1000 mg/kg body wt during an 90 day study (US EPA, 2004b).

An inhalation study on rat gave a LC >203 g/kg/m 3/1 hour, with effects on sense organs (olfaction) and the respiratory system (dyspnea and pulmonary emboli). EBTPI was dermally administered to rabbit, with a LD50 >2g/kg (effects not reported) (RTECS, 2008).

Developmental or reproductive toxicity where studied using Saytex BT 93 (EBTPI) and administered to two groups of 20 mated female New Zealand white rabbits each by gavage at a dose of 0 or 1,000 mg/kg/day on gestation days 7-19. The females were sacrificed on gestation day 29 and subjected to a cesarean section. No maternal mortality, abortions or clinical signs of toxicity were observed during the study. Maternal body weights, weight gain, food consumption, necropsy observations and cesarean section data were generally comparable among the groups. No treatment-related malformations or developmental variations in the fetuses were observed. The maternal and fetal NOEL was 1,000 mg/kg/day (TOXNET, 2008).

Bioaccumulation, degradation and fate: EBTPI was not readily biodegradable by activated sewage sludge over a 2-day period when tested under Japanese MITI/OECD Ready Biodegradability 301C Modified MIT1 guidelines (US EPA, 2004b). The BCF was studied

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) in the Japanese Carp according to OECD guidelines (305C). No bioconcentration was observed for EBTPI during the 6 week study (BCF = <0.3 to <33) (Hardy, 2004).

Environmental levels: No environmental levels have to our knowledge been reported for EBTPI.

Emissions and monitoring data in the Nordic countries: No data available.

Update : No further updates on this compound, nor its presence, could be found in the literature.

Tetrabromophtalic anhydride

Br O Br O

Br O Br

Characteristics of the compound: CAS No.: 632-79-1 CA Index name: 1,3-Isobenzofurandione, 4,5,6,7-tetrabromo- Molecular Formula: C8 Br4 O3 Molecular Weight (g/mol) 463.70 a Melting Point/range ( °C) 274-275 b, 278-280 b; 279.5-280.5 b Boiling Point/range ( °C) 540.5±50.0 a Vapour Pressure (Pa (25 °C)) 1.27E-9a Water Solubility (g/l (25 °C)) 2.4E-02 a a Partition Coefficient (log P ow ) 3.779±0.698 a Partition Coefficient (log K oc ) 3.43

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04) b) Data from SciFinder originating from experimentally determined properties

Toxicity: Oral administration to rat resulted in a LD 50 >10g/kg, no details on effects were reported (RTECS, 2008). An inhalation study on rat of tetra-bromophtalic anhydride resulted in a LC of >10.92 g/m 3/4 hours and toxic effects resulting in changes in motor activity. Another study showed an LCLo of 2 g/m 3/4 hour/3 weeks (intermittent), with pathologial effects on lung weight, changes in structure or function of salivary glands and changes in liver weight. Another study resulted in a lowest observed toxicity of 50ug/m 3/5 days (intermittent), resulting in a increase in humoral immune response (RTECS, 2008). Dermal administration of tetrabromo-phthalic anhydride on rabbit resulted in a determined LD 50 of

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

>10 g/kg with behavioural effects such as somnolence and a general depressed activity. Similarly, dermal administration on guinea pig resulted in a LD 50 of >1 g/kg (no details on effect). Another study on rabbit resulted in a lowest observed toxicity of 100 g/kg/4 weeks (intermittent), resulting in behavioural effects such as ataxia, weight loss or decreased weight gain and death (RTECS, 2008).

Bioaccumulation, degradation and fate: tetrabromophtalic anhydride was proposed to be one of 120 high production chemicals which are structurally similar to known arctic contaminants and/or have partitioning properties that suggests they are potential arctic contaminants (Brown and Wania, 2008).

Environmental levels: No reported studies in the peer reviewed literature.

Emissions and monitoring data in the Nordic countries: No data available.

Update : No further updates on this compound, nor its presence, could be found in the literature.

Bis(2 – ethylhexyl)tetrabromophtalate) (TBPH)

Et O O Et n-Bu CH CH 2 O C Br C O CH 2 CH Bu-n

Br Br Br

Characteristics of the compound: CAS No.: 26040-51-7 CA Index name: 1,2-Benzenedicarboxylic acid, 3,4,5,6-tetrabromo-, 1,2-bis(2-ethylhexyl) ester Molecular Formula: C24 H34 Br4 O4 Molecular Weight (g/mol)) 706.14 a Melting Point/range ( °C) Not available Boiling Point/range ( °C) 584.8±45.0 a Vapour Pressure (Pa (25 °C)) 1.55E-11a Water Solubility (g/l (25 °C)) 1.60E-06 a a Partition Coefficient (log P ow ) 10.084±0.938 a Partition Coefficient (log K oc ) 6.86

a) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04)

Toxicity: No data available

Update: Hardy et al. (Hardy et al., 2008) proposed estimating the exposure risk to children using earlier derived studies on bis(2-ethylhexyl)-phthalate (DEHP) as a surrogate standard, and proposed a response factor at 20 µg/kg body weight/day. This was criticized by Staplenton et al.

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

(Stapleton et al., 2008), who among other reasons argued that non-brominated compounds do not make good surrogates for brominated compounds.

Bioaccumulation, degradation and fate: No data available.

Update : Photodegradation of TBPH in various solvents and with natural sunlight was investigated (Davis and Stapleton, 2009). The halflife was notably slower than for PBDEs, ranging from 145 – 220 min (as opposed to 3.6 – 13 minutes for diverse nonaBDEs and deca BDE). Debromination products were primarily formed.

Environmental levels: Dust collected from 19 homes in Boston, Massachusetts, showed TBPH and the decarboxylated form of TBPH which was identified to be 2-ethylhexyl- 2,3,4,5-tetrabromobenzoate (TBB) and the dominant brominated compound in Firemaster 550. TBPH was detected in house hold dust at ranges 1.5-10600 ng/g dust wt at levels comparable to HBCDD (Stapleton et al., 2008). TBPH and TBB was further detected in biosolids from two San Francisco bay area waste water treatment plants (WWTP) ranging from 40 to 1412 ng/g dry wt and 57 to 515 ng/g dry wt, respectively (Klosterhaus et al., 2008). Firemaster 550 is a replacement product for PentaBDE (Venier and Hites, 2008;Chemtura, 2008) and was introduced to the market in 2003 (Stapleton et al., 2008).

Update : Subsequent studies on household dust found levels of < 300 – 47,110 ng/ g dw, with a geometric mean of 659 ng / g dw. In a study on marine mammals in Hong Kong, levels ranged from < 0.04 to 5.3 ng / g lw in humpack dolphin (median 0.51 ± 0.13 ng / g lw), and from < 0.04 – 3859 ng / g lw (342 ± 883 ng / g lw) in finless porpoise (Lam et al., 2009).

Emissions and monitoring data in the Nordic countries: No data available.

2 – ethylhexyl – 2,3,4,5 – tetrabromobenzoate (TBB)

Note : Referred to also as EHTBB.

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Characteristics of the compound: CAS No.: 183658-27-7 CA Index name: a) SPARC online calculator (accesses February, 2010 at http://sparc.chem.uga.edu/sparc/search/searchcas.cfm)

Toxicity: Hardy et al. (Hardy et al., 2008) proposed estimating the exposure risk to children using earlier derived studies on bis(2-ethylhexyl)-phthalate (DEHP) as a surrogate standard, and proposed a response factor at 20 µg/kg body weight / day, similar to the above estimation for TBPH. This was criticized Staplenton et al. (Stapleton et al., 2008).

Bioaccumulation, degradation and fate: Photodegradation of TBB in various solvents and with natural sunlight was investigated (Davis and Stapleton, 2009). The half-life was notably slower than for PBDEs and only slightly slower than TBPH (see above), ranging from 86 – 163 minutes.

Environmental levels: Dust collected from 19 homes in Boston, Massachusetts, showed TBPH and the decarboxylated form of TBPH which was identified to be 2-ethylhexyl- 2,3,4,5-tetrabromobenzoate (TBB) and the dominant brominated compound in Firemaster 550. TBB was detected in house hold dust at ranges 1.5-15030 ng/g dust (geometric mean of 322 ng / g dust) at levels comparable to HBCDD (Stapleton et al., 2008). In a follow up study, the range of levels in diverse houses ranged from < 450 to 75,000 ng / g dust, with a median of 840 ng / g dust.

TBB was further detected in biosolids from two San Francisco bay area waste water treatment plants (WWTP) ranging from 57 to 515 ng/g dry wt, respectively (Klosterhaus et al., 2008). Firemaster 550 is a replacement product for PentaBDE (Venier and Hites, 2008;Chemtura, 2008) and was introduced to the market in 2003 (Stapleton et al., 2008).

In a study on marine mammals from Hong Kong, levels of TBB were below detection in humpback dolphin and range from < 0.04 to 70 ng / g lw (with a mean of 5.6 ± 17 ng/g lw) for porpoise (Lam et al., 2009).

Emissions and monitoring data in the Nordic countries : No data available

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

12. Appendix IV – Description of PFCs included in the screening

The PFCs looked in this screening include PFSAs of various chain-length and 6:2 FTS. Some introductory information of these particular compounds was provided in Chapter 3.3 and references therein. Here more additional specific information will be given on the following topics:

● Chemical structure and physical data ● Toxicity data ● Bioaccumulation, degradation and fate ● Environmental levels (published data) ● Emissions and monitoring data from Norway

Note that as these chemicals are all structurally similar and tend to have similar sources, the information will be presented for all compounds simultaneously, rather than individually, as in the case of the new BFRs.

Unlike the neutral BFRs, PFSAs are ionic surfactants. Accounting for the environmental partitioning behavior is much more challenging than neutral compounds, as these compounds can exist in several different species and as such can sorb substantially to components that neutral compounds do not sorb substantially to. These include the air-water interface, proteins and certain mineral surfaces. Thus, the list of relevant physical-chemical properties needed to characterize the partitioning behavior of PFSAs is inherently longer, and more complex, than BFRs.

Regarding the different species that PFSAs can exist as in the environment: 1) they can form salts with various cations ( Error! Reference source not found. a), and depending on the cation different relevant physical chemical properties of the PFSA species (e.g. vapour pressure) will result; 2) in aqueous environments PFSAs can exist in a neutral acidic form or as a negatively charged ion, e.g. via the following acid dissociation equation for PFOS.

+ C8F17 S O 3-H + H 2O → C 8F17 S O 3- + H 3O + (PFOS + H 2O → PFOS- + H 3O ) (x)

With a pKa that is most likely < 0, PFSAs are such strong acids that they (as single molecules) are assumed to exist in ambient water almost entirely as negatively charged ions (unlike TBA in which both the neutral and anionic form can be found in abundance in the environment). Thirdly, to further complicate the situation, it is also possible that PFSAs like other surfactants can form aggregates (multi-molecular structures which share protons or cations) and micelles under environment conditions. This occurs in aqueous solution once certain (pH and ionic strength dependant) threshold concentrations are reached, such as the critical micelle concentration (cmc), at which micelles are formed. The formation of pre- micellular aggregates at very low concentrations has been identified as a critical issue when dealing with PFCAs (Arp and Goss, 2009; Cheng et al., 2009). Research so far for PFSA in pure water, however, indicates that micelle formation and aggregation does not occur to a noticeable extent, as these compounds become insoluble before the critical micelle concentration (cmc) is reached (Campbell et al., 2009). Further, no aggregates have been

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010) observed for a study that considered a wide concentration range (Cheng et al., 2009). In any case, until this is ruled out for natural waters, such as sea water, marine aerosols and precipitation, the possibility of micelle formation in the environment cannot be ruled out.

a) b)

Figur 21. Examples of potential, alternative forms of PFSAs in the environment, such as a) salts(Image obtained from http://chm.pops.int/ ), and b) an micelle such as on the right (image obtained from http://en.wikipedia.org/wiki/Micelle ), which are formed by surfactants once the criticial micelle concentration (cmc) is exceeded (pre-micellular aggregates may form at lower concentrations).

To quantify partitioning between the air and water phase, the octanol and water phase or between any phase and water when it comes to ionizing species, it is necessary to account for the partitioning of both the neutral and ionic forms of the molecule. For organic acids, such as PFSA and TBP, the octanol-water and air-water distribution coefficients of both the ionic and neutral forms can be calculated as:

pH-pKa -1 Dow = Pow (1 + 10 ) (x) pH-pKa -1 Daw = Paw (1 + 10 ) for organic acids (x)

Where “D” is used instead of “P” to signify we are talking about a multiple species distribution, with D ow refers to the simultaneous octanol-water distribution of a compound that exists in the neutral and ionic form in water but only in the neutral form in octanol, and Daw refers to the simultaneous air-water distribution of a compound that exists in the neutral and ionic form in water but only in the neutral form in air. If the species was known to aggregate or form micelles under certain conditions, it would be ideal to add terms to describe this to D ow and D aw as well, however, these conditions are currently unknown.

Further, as partitioning to the water surface can be highly substantial in the environment for perfluorinated surfactants (Arp and Goss, 2009; Psillakis et al., 2009), it is necessary to additionally quantify air-water interface partitioning. For surfactants, this is typically described with a Langmuir isotherm:

K c c === ΓΓΓ L water (4) surface MAX +++ 1 K Lc water

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

2 Where Γmax is the maximum surface coverage (mol/m water surface ), K L is the Langmuir 3 sorption constant (m water /mol), c surface is the concentration adsorbed at the air/water interface 2 3 (mol/m water surface ) and c water is the concentration in the bulk water phase (mol/m water ).

Due to difficulties in experimentally determining the environmentally relevant physical chemical parameters for PFSAs, only a handful of high-quality data exists. Further, due to the limited amount of such data for PFCs in general, it is still an open question how good traditional chemical property prediction models are for estimating the properties of these compounds, as it was with BFRs. Thus, all estimated values (and even some experimental values) presented here and elsewhere in the literature must be treated with caution. A selection of the best known experimental data, and a selection of some estimated data, is presented in

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 50 and Table 51 respectively. More information on challenges and issues on estimating the physical chemical properties of PFSAs can be found in several publications in the literature, particularly by Goss et al (e.g. Arp and Goss, 2009; Arp et al., 2006; Goss, 2008; Goss and Arp, 2009; Goss and Bronner, 2006) and Rayne and Forest (e.g. Rayne and Forest, 2009b, c; Rayne et al., 2009). Similarly, active discussions about what this uncertainty means for the understanding of the environmental fate of these compounds are currently ongoing (e.g. Armitage et al., 2009b; Rayne and Forest, 2009a).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Table 50 Experimentally determined properties for PFSAs (note, the data in this table is dependent on ionic strength, temperature, pH and co solvents)

d e e CAS Name Formula pK a cmc Cmax Kl Γ pis Cwsat log K aw log K oc

-1 (mM) (mM) (M ) (mol/m 2) (Pa) (mM) (neutral) (ionized)

a 375-73-5 PFBuS C4HF 9O3S 300.10 < 1 223 4.04E+01 4.40E-06 n.d. n.d.

a 355-46-4 PFHxS C6HF 13 O3S 400.11 < 1 57 1.58E+02 5.70E-06 n.d. n.d.

a b h g 1763-23-1 PFOS C8HF 17 O3S 500.13 < 1 1.1 4.8 1.87E+03 5.00E-06 n.d. n.d. 2.57 - 3.3

c f g 2795-39-3 PFO-Ka C8F17 O3S-K 538.23 2.0 3.3E-4 0.56

a h 335-77-3 PFDS C10 HF 21 O3S 600.14 < 1 0.85 1.06E+04 5.00E-06 n.d. n.d. 3.53

a 27619-97-2 6:2 FTS C8H5F13 O3S 428.17 < 1 a) Measurements by Cheng et al (Cheng et al., 2009) confirmed the pKa of PFOS < 1 (determining pKa below 1 is experimentally difficult). The remaining PFSA are assumed to be < 1 out of analogy, as CF 2 groups more than 3 CF 2 units away should not affect the acidity, based on analogy considerations from (Goss, 2008); b) Critical micelle concentration (cmc) (Gente et al., 2000); note, at pH = 7 and pure water maximum solubility levels, c max , occur below the cmc (Campbell et al., 2009), though levels reported cmc levels here are below the reported c max . c) Value quoted from the chemical distributor Fluka (Yuan et al., 2001), see note for b. d) maximum estimated concentrations, from Cambell et al (Campbell et al., 2009), note value for PFDS is extrapolated here assuming a correlation between log K L and the number of -(CF2)- groups. e) from Cambell et al (Campbell et al., 2009), PFDS value derived here as with c max . f) vapour pressure of solid, from Giesy, J.; Mabury, S.; Martin, J.; Kannan, K.; Jones, P.;Newsted, J.;Coady, K. Perfluorinated compounds in the Great Lakes. Persistent Organic Pollutants in the Great Lakes; Hites,R., Ed.; Springer: New York, NY, USA, 2006; 391–438. g) solid (not subcooled liquid) solubility (Pan et al., 2009). h) from Higgens and Luthy (Higgins and Luthy, 2006).

Table 51 Selection of some estimated environmentally relevant physical-chemical properties for PFSAs and 6:2 FTS.

e) e) Name pK a p* iL CwLsat log K aw (-) log D aw (-) log K ow (-) log D ow (-) SPARC b) SPARC b) SPARC b) SPARC b) ACD d) SPARC b) SPARC b)

a) b) c) neutral neutral neutral ionic neutral neutral ionic Analogy SPARC PM6 (Pa) (mM) (pH < -3) (pH=7) (pH < -3) (pH < -3) (pH=7)

PFBuS -3 0.14 -5.5 73.2 489.78 -4.22 -11.4 3.68+/- 0.91 1.25 -5.89

PFHxS -3 0.14 -5.5 27.3 6.03 -2.74 -9.9 5.25+/- 0.97 3.27 -3.87

PFOS -3 0.14 -5.5 13 0.030 -0.76 -7.9 7.03+/- 1.01 5.5 -1.64

PFDS -3 0.14 -5.5 6.94 5.6E-05 1.7 -5.4 8.81+/- 1.05 8.09 0.95

6:2 FTS -3 0.36 n.d. 0.31 12.30 -3.99 -11.4 3.47+/- 0.96 3.98 -3.16 n.d: not detected a) the pKa of PFSAs are commonly assumed to be in the approximately -3, as this is the pKa generally assigned to organosulphonic acid (i.e. R-S(O) 3-H, e.g. www.chem.wisc.edu/areas/organic/index-chem.htm ); though, note the authors have not found the experiments where this value comes from, likely because the pKa of strong acids are quite difficult to measure. b) From the SPARC online calculator, accessed February 2010 ( http://sparc.chem.uga.edu/sparc/ ). c) Rayne et al., 2009 d) Data from SciFinder originating from calculated properties (ACD/labs Software V9.04) e) refers to the distribution of both the neutral and ionic species (see main text).

Environmental screening of selected ”new” brominated flame retardants and polyfluorinated compounds 2009 (TA-2625/2010)

Klima- og forurensningsdirektoratet Postboks 8100 Dep, 0032 Oslo Besøksadresse: Strømsveien 96

Telefon: 22 57 34 00 Telefaks: 22 67 67 06 E-post: [email protected] www.klif.no

Om Statlig program for forurensningsovervåking

Statlig program for forurensningsovervåking omfatter overvåking av forurensningsforholdene i luft og nedbør, skog, vassdrag, fjorder og havområder. Overvåkingsprogrammet dekker langsiktige undersøkelser av:

• overgjødsling • forsuring (sur nedbør) • ozon (ved bakken og i stratosfæren) • klimagasser • miljøgifter

Overvåkingsprogrammet skal gi informasjon om tilstanden og utviklingen av forurensningssituasjonen, og påvise eventuell uheldig utvikling på et tidlig tidspunkt. Programmet skal dekke myndighetenes informasjonsbehov om forurensningsforholdene, registrere virkningen av iverksatte tiltak for å redusere forurensningen, og danne grunnlag for vurdering av nye tiltak. Klima- og forurensningsdirektoratet er ansvarlig for gjennomføringen av overvåkingsprogrammet.

SPFO-rapport xxxx/xxxx TA-xxxx/xxxx