Persistent Organic Chemicals in the Environment - Status and Trends in the Pacific Basin Countries II Temporal Trends 9780841231979, 0841231974, 9780841231993, 0841231990, 9780841231962, 9780841231986

449 111 5MB

English Pages [263] Year 2016

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Persistent Organic Chemicals in the Environment - Status and Trends in the Pacific Basin Countries II Temporal Trends
 9780841231979, 0841231974, 9780841231993, 0841231990, 9780841231962, 9780841231986

Table of contents :
Content: V. 1. Contamination status --
v. 2. Temporal trends

Citation preview

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.fw001

Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II Temporal Trends

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.fw001 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

ACS SYMPOSIUM SERIES 1244

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.fw001

Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II Temporal Trends Bommanna G. Loganathan, Editor Murray State University, Murray, Kentucky

Jong Seong Khim, Editor Seoul National University, Seoul, Korea

Prasada Rao S. Kodavanti, Editor U.S. Environmental Protection Agency, Research Triangle Park, North Carolina

Shigeki Masunaga, Editor Yokohama National University, Yokohama, Japan Sponsored by the ACS Division of Environmental Chemistry, Inc.

American Chemical Society, Washington, DC Distributed in print by Oxford University Press

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.fw001

Library of Congress Cataloging-in-Publication Data Names: Loganathan, Bommanna G., editor. | American Chemical Society. Division of Environmental Chemistry. Title: Persistent organic chemicals in the environment : status and trends in the Pacific Basin countries / Bommanna G. Loganathan, Murray State University, Murray, Kentucky [and three others], editor ; sponsored by the ACS Division of Environmental Chemistry. Description: Washington, DC : American Chemical Society, [2016]- | Series: ACS symposium series ; 1243, 1244 | Includes bibliographical references and index. Contents: v. 1. Contamination status -- v. 2. Temporal trends Identifiers: LCCN 2016054891 (print) | LCCN 2016055072 (ebook) | ISBN 9780841231979 (v. 1) | ISBN 9780841231993 (v. 2) | ISBN 9780841231962 (v.1) (ebook) | ISBN 9780841231986 (v.2) (ebook) Subjects: LCSH: Organic compounds--Environmental aspects--Pacific Area. | Persistent pollutants--Pacific Area. Classification: LCC TD196.O73 P46 2016 (print) | LCC TD196.O73 (ebook) | DDC 628.5/2--dc23 LC record available at https://lccn.loc.gov/2016054891

The paper used in this publication meets the minimum requirements of American National Standard for Information Sciences—Permanence of Paper for Printed Library Materials, ANSI Z39.48n1984. Copyright © 2016 American Chemical Society Distributed in print by Oxford University Press All Rights Reserved. Reprographic copying beyond that permitted by Sections 107 or 108 of the U.S. Copyright Act is allowed for internal use only, provided that a per-chapter fee of $40.25 plus $0.75 per page is paid to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. Republication or reproduction for sale of pages in this book is permitted only under license from ACS. Direct these and other permission requests to ACS Copyright Office, Publications Division, 1155 16th Street, N.W., Washington, DC 20036. The citation of trade names and/or names of manufacturers in this publication is not to be construed as an endorsement or as approval by ACS of the commercial products or services referenced herein; nor should the mere reference herein to any drawing, specification, chemical process, or other data be regarded as a license or as a conveyance of any right or permission to the holder, reader, or any other person or corporation, to manufacture, reproduce, use, or sell any patented invention or copyrighted work that may in any way be related thereto. Registered names, trademarks, etc., used in this publication, even without specific indication thereof, are not to be considered unprotected by law. PRINTED IN THE UNITED STATES OF AMERICA Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.fw001

Foreword The ACS Symposium Series was first published in 1974 to provide a mechanism for publishing symposia quickly in book form. The purpose of the series is to publish timely, comprehensive books developed from the ACS sponsored symposia based on current scientific research. Occasionally, books are developed from symposia sponsored by other organizations when the topic is of keen interest to the chemistry audience. Before agreeing to publish a book, the proposed table of contents is reviewed for appropriate and comprehensive coverage and for interest to the audience. Some papers may be excluded to better focus the book; others may be added to provide comprehensiveness. When appropriate, overview or introductory chapters are added. Drafts of chapters are peer-reviewed prior to final acceptance or rejection, and manuscripts are prepared in camera-ready format. As a rule, only original research papers and original review papers are included in the volumes. Verbatim reproductions of previous published papers are not accepted.

ACS Books Department

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.pr001

Preface Environmental pollution by man-made Persistent Organic Chemicals (POCs) has been a serious global issue for over half a century. Exposure to POCs may result in health effects, including, endocrine disruption leading to birth defects, intellectual disability, low testosterone, childhood obesity, autism and attention deficit hyperactivity disorder (ADHD). Therefore, POCs have been the subject of an intensive regional, national, and international efforts to limit the production, use, and disposal of these chemicals. Since POCs are ubiquitous and recalcitrant, and cause long-term effects on wildlife and humans, trend monitoring studies are valuable in making clear the behavior and fate of these compounds and to protect our environment and living resources. The Pacific Basin is a unique geographical region representing tropical, temperate and polar zones. This region is home to 2/3 of the world’s population and consists of rapidly growing economies (countries) and highly developed countries. Due to this diversity of climatic and socio-economic conditions, environment and biota in different countries in this basin have varying degrees of environmental contamination and effects on wildlife and humans. The Pacific Rim countries play a pivotal role in governing global POC contamination and resulting harmful health effects. Because articles on POCs and their effect on environment and health are published in a large number of different journals, it is useful to have a book that includes original papers and reviews on the latest advances by well-known scientists in the field, especially focusing on the countries in the Pacific Rim. The two volumes of this book satisfies this need. The two volumes are based on the successful symposium on “Status and Trends of Persistent Organic Chemicals in the Environment”. The symposium took place at PACIFICHEM 2015, International Chemical Congress of Pacific Basin Societies, December 15-20, 2015 in Honolulu, Hawaii. The symposium brought together an impressive group of leading experts in the field, covering a broad spectrum of expertise in contamination status and temporal trends of POCs from countries in the Pacific Rim. Eighteen platform presentations and nine posters were presented. The presentations created an exciting and dynamic forum for highlighting current contamination profiles and as well as future trends, which formed the foundation of this two-volume book. All of the symposium speakers were invited to submit chapters to this book. We were pleased that the majority contributed chapters. Other internationally respected researchers contributed additional chapters in order to strengthen the coverage of classical and emerging contaminant statuses and trends in the Pacific Rim countries. A total of 20 chapters are included in the two volumes of the book. Volume 1 focuses on contamination status including human exposure to POCs and Volume 2 focuses on the temporal trends and future perspectives. Topics covered in Volume 1 include an overview of POCs contamination status and trends in the Pacific Basin Countries (Chapter 1); human exposure to brominated flame

ix Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.pr001

retardants (Chapter 2); persistent toxic substances in Vietnam (Chapter 3); dietary exposure to a variety of organohalogen pollutants and heavy metals in Tokyo, Japan (Chapter 4) and Georgia, USA (Chapter 5); e-waste and associated environmental contamination in the Asia/Pacific Region (Chapter 6), including a case study on dioxin and furan exposure to e-waste workers in India (Chapter 7); POCs in sediments (Chapter 8), soil and atmosphere of South Korea (Chapter 9); and new research on sequestration and redistribution of emerging and classical persistent organic pollutants by polystyrene (Chapter 10). Topics covered in Volume 2 include lessons learned from three decades monitoring contaminants in Pacific Basin wildlife samples from the USA’s Marine Environmental Specimen Bank (Chapter 1); spatial and temporal trends of brominated flame retardants (Chapter 2), PCBs, pesticides, and dioxins/furans, in the environment and biota in the USA, Colombia (Chapter 3), China (Chapter 4), Korea (Chapter 5), and Japan (Chapter 6 and Chapter 7); emission of emerging pharmaceutical contaminants in the USA (Chapter 8) and Vietnam (Chapter 10); and possible application of bio-analytical assays in the biological impact assessment of persistent organic pollutants in Mangrove sediments in Southeast Asia with particular reference to Malaysia (Chapter 9). The collection of chapters in these volumes may serve as a reasonable representation of current and future trends of POCs in the Pacific Basin countries. It is hoped that the book can inspire students and researchers, as well as professionals, to facilitate the understanding of the environmental and biological behavior of these persistent chemicals and to help in the development of strategies and practices for protecting the global environment for future generations. We would like to express our gratitude to all of the authors who took time to prepare their manuscripts and to the many reviewers for their valuable time and expertise in the peer review process. Special thanks to Dr. Kevin Miller for his timely help in reviews and language corrections. Thanks are also due to the editorial team at ACS Books, particularly Bob Hauserman, Elizabeth Hernandez, Arlene Furman, and Aimee Greene for their efficient handling of manuscripts.

Bommanna G. Loganathan Department of Chemistry/Watershed Studies Institute, Murray State University Murray, Kentucky 42017, United States Jong Seong Khim School of Earth and Environmental Sciences, Seoul National University Seoul 08826, Republic of Korea Prasada Rao S. Kodavanti Neurotoxicology Branch, NHEERL/ORD, U.S. Environmental Protection Agency Research Triangle Park, North Carolina 27711, United States Shigeki Masunaga Faculty of Environment & Information Sciences, Yokohama National University Yokohama-240-8501, Japan x Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

Chapter 1

Lessons Learned from Monitoring Organic Contaminants in Three Decades of Marine Samples from the Pacific Basin Archived at the USA’s Marine Environmental Specimen Bank Stacy S. Vander Pol,*,1 John R. Kucklick,1 Jennifer M. Lynch,2 Rebecca S. Pugh,1 Jared M. Ragland,1 Jessica L. Reiner,1 Jennifer Trevillian,1 and Michele M. Schantz3 1Chemical Sciences Division,

National Institute of Standards and Technology, Hollings Marine Laboratory, 331 Fort Johnson Rd., Charleston, South Carolina 29412, United States 2Chemical Sciences Division, National Institute of Standards and Technology, Kaneohe, Hawaii 96744, United States 3Chemical Sciences Division, National Institute of Standards and Technology, Gaithersburg, Maryland 20899, United States *E-mail: [email protected]

The USA’s Marine Environmental Specimen Bank (ESB) has archived marine wildlife collections dating back to 1976. Numerous lessons have been learned including collecting the correct species and tissues for environmental contaminant monitoring, developing protocols for mitigating sample contamination, and ensuring that samples can be used for new analytes and techniques. Investigations of organochlorine contaminants in several collections from the Pacific basin for species, regional, and temporal trends revealed that α-hexachlorocyclohexane (HCH) declined for all species/regions and was lowest in samples from Hawaii while polybrominated diphenyl ether (PBDE) 47 significantly increased in Alaskan marine mammals with the highest levels in California sea lions and adult male cetaceans that stranded in Hawaii. Chlordanes and dichloro-diphenyl-trichloroethanes (DDTs) declined except for beluga whales, and polychlorinated Not subject to U.S. Copyright. Published 2016 by American Chemical Society Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

biphenyls (PCBs) significantly declined for only common and thick-billed murres from St. George Island, Alaska and common murres from St. Lazaria Island, Alaska. The Marine ESB is also in the process of ensuring easy access to sample information and previous analytical results for other researchers to use this invaluable resource.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

Introduction Trend monitoring studies are invaluable in describing historical and current contamination and modeling future trends by legacy persistent organic pollutants (POPs), e.g. polychlorinated biphenyls (PCBs) and past-use organochlorine pesticides, as well as contaminants of emerging concern (1, 2). Determining trends in environmental contaminants requires access to quality samples, and especially in the case of contaminants that were not historically monitored, environmental specimen banks (ESBs) are a logical choice for obtaining samples for this purpose. ESBs are facilities that participate in long-term preservation of environmental samples. There is an international network of ESBs (www.inter-esb.org) and several journal issues have been devoted to the description and role of ESBs in environmental monitoring [Science of the Total Environment 1993 Vol. 139–140, Chemosphere 1997 Vol. 34(9–10), Journal of Environmental Monitoring 2006 Vol. 8(8), Interdisciplinary Studies on Environmental Chemistry 2010 Vol. 4]. ESBs have thus been shown to be valuable resources of samples to investigate trends of contaminants as a function of sample type, species, location and time. Monitoring contaminants in the marine environment is challenging due to the remoteness and the difficulty of obtaining samples. The USA’s Marine ESB has sample collections dating back to 1976 for mussels and oysters, 1987 for marine mammals, 1999 for seabirds and since 2011 for sea turtles. However, many of these collections are opportunistic. Descriptions of the Marine ESB and these programs (except for the sea turtles) were previously reported by Pugh et al. (3) Currently the Marine ESB collection holds over 100,000 aliquots in liquid nitrogen vapor-phase freezers (-150 °C) from 12,000 animals. Figure 1 provides a general overview of the Marine ESB programmatic sampling locations. Lessons learned from monitoring organic contaminants from some of these marine samples banked at the Marine ESB are presented here.

2 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

Figure 1. Marine Environmental Specimen Bank (ESB) programs and generalized sampling locations.

The Right Species and Tissue Choice Is Important In monitoring contaminant trends, careful choice of species and tissue is crucial to ensure the sample set accurately addresses study goals. For instance, if a researcher is interested in recent contaminant exposure, blood would be a better tissue than blubber/fat, which is more reflective of long-term contaminant exposure except in cases of re-mobilization due to starvation (4, 5). Likewise, if a researcher is interested in contaminants from a specific region, a non-migratory species should be chosen, or the migratory species should be sampled later in the season to ensure the majority of contaminants found in the sample are representative of the study region (e.g. Kucklick et al. (6)). Several guides to choosing species and tissues for biomonitoring have been published previously (7–9). Provided below are examples from the Marine ESB on why the choice of species and tissue can be extremely important when assessing environmental contaminant trends. One component of the Marine ESB, the Seabird Tissue Archival and Monitoring Project (STAMP) began collecting eggs of murres (Uria spp.) and black-legged kittiwakes (Rissa tridactyla) in 1999. The rationale and protocols for these collections was described in detail by York et al. (10) Briefly, murres were chosen as deep diving with a clutch of only one large egg that that may be replaced if lost early in the season, so eggs were collected as early as possible. Kittiwakes were chosen to represent a different guild; surface-feeders that prey on 3 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

small fish and euphausiids and lay up to three smaller eggs that were to be banked as a clutch. In 2004, glaucous gull (Larus hyperboreus) and glaucous-winged gull (L. glaucescens) eggs were added in partnership with the Bureau of Indian Affairs - Alaska Regional Subsistence Branch. Gulls are opportunistic predators and scavengers (including marine mammals and human refuse) and lay clutches of two or three eggs, which are an important food source for many Alaskan Natives. While contaminant information on gulls is important from a human health perspective and to establish baseline ranges of ecological values, using gull eggs for long-term biomonitoring is not advisable. The inability to always archive entire clutches (e.g. collection before the entire clutch was laid, after wild predation, or egg breakage in transport or processing) can lead to difficulty in accounting for laying order effects on contaminants (11). The wide range of prey consumed by gulls also results in more variable contaminant concentrations, making the interpretation of spatial and temporal trends more challenging. Life history is a major influence on contaminant burdens (12, 13). A priori knowledge of the questions that may be asked of a sample set is crucial for determining the correct species and tissue to sample as well as the metadata that needs to be recorded for future researchers to be able to properly choose samples from the specimen bank.

Elimination of All Sample Contamination Is Impossible, So Protocols Should Be Well Documented and Blanks Created Developing standardized protocols, Standard Operating Procedures (SOPs), or Standard Operating Guidelines (SOGs), is critical to the success of long-term ESB programs. The emphasis is to maintain, through optimal long-term preservation, collections of high-quality specimens that can be used for deferred analysis and evaluation (14). To provide the highest quality specimens for continuous time-trend studies and spatial monitoring based on materials collected repeatedly over a long period of time (i.e., decades), all procedures must follow standardized protocols. When developing SOPs many criteria should be considered, including the type of sample to collect, as previously discussed; how, if necessary, the sample should be processed; and how it should be stored (e.g. container type and storage condition). A pilot study for testing protocols should be conducted to determine if the protocol is feasible and if pre-analytical variables will have an effect on sample collection and processing. The success of the collection and storage protocols that are ultimately written depend on the success of the pilot study. All written protocols should also be updated on a regular basis (e.g. annually), made readily accessible for anyone providing samples to the ESB, and to ensure compliance by the contributor, a training program should be established. A well-planned and executed SOP or protocol can determine the success of an ESB by standardizing specimen handling procedures and ensuring uniformity and reproducibility of the processed specimens. In any long-term study, successful completion absolutely requires strict adherence to continuity of defined protocols. 4 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

Taking into consideration the field collection conditions as well as the possibility of extraneous sources of contamination, it is also recommended that a ‘field blank’ or ‘reference blank’ also be taken during the time of sample collection. Clean water from a water purification system can be collected in place of the tissue or fluid sample and should be handled in the exact same manner that the sample was collected, processed, and stored. Polytetrafluoroethylene (PTFE)-based plastic containers (e.g., TeflonTM) have been widely used as a storage container for archived samples as this material is inert and can withstand liquid nitrogen temperatures. PTFE-based plastics were also thought to be non-contaminating to the samples. However, since PTFE-based plastics were first widely used in ESBs, there have been numerous studies showing that PTFE-like compounds are pervasive in wildlife arising through food web exposure (see review by Reiner and Place (15)). This poses a challenge for the Marine ESB and other banks that have stored and processed samples using PTFE-based products. A thorough study was undertaken to estimate the amount of perfluorinated alkyl acids (PFAAs; historically the major component of PTFE-based plastics) that may have contaminated the samples. Negligible amounts of perfluorooctane sulfonate (PFOS) and perfluorononanoic acid (PFNA; approximately 0.2 ng/g and 0.1 ng/g, respectively) were found to leach from materials used during sample processing, but perfluorooctanoic acid (PFOA) was estimated to leach up to 1 ng/g (16). While this PFAA contamination from sample processing and storage is minimal, in the Marine ESB, some samples are now being stored in both PTFE-based plastics and polypropylene containers to eliminate some of the PFAAs contamination concerns, but still allow for organic analysis, especially of phthalates, for future temporal trend studies.

There Will Always Be Something New To Study An average of 4000 new chemicals are added to the Chemical Abstracts Service (CAS) REGISTRY (www.cas.org) every day (17). While only a small percentage of these chemicals are expected to be produced in high enough volume and have the physical-chemical properties to become persistent, bioaccumulative, toxic compounds (PBTs) (18, 19), this still leaves hundreds of new compounds each year that could be PBTs. Samples from the Marine ESB have been analyzed retrospectively for measuring several classes of contaminants of emerging concern and should also prove useful for future unknown PBTs as well. PFAAs have over 200 industrial and commercial uses, the most well-known being stain-resistance (20), and have been produced since the 1950s but were not known to be PBTs until the early 2000s. Reiner and Place (15) reviewed the studies of PFAAs in wildlife including retrospective studies of numerous species from international specimen banks. From the Marine ESB, beluga whales (Delphinapterus leucas) (21), northern fur seals (Callorhinus ursinus) (22), and five species of sea turtles (23, 24) have had retrospective studies published. Seabird eggs have also been analyzed for PFAAs, but many were below the detection limit (25). 5 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

Flame retardants have also been used for decades and were known to be PBTs in the 1980s (26). Brominated flame retardants (BFRs) include polybrominated diphenyl ethers (PBDEs) and hexabromocyclododecanes (HBCDs). Several reviews provide more in-depth information on flame retardants and in particular BFRs (26–30). While others were reporting environmental concentrations of BFRs in the early 1990s, NIST did not begin to focus on BFRs until after assigning values to Standard Reference Materials (SRMs) (31). Since then, samples from various species in or associated with the Marine ESB have been analyzed for BFRs, including white-sided dolphins (Lagenorhynchus acutus) (32, 33), California sea lions (Zalophus californianus) (34), beluga whales (35), northern fur seals (22), common and thick-billed murres (Uria aalge and U. lomvia) (36–40), glaucous and glaucous-winged gulls (37, 38, 41), five species of sea turtles (16, 42–47), and 16 species of cetaceans that stranded in Hawaii (48). In the near future, chlorinated and phosphorus flame retardants as reviewed by Marvin et al. (49) will also be examined. Beyond chemical contaminants, measurement methods are continually being developed to detect disease agents and screen for far-reaching changes to biological systems through omics techniques. Cryogenically stored tissue samples can be useful for these research topics and a sea turtle tissue bank has been developed with these uses in mind. The collection and banking of sea turtle tissues from the Pacific Basin began in 2011 under a project known as Biological and Environmental Monitoring and Archival of Sea Turtle Tissues (BEMAST) (50). Plasma samples archived by the BEMAST project were used to examine differences in the array of small molecule metabolites, using metabolomics, between green sea turtles (Chelonia mydas) with and without fibropapillomatosis (FP) (51). FP is a disease associated with a herpesvirus that causes the growth of debilitating tumors that can impede sight, foraging, and movement in threatened and endangered sea turtles. In the Hawaiian Islands, the disease prevalence peaked in the mid 1990’s (≈50 % at one site) and has been declining since (52). No significant differences were seen between the metabolome of turtles with or without tumors, but the methods developed from these preliminary samples (51) could be applied to a much larger sample size now available in the specimen bank. Currently, green sea turtle plasma samples archived by BEMAST from across the Pacific Ocean are being analyzed for gene expression through transcriptomics methods (53). These results will be compared to FP prevalence and POPs measured previously in these same turtles (44) to determine if this disease or these contaminants influence the expression of certain genes.

Analytical Methods Improve over Time Analytical methods are constantly evolving as critically reviewed by Muir and Sverko (54). Methods are changing so quickly that Analytical Chemistry now devotes the first issue of each year to new analytical methods and techniques. While not specific to the Marine ESB, a classic example of analytical changes over time is that involving PCB analysis. PCBs are highly bioaccumulative and frequently monitored in ESB samples. One of the major changes in PCB analysis 6 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

came as a result of the availability of individual PCB congeners (55) and the use of the congeners for developing methods using capillary column methods (e.g. Schulz et al. (56)). Once standards and methods became available, the newer congener analysis using capillary-column gas chromatography (GC) was soon favored over Aroclor standards for PCB analysis that relied on packed-column GC. However, values obtained from the newer PCB congener methods were less than half of those based on Aroclors derived using packed column GC (57). Using ESB samples, total PCB concentrations measured using older, non-capillary column GC have been adjusted to values generated using congener specific analysis (58). A set of common murre eggs archived in the ESB at the Swedish Museum of Natural History collected prior to 1988 that had total PCB values based on older methods were re-analyzed using congener-specific PCB methods. The observed relationship between total PCBs generated using both old and new methods allowed for the Swedish ESB to adjust older PCB data and construct a temporal record for total PCB concentrations in common murre eggs stretching back to the late 1960s. While the use of matrix-matched certified reference materials with each batch of samples does assist in accurately accounting for many analytical method changes, in this case, the ability to retrospectively compare values from different methods using banked samples was invaluable.

Properly Collected and Stored Specimen Bank Samples Allow Researchers To Retrospectively Obtain Baseline Values and Compare Species and Locations One of the first retrospective studies from the Marine ESB examined PBDEs, HBCDs and the naturally occurring methoxylated polybrominated diphenyl ethers (MeO-BDEs) in blubber from California sea lions that stranded between 1993 and 2003 (34). Additional retrospective studies on blubber from Alaskan beluga whales subsistently harvested between 1989 and 2006 examined PFAAs, PBDEs, HBCDs, as well as PCBs and organochlorine pesticides (21, 35). The most recent study also examined these compounds in blubber from subadult male northern fur seals that were harvested on St. Paul Island, Alaska between 1987 and 2007 with paired livers being analyzed for PFAAs and vitamins A and E (22). While no temporal trends were available, PCBs, organochlorine pesticides, PBDEs, and HBCDs baseline values were assigned for 16 species of cetaceans that stranded in the Main Hawaiian Islands from 1997 to 2011 (48). These studies demonstrated the usefulness of the Marine ESB to examine new compounds of interest as well as retrospective analysis with newer instrumentation. The STAMP component has analyzed samples in batches throughout the collections with SRM 1946 Lake Superior Fish Tissue and an in-house murre egg homogenate (59). Beginning with the 2001 collections, PBDEs were added to the list of routine analytes. PFAAs were retrospectively examined, but many were below the detection limit (25). Black-footed albatross (Phoebastria nigripes) and Laysan albatross (P. immutabilis) egg collections from Hawaii began in 2010 to supplement the Alaskan egg data. 7 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

The BEMAST project has collected blood and scutes from live captures of green sea turtles, hawksbill sea turtles (Eretmochelys imbricata), and leatherback sea turtles (Dermochelys coriacea). Fresh dead specimens of these species, as well as olive ridley sea turtles (Lepidochelys olivacea) and loggerhead sea turtles (Caretta caretta), are also sampled during necropsies for scutes, fat, muscle, bile, liver, blubber, GI tract and if available, follicles, shelled eggs, and fibropapilloma (FP) lesions, as well as skin from leatherbacks. Green and hawksbill sea turtle nests are also excavated after emergence and unhatched eggs collected (50). Several BEMAST samples have been analyzed for POPs and heavy metals (16, 44, 60, 61). In one study, POPs and other halogenated phenolic compounds were found to not be responsible for initiating the disease FP in Hawaiian green sea turtles (44). For the purposes of further discussing organic contaminant temporal trends in the North Pacific basin, selected data from the studies on blubber from California sea lions (only males) (34), northern fur seals (juvenile males) (22), belugas (adult males) (35), and eggs from murres (25, 36–40, 62) archived by the Marine ESB were examined. Temporal trends data previously published for the marine mammals may differ slightly from the results shown here due to the selection of only certain samples to limit confounding factors, such as gender and age, and the use of a wet mass basis rather than lipid normalized basis as previously published for some of the data. Lipid content was not available for some samples, necessitating the use of wet mass basis, and the use of lipid normalization may cause additional bias as well, especially when contaminant and lipid levels are not correlated (63). The recently collected albatross (25) and sea turtle (16, 44) samples were excluded as these have fewer years available for temporal trends and the sea turtle data is from plasma further confounding the comparisons. In addition, the 16 species of cetaceans stranded in the Hawaiian Islands (48) were excluded due to lack of temporal comparisons within one species, gender, and age class. However, the data from the albatross and male cetacean samples are shown in Table 1 for spatial comparisons. Summary statistics on ng/g wet mass basis followed methods recommended in Helsel (64). Briefly, non-parametric Kaplan-Meier estimates were used for data sets with >50% detection, the robust maximum likelihood or robust regression on order statistic estimates were used for data sets with 20-50% detection, and for data sets with 0.05). Beluga whales from the Chukchi Sea had a very large negative slope but this was also not significant along with the common murres from Bluff, Alaska. Results were similar for DDT and its metabolites. DDT was first synthesized in 1874 and used as an insecticide since 1939 (71). While production and use was limited by the Stockholm Convention beginning in 2004, DDT continues to be produced and used to treat malaria (72). For DDTs, the Bluff common murre eggs showed a significant (p < 0.05) decline, but those from St. Lawrence Island, Alaska were not significant. PCBs were produced from 1930 until 1993 for use as coolants and lubricants in electrical equipment (73). The sum of 18 major PCB congeners (out of a possible 209) showed significant (p < 0.05) negative declines for only common and thick-billed murres from St. George Island, Alaska and common murres from St. Lazaria Island, Alaska. Beluga whales from both Cook Inlet and the Chukchi Sea and common murre eggs from St. Lawrence, Island and Bluff showed positive, but non-significant (p > 0.05) trends. Due to species and tissue differences, geographical differences are difficult to examine. However, with the exception of PBDE 47 discussed above, these selected contaminants were highest in beluga whales from the Chukchi Sea and the cetaceans stranded in Hawaii compared to all other species and locations shown in Table 1. Beluga whales from Cook Inlet were similar in concentration to Northern fur seals from St. Paul Island, Alaska. A more thorough comparison of the beluga populations is available in Hoguet et al. (35) Among the murre eggs, St. Lazaria Island in the Southeast Gulf of Alaska had the highest levels of PCBs, DDTs, and PBDE 47. The PBDE 47 levels in St. Lazaria murres were similar to beluga whales and northern fur seals (Table 1). Chlordanes and α-HCH were similar among all locations of murre eggs. The Laysan albatross eggs from Hawaii had lower levels of α-HCH and PBDE 47 compared to the murre eggs, but higher levels of PCBs, DDTs, and chlordanes. The cetaceans stranded in Hawaii also had very low levels of α-HCH, reflecting the trend discussed by Li and Macdonald for α-HCH to dominate at higher latitudes (65). While PBDE 47 in the Hawaiian Laysan albatross eggs was generally below detection, the cetaceans stranded in Hawaii had higher levels than the Alaskan marine mammals. A manuscript detailing the temporal changes and geographical differences of contaminants in murre eggs is in preparation, but distinctive time point geographical differences are detailed in Vander Pol et al. (36–40, 62)

ESBs Are for the Common Good, But Researchers Need Easy Access To Sample Information and Previous Contaminant Results The Marine ESB samples are available to other researchers via tissue access policies. One project is federally mandated by United States Law and is available at: https://mmhsrp.nmfs.noaa.gov/tissbk. Other projects have policies established 12 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

within written protocols and procedures and are provided in publicly available reports (50, 74). These policies are similar to the data published by the Japanese Environmental Specimen Bank (75), Antarctic Environmental Specimen Bank (76), and German Environmental Specimen Bank (UPB) (77). A new data platform that is currently being developed is the Marine Sample Tracking and Analytical Reporting (STAR). Marine STAR will make it easier for researchers to not only determine which samples are available in the Marine ESB, but also view the results of previous chemical analysis with links to publications via a web-based interface. Currently only the UPB has interactive contaminant data available to the public. Similar resources applied to all specimen banks would greatly enhance our understanding of global trends regarding environmental contaminants in biotic matrices.

Acknowledgments The authors would like to thank all those who assisted in collections and processing for the Marine ESB samples as well as those who provided a critical review of this chapter.

Disclaimer Certain commercial equipment or instruments are identified in the paper to specify adequately the experimental procedures. Such identification does not imply recommendations or endorsement by the NIST nor does it imply that the equipment or instruments are the best available for the purpose.

References 1. 2.

3.

4.

5.

Loganathan, B. G.; Lam, P. K. S., Global Contamination Trends of Persistent Organic Chemicals; CRC Press: Boca Raton, FL, 2012; p xviii, 638 p. Muir, D.; Braune, B.; DeMarch, B.; Norstrom, R.; Wagemann, R.; Lockhart, L.; Hargrave, B.; Bright, D.; Addison, R.; Payne, J.; Reimer, K. Spatial and Temporal Trends and Effects of Contaminants in the Canadian Arctic Marine Ecosystem: A Review. Sci. Total Environ. 1999, 230, 83–144. Pugh, R. S.; Moors, A. J.; Rust, L. B.; Porter, B. J.; Becker, P. R. In Interdisciplinary Studies on Environmental Chemistry Chemistry Vol 4: Environmental Specimen Bank: Exploring Possibility of Setting-up ESBs in Developing Countries; Isobe, T.; Nomiyama, K.; Suramanian, A.; Tanabe, S., Eds. TerraPub: Tokyo, 2010; pp 33−41. Keller, J. M.; Kucklick, J. R.; Harms, C. A.; McClellan-Green, P. D. Organochlorine Contaminants in Sea Turtles: Correlations between Whole Blood and Fat. Environ. Toxicol. Chem. 2004, 23, 726–738. Yordy, J. E.; Wells, R. S.; Balmer, B. C.; Schwacke, L. H.; Rowles, T. K.; Kucklick, J. R. Partitioning of Persistent Organic Pollutants between Blubber 13

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

6.

7.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

8.

9.

10.

11.

12.

13.

14.

15. 16. 17.

18.

and Blood of Wild Bottlenose Dolphins: Implications for Biomonitoring and Health. Environ. Sci. Technol. 2010, 44, 4789–4795. Kucklick, J.; Schwacke, L.; Wells, R.; Hohn, A.; Guichard, A.; Yordy, J.; Hansen, L.; Zolman, E.; Wilson, R.; Litz, J.; Nowacek, D.; Rowles, T.; Pugh, R.; Balmer, B.; Sinclair, C.; Rosel, P. Bottlenose Dolphins as Indicators of Persistent Organic Pollutants in the Western North Atlantic Ocean and Northern Gulf of Mexico. Environ. Sci. Technol. 2011, 45, 4270–4277. Hollamby, S.; Afema-Azikuru, J.; Waigo, S.; Cameron, K.; Rae Gandolf, A.; Norris, A.; Sikarskie, J. G. Suggested Guidelines for Use of Avian Species as Biomonitors. Environ. Monit. Assess. 2006, 118, 13–20. OPSAR JAMP Guidelines for Monitoring Contaminants in Biota (Agreement 1999-02, Update 2010, Revision 2012). http://www.ospar.org/ documents?d=32414 (accessed July 15, 2016). Wise, S. A.; Koster, B. J. Considerations in the Design of an Environmental Specimen Bank: Experiences of the National Biomonitoring Specimen Bank Program. Environ. Health Perspect. 1995, 103, 61–67. York, G. W.; Porter, B. J.; Pugh, R. S.; Roseneau, D. G.; Simac, K.; Becker, P. R.; Thorsteinson, L. K.; Wise., S. A. Seabird Tissue Archival and Monitoring Project: Protocol for Collecting and Banking Seabird Eggs; NISTIR 6735; National Institute of Standards and Technology: Gaithersburg, MD, 2001, 2001. Pastor, D.; Jover, L.; Ruiz, X.; Albaigés, J. Monitoring Organochlorine Pollution in Audouin’s Gull Eggs: The Relevance of Sampling Procedures. Sci. Total Environ. 1995, 162, 215–223. Ylitalo, G. M.; Matkin, C. O.; Buzitis, J.; Krahn, M. M.; Jones, L. L.; Rowles, T.; Stein, J. E. Influence of Life-History Parameters on Organochlorine Concentrations in Free-Ranging Killer Whales (Orcinus orca) from Prince William Sound, AK. Sci. Total Environ. 2001, 281, 183–203. Yordy, J. E.; Wells, R. S.; Balmer, B. C.; Schwacke, L. H.; Rowles, T. K.; Kucklick, J. R. Life History As a Source of Variation for Persistent Organic Pollutant (POP) Patterns in a Community of Common Bottlenose Dolphins (Tursiops truncatus) Resident to Sarasota Bay, FL. Sci. Total Environ. 2010, 408, 2163–2172. Becker, P. R.; Wise, S. A. The U.S. National Biomonitoring Specimen Bank and the Marine Environmental Specimen Bank. J. Environ. Monit. 2006, 8, 795–799. Reiner, J. L.; Place, B. J. In Toxicological Effects of Perfluoroalkyl and Polyfluoroalkyl Substances; Springer, 2015; pp 127−150. Lynch, J. M.; et al. National Institute of Standards and Technology, unpublished. Binetti, R.; Costamagna, F. M.; Marcello, I. Exponential Growth of New Chemicals and Evolution of Information Relevant to Risk Control. Ann Ist Super Sanita 2008, 44, 13–5. Howard, P. H.; Muir, D. C. G. Identifying New Persistent and Bioaccumulative Organics Among Chemicals in Commerce. Environ. Sci. Technol. 2010, 44, 2277–2285. 14

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

19. Howard, P. H.; Muir, D. C. G. Identifying New Persistent and Bioaccumulative Organics Among Chemicals in Commerce II: Pharmaceuticals. Environ. Sci. Technol. 2011, 45, 6938–6946. 20. Lau, C. In Toxicological Effects of Perfluoroalkyl and Polyfluoroalkyl Substances; Springer: Cham, Switzerland, 2015; pp 1−21. 21. Reiner, J. L.; O’Connell, S. G.; Moors, A. J.; Kucklick, J. R.; Becker, P. R.; Keller, J. M. Spatial and Temporal Trends of Perfluorinated Compounds in Beluga Whales (Delphinapterus leucas) from Alaska. Environ. Sci. Technol. 2011, 45, 8129–8136. 22. Reiner, J. L.; Becker, P. R.; Gribble, M. O.; Lynch, J. M.; Moors, A. J.; Ness, J.; Peterson, D.; Pugh, R. S.; Ragland, T.; Rimmer, C.; Rhoderick, J.; Schantz, M. M.; Trevillian, J.; Kucklick, J. R. Organohalogen Contaminants and Vitamins in Northern Fur Seals (Callorhinus ursinus) Collected During Subsistence Hunts in Alaska. Arch. Environ. Contam. Toxicol. 2015, 70, 96–105. 23. Keller, J. M.; Ngai, L.; McNeill, J. B.; Wood, L. D.; Stewart, K. R.; O’Connell, S. G.; Kucklick, J. R. Perfluoroalkyl Contaminants in Plasma of Five Sea Turtle Species: Comparisons in Concentration and Potential Health Risks. Environ. Toxicol. Chem. 2012, 31, 1223–1230. 24. O’Connell, S. G.; Arendt, M.; Segars, A.; Kimmel, T.; Braun-McNeill, J.; Avens, L.; Schroeder, B.; Ngai, L.; Kucklick, J. R.; Keller, J. M. Temporal and Spatial Trends of Perfluorinated Compounds in Juvenile Loggerhead Sea Turtles (Caretta caretta) along the East Coast of the United States. Environ. Sci. Technol. 2010, 44, 5202–5209. 25. Vander Pol, S. S.; et al. National Institute of Standards and Technology, unpublished. 26. de Wit, C. A. An Overview of Brominated Flame Retardants in the Environment. Chemosphere 2002, 46, 583–624. 27. Covaci, A.; Voorspoels, S.; de Boer, J. Determination of Brominated Flame Retardants, With Emphasis on Polybrominated Diphenyl Ethers (PBDEs) in Environmental and Human Samples—A Review. Environ. Int. 2003, 29, 735–756. 28. Kodavanti, P. R. S.; Loganathan, B. G. In Biomarkers in Toxicology, Academic Press: Boston, 2014; pp 433−450. 29. Rahman, F.; Langford, K. H.; Scrimshaw, M. D.; Lester, J. N. Polybrominated Diphenyl Ether (PBDE) Flame Retardants. Sci. Total Environ. 2001, 275, 1–17. 30. Sjödin, A.; Patterson, D. G., Jr.; Bergman, Å. A Review on Human Exposure to Brominated Flame Retardants—Particularly Polybrominated Diphenyl Ethers. Environ. Int. 2003, 29, 829–839. 31. Kucklick, J. R.; Tuerk, K. J. S.; Vander Pol, S. S.; Schantz, M. M.; Wise, S. A. Polybrominated Diphenyl Ether Congeners and Toxaphene in Selected Marine Standard Reference Materials. Anal. Bioanal. Chem. 2004, 378, 1147–1151. 32. Peck, A. M.; Pugh, R. S.; Moors, A.; Ellisor, M. B.; Porter, B. J.; Becker, P. R.; Kucklick, J. R. Hexabromocyclododecane in White-Sided Dolphins: 15 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

33.

34.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

35.

36. 37.

38.

39.

40.

41.

42.

Temporal Trend and Stereoisomer Distribution in Tissues. Environ. Sci. Technol. 2008, 42, 2650–2655. Tuerk, K. J. S.; Kucklick, J. R.; Becker, P. R.; Stapleton, H. M.; Baker, J. E. Persistent Organic Pollutants in Two Dolphin Species with Focus on Toxaphene and Polybrominated Diphenyl Ethers. Environ. Sci. Technol. 2005, 39, 692–698. Stapleton, H. M.; Dodder, N. G.; Kucklick, J. R.; Reddy, C. M.; Schantz, M. M.; Becker, P. R.; Gulland, F.; Porter, B. J.; Wise, S. A. Determination of HBCD, PBDEs and MeO-BDEs in California Sea Lions (Zalophus californianus) Stranded between 1993 and 2003. Mar. Pollut. Bull. 2006, 52, 522–531. Hoguet, J.; Keller, J. M.; Reiner, J. L.; Kucklick, J. R.; Bryan, C. E.; Moors, A. J.; Pugh, R. S.; Becker, P. R. Spatial and Temporal Trends of Persistent Organic Pollutants and Mercury in Beluga Whales (Delphinapterus leucas) from Alaska. Sci. Total Environ. 2013, 449, 285–294. Vander Pol, S. S.; Becker, P. R. Monitoring Contaminants in Seabirds: The Importance of Specimen Banking. Mar. Ornithol. 2007, 35, 113–118. Vander Pol, S. S.; Becker, P. R.; Berail, S.; Day, R. D.; Donard, O. F. X.; Hobson, K. A.; Moors, A. J.; Pugh, R. S.; Rust, L. B.; Roseneau, D. G. Seabird Tissue Archival and Monitoring Project: Egg Collections and Analytical Results for 2006−2009; NISTIR 7872; National Institute of Standards and Technology: Gaithersburg, MD, 2012. Vander Pol, S. S.; Becker, P. R.; Day, R. D.; Ellisor, M. B.; Guichard, A.; Moors, A. J.; Point, D.; Pugh, R. S.; Roseneau, D. G. Seabird Tissue Archival and Monitoring Project: Egg Collections and Analytical Results for 2002–2005; NISTIR 7562; National Institute of Standards and Technology: Gaithersburg, MD, 2009. Vander Pol, S. S.; Becker, P. R.; Kucklick, J. R.; Pugh, R. S.; Roseneau, D. G.; Simac, K. S. Persistent Organic Pollutants in Alaskan Murre (Uria spp.) Eggs: Geographical, Species, and Temporal Comparisons. Environ. Sci. Technol. 2004, 38, 1305–1312. Vander Pol, S. S.; Christopher, S. J.; Roseneau, D. G.; Becker, P. R.; Day, R. D.; Kucklick, J. R.; Pugh, R. S.; Simac, K. S.; York, G. W. Seabird Tissue Archival and Monitoring Project: Egg Collections and Analytical Results for 1999−2002; NISTIR 7029; National Institute of Standards and Technology: Gaithersburg, MD, 2003. Vander Pol, S. S.; Becker, P. R.; Ellisor, M. B.; Moors, A. J.; Pugh, R. S.; Roseneau, D. G. Monitoring Organic Contaminants in Eggs of Glaucous and Glaucous-Winged Gulls (Larus hyperboreus and Larus glaucescens) from Alaska. Environ. Pollut. 2009, 157, 755–762. Alava, J. J.; Keller, J. M.; Wyneken, J.; Crowder, L.; Scott, G.; Kucklick, J. R. Geographical Variation of Persistent Organic Pollutants in Eggs of Threatened Loggerhead Sea Turtles (Caretta caretta) from Southeastern United States. Environ. Toxicol. Chem. 2011, 30, 1677–1688.

16 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

43. Clukey, K. E. Assessment of Plastic Ingestion and Persistent Organic Pollutants in Sea Turtles across the Pacific Ocean. Master’s Thesis, University of Hawaii, Honolulu, HI, May 2016. 44. Keller, J. M.; Balazs, G. H.; Nilsen, F.; Rice, M.; Work, T. M.; Jensen, B. A. Investigating the Potential Role of Persistent Organic Pollutants in Hawaiian Green Sea Turtle Fibropapillomatosis. Environ. Sci. Technol. 2014, 48, 7807–7816. 45. Ragland, J. M.; Arendt, M. D.; Kucklick, J. R.; Keller, J. M. Persistent Organic Pollutants in Blood Plasma of Satellite-Tracked Adult Male Loggerhead Sea Turtles (Caretta caretta). Environ. Toxicol. Chem. 2011, 30, 1549–1556. 46. Stewart, K. R.; Keller, J. M.; Templeton, R.; Kucklick, J. R.; Johnson, C. Monitoring Persistent Organic Pollutants in Leatherback Turtles (Dermochelys coriacea) Confirms Maternal Transfer. Mar. Pollut. Bull. 2011, 62, 1396–1409. 47. Swarthout, R. F.; Keller, J. M.; Peden-Adams, M.; Landry, A. M.; Fair, P. A.; Kucklick, J. R. Organohalogen Contaminants in Blood of Kemp’s Ridley (Lepidochelys kempii) and Green Sea Turtles (Chelonia mydas) from the Gulf of Mexico. Chemosphere 2010, 78, 731–741. 48. Bachman, M. J.; Keller, J. M.; West, K. L.; Jensen, B. A. Persistent Organic Pollutant Concentrations in Blubber of 16 Species of Cetaceans Stranded in the Pacific Islands from 1997 through 2011. Sci. Total Environ. 2014, 488–489, 115–123. 49. Marvin, C. H.; Sverko, E.; Tomy, G. T. In Dioxin and Related Compounds: Special Volume in Honor of Otto Hutzinger; Alaee, M., Ed.; Springer International Publishing: Cham, 2016; pp 411−456. 50. Keller, J. M.; Pugh, R.; Becker, P. R. Biological and Environmental Monitoring and Archival of Sea Turtle Tissues (BEMAST): Rationale, Protocols, And Initial Collections of Banked Sea Turtle Tissues; NISTIR 7996; National Institute of Standards and Technology: Gaithersburg, MD, 2014. 51. Schock, T. B.; Keller, J. M.; Rice, M.; Balazs, G. H.; Bearden, D. W. Metabotyping of a Protected Non-Model Organism, Green Sea Turtle (Chelonia mydas), Using 1H NMR Spectroscopy and Optimized Plasma Methods for Metabolomics. Curr. Metabolomics 2013, 1, 279–290. 52. Chaloupka, M.; Balazs, G. H.; Work, T. M. Rise and Fall over 26 Years of a Marine Epizootic in Hawaiian Green Sea Turtles. J. Wildl. Dis. 2009, 45, 1138–1142. 53. Komoroske, L. M.; et al. National Oceanic and Atmospheric Administration Southwest Fisheries Sceince Center, unpublished. 54. Muir, D.; Sverko, E. Analytical Methods for PCBs and Organochlorine Pesticides in Environmental Monitoring and Surveillance: A Critical Appraisal. Anal. Bioanal. Chem. 2006, 386, 769–789. 55. Mullins, M. D.; Pochini, C. M.; McCrindle, S.; Romkes, M.; Safe, S. H.; Safe, L. M. High-Resolution PCB Analysis: Synthesis and Chromatographic Properties of All 209 PCB Congeners. Environ. Sci. Technol. 1984, 18, 468–476. 17 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

56. Schulz, D. E.; Petrick, G.; Duinker, J. C. Complete Characterization of Polychlorinated Biphenyl Congeners in Commercial Aroclor and Clophen Mixtures by Multidimensional Gas Chromatography-Electron Capture Detection. Environ. Sci. Technol. 1989, 23, 852–859. 57. Turle, R.; Norstrom, R. J.; Collins, B. Comparison of PCB Quantitation Methods: Re-Analysis of Archived Specimens of Herring Gull Eggs from the Great Lakes. Chemosphere 1991, 22, 201–213. 58. Odsjö, T.; Bignert, A.; Olsson, M.; Asplund, L.; Eriksson, U.; Häggberg, L.; Litzén, K.; de Wit, C.; Rappe, C.; Åslund, K. The Swedish Environmental Specimen Bank — Application in Trend Monitoring of Mercury and Some Organohalogenated Compounds. Chemosphere 1997, 34, 2059–2066. 59. Vander Pol, S. S.; Ellisor, M. B.; Pugh, R. S.; Becker, P. R.; Poster, D. L.; Schantz, M. M.; Leigh, S. D.; Wakeford, B. J.; Roseneau, D. G.; Simac, K. S. Development of a Murre (Uria spp.) Egg Control Material. Anal. Bioanal. Chem. 2007, 387, 2357–2363. 60. Bryan, C. E.; et al., National Institute of Standards and Technology, unpublished. 61. Day, R. D.; et al., National Institute of Standards and Technology, unpublished. 62. Vander Pol, S. S.; Hobson, K. A.; Becker, P. R.; Day, R. D.; Ellisor, M. B.; Pugh, R. S.; Roseneau, D. G. Geographic Differences in Organic Contaminants and Stable Isotopes (d13C, d15N) in Thick-Billed Murre (Uria lomvia) Eggs from Alaska. J. Environ. Monit. 2011, 13, 699–705. 63. Hebert, C. E.; Keenleyside, K. A. To Normalize or Not to Normalize? Fat Is the Question. Environ. Toxicol. Chem. 1995, 14, 801–807. 64. Helsel, D. R. In Statistics for Censored Environmental Data Using Minitab® and R; John Wiley & Sons, Inc.: Hoboken, NJ, 2011; pp 268−296. 65. Li, Y. F.; Macdonald, R. W. Sources and Pathways of Selected Organochlorine Pesticides to the Arctic and the Effect of Pathway Divergence on HCH Trends in Biota: A Review. Sci. Total Environ. 2005, 342, 87–106. 66. Kannan, K.; Moon, H.-B.; Yun, S. H.; Agusa, T.; Thomas, N. J.; Tanabe, S. Chlorinated, Brominated, and Perfluorinated Compounds, Polycyclic Aromatic Hydrocarbons and Trace Elements in Livers of Sea Otters from California, Washington, and Alaska (USA), and Kamchatka (Russia). J. Environ. Monit. 2008, 10, 552–558. 67. McKinney, M. A.; Cesh, L. S.; Elliott, J. E.; Williams, T. D.; Garcelon, D. K.; Letcher, R. J. Brominated Flame Retardants and Halogenated Phenolic Compounds in North American West Coast Bald Eaglet (Haliaeetus leucocephalus) Plasma. Environ. Sci. Technol. 2006, 40, 6275–6281. 68. Daley, R. E.; Blum, A. In Understanding the Role of California Flammability Standard TB117 in the History of Flame Retardant Biomonitoring and Health Findings; Proceedings of the 12th International Conference on Envrionmental Science and Technology, Rhodes, Greece, Rhodes, Greece, 8−10 September 2011; pp B-174−B-180.

18 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch001

69. Li, Y. F.; Bidleman, T. F.; Barrie, L. A.; McConnell, L. L. Global Hexachlorocyclohexane Use Trends and Their Impact on the Arctic Atmospheric Environment. Geophys. Res. Lett. 1998, 25, 39–41. 70. Dearth, M. A.; Hites, R. A. Complete Analysis of Technical Chlordane Using Negative Ionization Mass Spectrometry. Environ. Sci. Technol. 1991, 25, 245–254. 71. Environmental Health Criteria: DDT and Its Derivatives; World Health Organization, 1979. 72. van den Berg, H. Global Status of DDT and Its Alternatives for Use in Vector Control to Prevent Disease. Environ. Health Perspect. 2009, 117, 1656–1663. 73. Breivik, K.; Sweetman, A.; Pacyna, J. M.; Jones, K. C. Towards a Global Historical Emission Inventory for Selected PCB Congeners — a Mass Balance Approach: 1. Global Production and Consumption. Sci. Total Environ. 2002, 290, 181–198. 74. Rust, L.; Pugh, R.; Moors, A.; Vander Pol, S.; Roseneau, D.; Becker, P. Seabird Tissue Archival and Monitoring Project (STAMP): Project Overview and Updated Protocols for Collecting and Banking Seabird Eggs; NISTIR 7678; National Institute of Standards and Technology: Gaithersburg, MD, 2010. 75. Center for Marine Environmental Studies Ehime University Environmental Specimen Bank (es-BANK). http://esbank-ehime.com/dnn/Home/ RequestForm/tabid/106/Default.aspx (accessed July 15, 2016). 76. Banca Campioni Ambientali Antartici (Antarctic Environmental Specimen Bank) Home Page. http://www.bcaa.unige.it/introframe.asp?lang=english (accessed July 15, 2016). 77. Umwelt Probenbank des Bundes (German Environmental Specimen Bank) Data Search. https://www.umweltprobenbank.de/en/documents/10027 (accessed July 15, 2016).

19 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

Chapter 2

Brominated Flame Retardants: Spatial and Temporal Trends in the Environment and Biota from the Pacific Basin Countries Prasada Rao S. Kodavanti*,1 and Bommanna G. Loganathan2 1Toxicity

Assessment Division, NHEERL/ORD, U.S. Environmental Protection Agency, Research Triangle Park, North Carolina 27711, United States 2Department of Chemistry and Watershed Studies Institute, Murray State University, 1201 Jesse D. Jones Hall, Murray, Kentucky 42071, United States *E-mail: [email protected]

Brominated flame retardants (BFRs) are used as additive or reactive components in a variety of polymers including high-impact polystyrene and epoxy resins, commercial products such as computers, electronics and electrical equipment, thermal insulation, textiles and furniture foam. There were over 75 different BFRs in the market; some of them were restricted/banned from production and use due to their environmental persistence, bioaccumulation and toxic effects on organisms. Of the many BFRs still on the market, brominated bisphenols, decabrominated diphenyl ethers, and cyclododecanes are three major classes which represent the highest production volumes. Recent studies have revealed that environmental contamination and toxic health effects by high production volume BFRs continues to be of concern. Trend monitoring studies are useful in understanding the historical perspectives, current status and also help to predict future trends of environmental contamination by these compounds. This chapter deals with the environmental contamination status and temporal trends of polybrominated diphenylethers in a © 2016 American Chemical Society Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

variety of environmental and biological matrices, including soil, sediment, wildlife, marine and terrestrial mammals from Pacific Basin countries.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

Introduction Persistent Organic Chemicals (POCs) are synthetic chemicals, either intentionally or unintentionally produced/released into the environment (1). Some of the POCs are pesticides while others are industrial products or unintended by-products resulting from industrial processes (combustion) or from consumer products (Figure 1). POCs remain in the environment for extended periods of time and several factors can contribute to a compound’s persistence in the ecosystem. POCs resist degradation through natural processes and can become concentrated in sediment, water or air. These compounds can be volatile (i.e. can vaporize in the air) or travel by water currents through the process of evaporation and re-deposition. These traits allow POCs to be transported over long distances, far from the original source of contamination. One of the main characteristics of POCs is that they build up in the fatty tissue of living organisms, with serious consequences for humans and wildlife. This affinity for lipid rich tissue allows these compounds to accumulate, persist due to their resistance to biological degradation, and bioconcentrate in biological tissues and biomagnify in the foodchain. Consequently, even though the level of exposure may be limited, POCs can eventually reach toxicologically relevant concentrations. Because of their ability to accumulate inside an organism, to be transported long-range, to persist in the environment, and to be toxic, POCs are considered a global threat. POCs are also known as persistent organic pollutants (POPs). Figure 1 shows the chemical structures of some of the legacy, unintentionally produced POCs (polychlorinated dibenzo- dioxins and furans) and emerging compounds of concern. Brominated flame retardants (BFRs) belong to a class of compounds known as organohalogens most of which are highly persistent in the environment (1–3). BFRs are currently the largest marketed flame retardant group due to their high performance efficiency and low cost. In the commercial market, more than 75 different BFRs are recognized. Some BFRs, such as the polybrominated biphenyls (PBBs), were removed from the market in the early 1970s after an incidental poisoning resulted in the loss of livestock due to the ingestion of PBB-contaminated animal feed, which demonstrated the toxicity of this BFR class (4). Tris (2,3-dibromopropyl) phosphate, commonly known as “Tris”, is another BFR that was removed from children’s clothing, including pajamas, due to its mutagenic and nephrotoxic effects (5). Of the BFRs still on the market, brominated bisphenols, decabrominated diphenyl ethers, and cyclododecanes are three major classes which represent the highest production volumes. These BFRs are used as additive or reactive components in a variety of polymers such as high-impact polystyrene and epoxy resins, which are then used in commercial products such as computers, electronics and electrical equipment, thermal insulation, textiles and furniture foam. 22 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

Figure 1. Chemical structures of some of the legacy, unintentionally produced POCs and emerging compounds of concern. Polybrominated diphenyl ethers (PBDEs) constitute an important group of flame retardants (Figure 1). PBDEs were added to consumer products to prevent them from catching fire or delay the ignition process if exposed to flame or heat. PBDEs are added to plastics, upholstery, fabrics and foams and are in common products such as computers, television sets, mobile phones, furniture, and carpet pads. Nearly 90% of electrical and electronic appliances contain PBDEs and the Bromine Science and Environmental Forum (BSEF) claims that adding flame retardants gives 15 times greater escape time in case of a fire (6). Although PBDEs are ubiquitous, they are primarily considered as indoor pollutants based on human exposure scenarios. They leach into the environment when household wastes decompose in landfills or are incompletely incinerated. Human health concerns stem from the fact that PBDEs are persistent, bioaccumulative and structurally related to PCBs (Figure 1). PBDE concentrations are rapidly increasing in the global environment and in human blood, breast milk, liver, as well as other fatty tissues. However, in defined areas like the European Union, PBDEs are leveling off or declining due to regulations on these compounds (7). Although these 23 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

chemicals are ubiquitous in the environment and bioaccumulate in wildlife and humans, information on their potential toxic effects are now accruing (8–12). Tetrabromobisphenol A (TBBPA; Figure 1) is a reactive BFR found in printed circuit boards and is added to several types of polymers. TBBPA is highly lipophilic (log Kow [octanol-water partition coefficient] = 4.5) and has low water solubility (0.72 mg/ml). TBBPA has been measured in the air (13), soil and sediment (14), but is generally not found in water samples. Unlike most of the PBDEs, TBBPA has a relatively short elimination half-life of about 2-4 days in blood and 3 weeks in adipose tissue of humans, and about 0.5 days in serum and 3 days in adipose tissue of adult rats (15). Hexabromocyclododecane (HBCD; Figure 1) is a non-aromatic brominated cyclic alkane, mainly used as an additive flame retardant in thermoplastic polymers with final applications in styrene resins (16). Like other BFRs, HBCD is highly lipophilic, with a log Kow of 5.6 and has low water solubility (0.0034 mg/L) (17). Studies have shown that HBCD is highly persistent, with a half-life of 3 days in air and 2,025 days in water (18), and is bioaccumulative with a bioconcentration factor of approximately 18,100 in fathead minnows (19). Although we discuss some aspects of new/emerging BFRs in general, the contamination status and temporal trends of polybrominated diphenyl ethers (PBDEs) are discussed at length in this chapter.

Sources and Environmental Contamination of BFRs Industrial scale production of PBDEs began in 1976 (20). Commercial formulations of PBDEs include, penta-, octa-, and deca-BDEs and their bromine content was about 71%, 79% and 83%, respectively. In 2004, severe restrictions were placed on the production of PBDEs. Prior to 2004, about 95% of penta-BDE was used as an additive flame retardant in polyurethane foam materials used in seat cushions, bedding mattresses, furniture etc. Octa-BDE was used as an additive flame retardant in plastics that are used for the manufacture of office equipment and computer casings. Deca-BDE was used in a variety of plastics and polystyrene used in the manufacture of televisions, mobile phones, audio-video equipment, and several other plastic materials. Since PBDEs are used as additive flame retardants and are not chemically bound to the materials, PBDEs easily escape the matrix via volatilization into the air. Environmental contamination can occur during the disposal of electronic waste via leaching and volatilization. E-waste recycling locations in Asian countries have been identified as major sources of PBDEs (21). Lower brominated PBDEs (mono-, to hexa- bromine substituted) that are present in both vapor and particulate matter are transported long distances (long-range transport) from the source. In the United States, waste water treatment plants (WWTP) and landfills are considered point sources of PBDEs entering the environment (22). High concentrations of PBDEs found in WWTP effluents and sludge may contaminate agricultural land as they are used as fertilizer. Over half of the sewage sludge produced annually in the USA is applied to land as fertilizer (23). Thus, application of sewage sludge may represent a source of exposure to humans and wildlife through direct contact or uptake by plants. Due to the hydrophobic nature, PBDEs tend to adsorb strongly 24 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

to sediments, soils and suspended materials in water, thus facilitating their transfer to aquatic biota. A survey of US foods showed that PBDE levels were highest in fish (median: 1,725 pg/g), followed by meat (283 pg/g), and lowest in dairy products (31.5 pg/g) (24). Significant levels of PBDEs may be found in outdoor air, even at rural locations. PBDE concentrations in indoor measures were 15–20 times higher than in outdoor air (25). PBDEs also enter coastal waters through industrial and municipal waste water outfalls, landfill leachate and atmospheric deposition from various sources (26, 27). TBBPA, a compound used for many years in plastics and known to have numerous endocrine disrupting effects in humans, rodent research models, and wildlife. TBBPA is a high volume production flame retardant used in electrical equipment, plastics, and home furnishings (sofas, chairs, carpeting), and is currently the most widely used BFR in the world (28). It is typically detected in house dust (29) and inhalation is thought to be a major route of exposure to humans. People working as computer technicians, in electronics dismantling, and smelting have been reported to have elevated blood TBBPA levels compared to measures reported for the general population (30). In fact, the European Food Safety Authority (EFSA) tested over 650 foodstuffs for TBBPA and found it to be non-quantifiable, even in fish and related products (31). However, others have measured TBBPA in fish and drinking water, human serum and breast milk (32, 33), and its disposition in the rat has been characterized (34). TBBPA is not currently on the list of chemicals for biomonitoring in the U.S. Hexabromocyclododecane (HBCD) is a brominated flame retardant (Figure 1). It is used in polystyrene foam thermal insulation in buildings. Other uses include automobile interior textiles, car cushions, upholstered furniture and insulation blocks in trucks, packaging materials, and housing for electric and electronic equipment. HBCD is produced in China, Europe, Japan, and the USA. The known current annual production is approximately 28,000 tons per year (35). HBCD’s toxicity and its harm to the environment are currently under investigation. HBCD has been detected in environmental samples such as birds, mammals, fish and other aquatic organisms as well as soil and sediment (36). HBCD has been detected in workplace air samples at levels up to 1,400 μg/kg in dust (37). Diet is considered an important source for HBCD exposure (38), especially for humans consuming large quantities of fish, which reportedly contains relatively high HBCD levels (1,110 ng/g lipid) (39, 40). In addition to diet, house dust is probably another important source of human exposure, since dust consists of high levels of HBCD (41).

Status and Trends of PBDEs in Pacific Basin Countries Sediments Sediments serve as sources and sinks for a variety of POPs, including PBDEs. PBDEs have been detected in the sediments of Great Lakes (42–45) and in coastal marine sediments of North America (46). In the Great Lakes, total PBDE concentrations in surface sediments ranged from 10 to 236 ng/g dry wt., with PBDE 209 contributing to major proportions (from 8.9 to 230 ng/g 25 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

(47). The National Oceanic and Atmospheric Administration (NOAA)’s Mussel Watch Project collected surface sediments throughout the U.S. coast (46) and showed that high levels of PBDEs were found in densely populated and industrial areas, such as Marina del Rey in California (88 ng/g dry wt). Relatively high PBDE concentrations have been reported in sediments collected from urban and industrial areas in Australia (4.71 ± 12.6 ng/g d.wt.) (48), Tokyo Bay, Japan, and in the Pearl River Delta, China (Figure 2a, 2b) (49, 50).

Figure 2a. Temporal trends of PBDEs in sediment cores from USA and Canada. Data: Lake Ontario (44) (ng/g dry wt. Data estimated from figure). Lake Michigan and Lake Erie (45) (ng/g wet wt. Data estimated from figure).

26 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

Sediment cores have been used to reconstruct the history of contamination by persistent organic pollutants (51). Figure 2a shows PBDEs concentrations determined in sediment cores collected from the Great Lakes and provided information on the doubling time, inventory, surface flux, and loading rates of PBDEs in sediment (42–45, 47). Qiu et al. described temporal trends of PBDE-209 and tPBDE (PBDE3-7 , i.e. sum of PBDE congeners with three to seven bromines: BDE-28, -47, -49, -99, -100, -116, -153, -154, -181, and 183 which mainly come from the penta- and octa- BDE products) flame retardants in a sediment core from Lake Ontario (44). Zhu and Hites traced the temporal trends of BFRs in sediment cores from Lakes Michigan and Erie (45). They showed temporal trends of PBDE-209 and other PBDEs (tPBDE: sum of BDE-17, -28, -47, -49, -66, -71, -85, -99, -100, -138,- 153, -154, -183, -190, -206) in sediment cores (Figure 2a). Hites reported that total concentrations of PBDEs (excluding PBDE-209) and PBDE-209 increased annually since the 1980s with the doubling time of PBDE-209 in sediments from Lakes Michigan, Huron, Erie, and Ontario of 19, 10, 10, and 13 years, respectively (47). Drage et al. reported a large increase in all PBDEs between 1980 and 2014 in sediment cores collected from the Sydney estuary in Australia (Figure 2b) (52). This trend was particularly prominent for PBDE-209, which was found in surface sediment at an average concentration of 42 ng/g dry wt. (21-65 ng g. d.wt) (52). tPBDE trend include sum of six BDE congeners (BDE-47, -99, -100, -153, -154, -and 183) (52). Sediment cores collected from Pacific basin countries showed that PBDE levels have been increasing annually, but at different rates (i.e., doubling times) (Figure 2b). PBDE-209, as the major component (>96%) of the deca-BDE technical mixture (49, 53), was also shown to increase in sediment cores in these countries. As shown in Figure 2b, the highest concentrations of PBDE-209 were found in sediment samples collected from Asia. Sediment cores collected from the Pearl River Delta in China (50) showed that PBDE-209 concentrations remained similar until 1990, and then increased notably, with a doubling time of 3–6 years, reflecting the high market demand for deca-BDE mixture after 1990 in China. The elevated concentrations of PBDE-209 in sediment suggests its high production and usage, and high Kow value (octanol-water partition coefficients, log Kow ~10), resulting in preferential partitioning to the sinking sediment particles. In addition, PBDE-47, -99, -153, and -154 have also been frequently reported in sediments.

27 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

Figure 2b. Temporal trends of PBDEs in sediment cores from the Asia/Pacific region. Data: Tokyo Bay (49), Pearl River estuary, South China (50) and Sydney harbor, Australia (52) (ng/g dry wt. data estimated from figure).

Soil and House Dust In the United States, concentrations of PBDE in floodplain soils from Michigan ranged from 0.02 to 55.1 ng/g d.wt (54). In China, a low PBDE concentration (0.01 ng/g dry wt) was detected in surface soils from the Tibetan Plateau (55) and high concentrations (~2700 ng/g) were found in soils collected near E-waste dismantling sites in Guangdong (56). Since PBDEs are added to plastics, upholstery, fabrics and foams and in common products such as computers, television sets, mobile phones, furniture, and carpet pads etc., they are primarily considered as indoor pollutants based on human exposure scenarios. Concentrations of PBDEs in house dust are much higher than in soil or sediment. House dust samples from North America had the highest PBDE levels (thousands of nanogram per gram), followed by dust samples from Eastern Asia and Australia (Figure 3). Relatively high PBDE levels were reported in dust collected from 28 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

houses and garages (57), television sets (300 μg/g and 72 μg/g) (58, 59), houses with several computers (6 μg/g) (59), and an E-waste recycling area (60) (Figure 3) (61–67).

Figure 3. PBDE concentrations (ng/g dry weight) in house dust and outdoor dust from several countries in the Pacific Basin. Median values (mean value was used when median value was not available) from References: Canada (61, 62); China (59); Japan (63); Singapore (64); Australia (65, 66); the United States (57, 62, 65–67). (Adapted/Reproduced with permission from Ref # (3), 2012, the Taylor and Francis Group LLC Books).

Wild Life Species Mussels Mussels and oysters are considered sentinel organisms and are widely used to assess spatial distribution and temporal trends of PBDEs in coastal environments in the United States and Asia (46, 68–70). The NOAA’s Mussel Watch Program reported a concentration range of PBDEs in mussels and oysters, from 1 to 270 ng/g lipid wt (46). Mean PBDE concentrations of mussels collected from San Francisco Bay were 2380 ng/g lipid wt (71). deBruyn et al. (2009) reported that concentrations of PBDEs in mussels collected near a municipal outfall in British Columbia were higher than in a reference area, with median values of 1000 and Philippines > India.

Temporal Trends of PBDEs in Marine Mammals Trend monitoring data are useful to understand the fate of PBDEs in open ocean ecosytems, a remote area from where these compounds are produced and used. Several studies have reported temporal trends of PBDEs in marine mammals (Table 2). No clear trend was found for PBDE concentrations in California sea lions collected during 1993–2006 from the California coast (121, 124), in sea otter livers collected during 1992–2002 from the California coast (125), in harbor seals collected during 1991–2005 from California (121). She et al. observed an increasing trend of PBDEs in harbor seal blubber from 1989 to 1998 with concentrations ranging from 0.09-8.3 µg/g lipid wt (126). Total PBDE concentrations increased by 10-fold in ringed seals collected during 1981–2000 from the Canadian Arctic (127) and in bottlenose dolphin collected during 1993–2004 from Florida coast, United States (79). PBDE concentrations in Guiana dolphin collected during 1993–2004 from Southeast Brazil showed increasing concentrations with time (130). 34 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

35

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

Table 2. Temporal trends of PBDEs in marine mammals from Pacific Basin

*

Location

Marine mammal

ΣPBDE Range (µg/g. lipid wt.)

Trend

Ref.

Florida coast, USA

Bottlenose dolphin

0.03-4.5

Increased with doubling time of 3-4 years between 1993-2004

(79)

California, USA

California Sea lion

0.45-4.74*

No clear trend, 1993-2003

(124)

California, USA

California, Sea lion

0.04-33.7*

No clear trend, 1994-2006

(121)

San Francisco Bay, USA

Harbor seal

0.09-8.3

Increased, 1989-1998

(126)

Arctic Canada

Ringed seal (male)

0.5-4.6**

Increased, 1981-2000

(127)

Southeast Brazil

Guiana dolphin

0.01-1.6

Increased, 1994-2006

(130)

Hong Kong, China

Indo-Pacific dolphin

0.10-51

No trend, 1997-2008

(128)

Taiji, Japan

Striped dolphin

0.01-0.09

Increased, 1978-2003

(131)

Pacific coast, Japan

Melon-headed whales

0.02-0.5

Increased, 1982-2006

(132)

South China sea

Finless porpoise

0.08-1.0

Increased, 1990-2001

(133)

Pacific coast, Japan

Northern fur seal

0.0003-0.1

Reached peak (1991-94) and decreased 50% (1998)

(120)

µg/g wet wt.

**

ng/g lipid wt.

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

A few studies showed decreasing PBDE concentrations in marine mammals after 2000. PBDE concentrations in archived blubber tissues of Northern fur seals from Sanriku, Japan, increased during 1972–1994 sampling (about 150 times in 1994 to that in 1972), and then decreased by 50% during 1997–1998 sampling (120). The available data indicated that PBDE concentrations in marine mammals increased from the 1970s to the mid-1990s in the Pacific Basin. A decreasing trend of PBDEs in marine mammals from the coast of Japan after the late 1990s may be due to the earlier restriction on the usage of penta- and octa-PBDE formulations. Further studies are needed to evaluate the trends of PBDEs in North America after restrictions have been imposed on the production of PBDEs since the mid-2000s. Moreover, time-trend studies are needed in Asia (especially China and India), as the economies of these countries are increasing considerably and there has been a heavy demand for PBDEs in Asia in recent years. As stated in earlier sections, PBDEs are used in the manufacture of electronic equipment and therefore, electronic waste (e-waste) disposal has led to serious environmental problems by PBDEs in developing countries. It was estimated that 50-80% of the e-waste generated in the United States are exported to Asia, mainly China (133). and 60-75% of the e-waste generated in EU is exported to Asia and Africa. Significant quantities of e-waste are also exported to India, Malaysia, Pakistan, Philippines, and Vietnam etc. Therefore, it is expected that the environmental levels of PBDEs and other flame retardants may continue to increase in future in countries in these Asia/Pacific region. Terrestrial Mammals Like organochlorines, PBDEs are lipophilic and hydrophobic compounds and readily bioaccumulate into aquatic and terrestrial organisms. PBDEs were detected in human specimens such as serum, adipose tissue and breast milk. Studies have shown that PBDE levels in human samples from North America are much higher (10-100 times) than the levels reported for Asia (135, 136). Hites performed meta-analysis of the human biomonitoring data for Asia and America and showed median PBDE levels of 3.5, 40 ng/g lipid wt. respectively (135). Temporal trend studies of PBDEs in human tissue samples from the USA and Asia are limited. PBDE concentrations in breast milk collected from Osaka, Japan during 1973-2000 (137) reached their peak in 1998 (2.3 ng/g lipid wt.) and then stabilized. However, human adipose tissue collected between 1970 and 2000 from Tokyo showed a 40-fold increase in concentration (0.03-1.3 ng/g lipid wt.) during that period (138). In contrast, no temporal variation was observed in breast milk samples from Brisbane, Australia. Average PBDE concentrations in breast milk collected in 2003 and2007 were 10.2 and 10 ng/g lipid wt. respectively (139). Dye et al. determined PBDEs in the serum of domestic cats (140). PBDEs were found in all cats, with mean PBDE concentrations of 4.3 and 10.5 ng/mL for young and old cats, respectively. PBDEs-47, -99, -207, and -209 were the dominant congeners. Liang et al. found very high concentrations of PBDEs in various tissues of foraging hens from an e-waste recycling area in South China (141). The highest PBDE concentrations were found in muscle (18,000 ng/g lipid wt), followed by fat, intestine, heart, liver, oviduct, gizzard, blood, skin, and ovum 36 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

(125 ng/g lipid wt). PBDE-209 was found in all samples (33–18,000 ng/g lipid wt) and was the dominant congener. PBDEs were also found in captive giant and red pandas from China (142). Total PBDE concentrations in panda ranged from 16.4 to 2160 ng/g lipid wt. PBDE-209 was the most abundant congener, followed by PBDEs-206, -207, -203, -47, and -153 (142).

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

Summary and Conclusions PBDEs are widespread and are detectable in various environmental media (air, water and food) and biota (aquatic and terrestrial animals including human blood, breast-milk, and fetuses) of Pacific Basin countries. As flame retardants, PBDEs were thought to save lives by reducing fire. However, the increasing levels of PBDEs in the environment associated with its potential toxicity to biota present an emerging risk for environment and human health. In May 2009, the commercial penta- and octa-PBDE mixtures have been added to the list of POPs by the United Nations Stockholm Convention. Commercial penta- and octa-PBDE mixtures were banned in the Europe and the United States. This is reflected in some samples, with concentrations being stable or in slow decline. However, e-waste containing PBDEs and other flame retardants disposal has led to serious emerging new environmental problems in the Asia/Pacific countries. The global release of PBDEs through e-waste disposal is estimated to be at least 20,000 metric tons per year with China accounting for 14,000 metric tons annually143. Global scale regulations, modernization of e-waste recycling processes and strategies for management of e-waste are needed to control the emission and to protect the environment, wildlife and human health.

Acknowledgments Authors thank Dr. Kevin Miller of Murray State University, Murray, KY and Dr. Riyaz Basha of University of North Texas Health Sciences Center, Fort Worth, TX for their excellent comments on an earlier version of this chapter. Authors also thank Mr. John Havel for excellent graphic assistance. The contents of this article has been reviewed by the National Health and Environmental Effects Research Laboratory of the US Environmental Protection Agency and approved for publication. Approval does not signify that the contents necessarily reflect the views and policies of the Agency nor does mention of trade names or commercial products constitute endorsement or recommendation for use.

References 1.

2.

Loganathan, B. G. In Global Contamination Trends of Persistent Organic Chemicals; Loganathan, B. G., Lam, P. K. S., Eds.; CRC Press, Taylor and Francis Group: Boca Raton, FL, 2012; pp 3−31. Kodavanti, P. R. S.; Loganathan B. G. PCBs, PBBs, and brominated flame retardants. In Biomarkers in Toxicology; Gupta, R. C., Ed.; Oxford Academic Press: San Diego, CA; 2014; pp 433−450. 37

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

3.

4. 5.

6.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

7.

8.

9.

10.

11.

12.

13.

14.

15.

Guo, Y.; Shaw, S. D.; Kannan, K. In Global Contamination Trends of Persistent Organic Chemicals; Loganathan, B. G., Lam, P. K. S., Eds.; CRC Press, Taylor and Francis Group: Boca Raton, FL, 2012; pp 32−56. Fries, G. F. Bioavailability of soil-borne polybrominated biphenyls ingested by farm animals. J. Toxicol. Environ. Health 1985, 16, 565–579. Soderlund, E.; Dybing, E.; Nelson, S. D. Nephrotoxicity and hepatotoxicity of tris (2,3-dibromopropyl)phosphate in the rat. Toxicol. Appl. Pharmacol. 1980, 56, 171–181. BSEF. Bromine Science and Environmental Forum. Legislation on BFRs in North America. http://www.bsef.com/regulation/north-america/ (accessed June 21, 2010). Law, R. J.; Allchin, C. R.; de Boer, J.; Covaci, A.; Herzke, D.; Lepom, P.; Morris, S.; Tronczynski, J.; de Wit, C. A. Levels and trends of brominated flame retardants in the European environment. Chemosphere 2006, 64, 187–208. Kodavanti, P. R. S.; Ward, T. R.; Ludewig, G.; Robertson, L. W.; Birnbaum, L. S. Polybrominated diphenyl ether (PBDE) effects in rat neuron cultures: 14C-PBDE accumulation, biological effects, and structure–activity relationships. Toxicol. Sci. 2005, 88, 181–192. Stoker, T. E.; Cooper, R. L.; Lambright, C. S.; Wilson, V. S.; Furr, J.; Gray, L. E. In vivo and in vitro anti-androgenic effects of DE-71, A commercial polybrominated diphenyl ether (PBDE) Mixture. Toxicol. Appl. Pharmacol. 2005, 207, 78–88. Costa, L. G.; Giordano, G.; Tagliaferri, S.; Caglieri, A.; Mutti, A. Polybrominated diphenyl ether (PBDE) flame retardants: Environmental contamination, human body burden and potential adverse health effects. Acta Biomed. 2008, 79, 172–183. Szabo, D. T.; Richardson, V. M.; Ross, D. G.; Diliberto, J. J.; Kodavanti, P. R. S.; Birnbaum, L. S. Effects of perinatal PBDE exposure on hepatic phase I, phase II, phase III, and deiodinase 1 gene expression involved in thyroid hormone metabolism in male rat pups. Toxicol. Sci. 2009, 107, 27–39. Kodavanti, P. R. S.; Coburn, C. G.; Moser, V. C.; Mac Phail, R. C.; Fenton, S. E.; Stoker, T. E.; Rayner, J. L.; Kannan, K.; Birnbaum, L. S. Developmental exposure to a commercial PBDE mixture, DE-71: Neurobehavioral, hormonal, and reproductive effects. Toxicol. Sci. 2010, 116, 297–312. Zweidinger R. A.; Cooper S. D.; Erickson M. D. In Monitoring Toxic Substances; Schuetzle D., Ed.; ACS Symposium Series 94; American Chemical Society: Washington, DC, 1979; pp 217–231. Watanabe, I.; Kashimoto, T.; Tatsukawa, R. The flame retardant tetrabromobisphenol A and its metabolite found in river and marine sediments in Japan. Chemosphere 1983, 12, 1533–1539. Geyer, H. J.; Schramm, K-W.; Darnerud, P. O.; Aune, M.; Ernst, A. F.; Fried, K. W.; Henkelmann, B.; Lenoir, D.; Schmid, P.; McDonald, T. Terminal elimination of half-lives of the brominated flame retardants TBBPA, HBCD, and lower brominated PBDEs in humans. Organohalogen Compd. 2004, 66, 3820–3825. 38

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

16. National Research Council. Hexabromocyclododecane. In Toxicological Risks of Selected Flame-Retardant Chemicals; National Academy Press: Washington, DC, 2000; pp 53–71. 17. MacGregor J. A.; Nixon W. B.; Hexabromocyclododecane (HBCD): Determination of n-Octanol/Water Partition Coefficient; Wildlife International, Ltd. Project No. 439C-104; Brominated Flame Retardant Industry Panel, Chemical Manufacturers Association: Arlington, VA, 1977. 18. Lyman W. J. Handbook of Chemical Property Estimation Methods; American Chemical Society: Washington, DC, 1990. 19. Veith, G. D.; Defoe, D. L. Measuring and estimating the bioconcentration factor of chemicals in fish. J. Fish. Res. Board Can. 1979, 36, 1040–1048. 20. Environmental Health Criteria 162: Brominated Diphenyl Ethers, 1994. IPCS (International Program on Chemical Safety). http://www.inchem.org/ documents/ehc/ehc/ehc162.htm (accessed June 2016). 21. Wang, X. M.; Ding, X.; Mai, B. X.; Xie, Z. Q.; Xiang, C. H.; Sun, L. G.; Sheng, G. Y.; Fu, J. M.; Zeng, E. Y. Polybrominated diphenyl ethers in airborne particulates collected during a research expedition from the Bohai Sea to the Arctic. Environ. Sci. Technol. 2005, 39, 7803–7809. 22. Hale, R. C.; La Guardia, M. J.; Harvey, E. P.; Mainor, T. M.; Duff, W. H.; Gaylor, M. O. Polybrominated diphenyl ether flame retardants in Virginia freshwater fishes (USA). Environ. Sci. Technol. 2001, 35, 4585–4591. 23. Biosolids Generation, Use and Disposal in the United States; EPA530-R-99009; U.S. Environmental Protection Agency: Washington, DC, 1999. 24. Schecter, A.; Papke, O.; Tung, K. C.; Staskal, D.; Birnbaum, L. S. Polybrominated diphenyl ethers contamination of United States food. Environ. Sci. Technol. 2004, 38, 5306–5311. 25. Butt, C. M.; Diamond, M. L.; Truong, J.; Ikonomou, M. G.; ter Schure, A. F. Spatial distribution of polybrominated diphenyl ethers in southern Ontario as measured in indoor and outdoor window organic films. Environ. Sci. Technol. 2004, 38, 724–731. 26. de Wit, C. A. An overview of brominated flame retardants in the environment. Chemosphere 2002, 46, 583–624. 27. Moon, H.-B.; Kannan, K.; Choi, M.; Choi, H.-G. Polybrominated diphenyl ethers (PBDEs) in marine sediments from industrialized bays of Korea. Marine Pollut. Bull. 2007, 54, 1402–1412. 28. Schauer, U. M.; Volkel, W.; Dekant, W. Toxicokinetics of tetrabromobisphenol A in humans and rats after oral administration. Toxicol. Sci. 2006, 91, 49–58. 29. Meeker, J. D.; Stapleton, H. M. House dust concentrations of organophosphate flame retardants in relation to hormone levels and semen quality parameters. Environ. Health Perspect. 2010, 118, 318–23. 30. Makinen, M. S.; Makinen, M. R.; Koistinen, J. T.; Pasanen, A. L.; Pasanen, P. O.; Kalliokoski, P. J.; Korpi, A. M. Respiratory and dermal exposure to organophosphorus flame retardants and tetrabrompbisphenol a at five work environments. Environ. Sci. Technol. 2009, 43, 941–947. 31. EFSA. Scientific opinion on Tetrabromobisphenol A (TBBPA) and its derivatives in food. EFSA J. 2011, 9, 2477 (1-67). 39 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

32. Fuji, Y.; Nishimura, E.; Kato, Y.; Harada, K. H.; Koisumi, A.; Hiraguchi, K. Dietary exposure to phenolic and methoxylated organohalogen contaminants in relation to their concentrations in breast milk and serum in Japan. Environ. Int. 2014, 63, 19–25. 33. Cariou, R; Antignac, JP; Zalko, D.; Berrebi, A.; Cravedi, J. P.; Maume, D.; Marchand, P.; Monteau, F.; Riu, A.; Andre, F.; le Bizec, B. Exposure assessment of French women and their newborns to tetrabromobisphenol-A: Occurrence measurements in maternal adipose tissue, serum, breast milk and cord serum. Chemosphere 2008, 73, 1036–1041. 34. Knudsen, G. A.; Sanders, J. M.; Sadik, A. M.; Birnbaum, L. S. Disposition and kinetics of tetrabromobisphenol A in female Wistar Han rats. Toxicol. Rep. 2014, 1, 214–223. 35. UNEP Stockholm Convention HBCD Risk Management Evaluation. http://chm.pops.int/Convention/POPsReviewCommittee/Chemicals/tabid/ 243/Default.aspx (accessed November 2016). 36. Covaci, A; Gerecke, AC; Law, RJ; Voorspoels, S; Kohler, M; Heeb, NV; Leslie, H; Allchin, CR; De Boer, J. Hexabromocyclododecanes (HBCDs) in the environment and humans: A review. Environ. Sci. Technol. 2006, 40, 3679–3688. 37. Leonards, P. E. G.; Santillo D.; Brigden K,; Veen, Ivd.; Hesselingen, J. V.; Boer, J. de.; Johnston, P. Brominated Flame Retardants in Office Dust Samples. [Abstract]. Presented at the Second International Workshop on Brominated Flame Retardants, May 14–16, 2001, Stockholm, Sweden. 38. Schecter, A.; Haffner, D.; Colacino, J.; Patel, K.; Papke, O.; Opel, M.; Birnbaum, L. S. Polybrominated diphenyl ethers (PBDEs) and hexabromocyclodecane (HBCD) in composite U.S. food samples. Environ. Health Perspect. 2010, 118, 357–362. 39. Janák, K.; Covaci, A.; Voorspoels, S.; Becher, G. Hexabromocyclododecane in marine species from the Western Scheldt Estuary: Diastereoisomerand enantiomer-specific accumulation. Environ. Sci. Technol. 2005, 39, 1987–1994. 40. Xian, Q.; Ramu, K.; Isobe, T.; Sudaryanto, A.; Liu, X.; Gao, Z.; Takahashi, S.; Yu, H.; Tanabe, S. Levels and body distribution of polybrominated diphenyl ethers (PBDEs) and hexabromocyclododecanes (HBCDs) in freshwater fishes from the Yangtze River, China. Chemosphere 2007, 71, 268–276. 41. Roosens, L.; Abdallah, M. A.; Harrad, S.; Neels, H.; Covaci, A. Exposure to hexabromocyclododecanes (HBCDs) via dust ingestion, but not diet, correlates with concentrations in human serum: Preliminary results. Environ Health Perspect. 2009, 117, 1707–1712. 42. Song, W. L.; Ford, J. C.; Li, A.; Mills, W. J.; Buckley, D. R.; Rockne, K. J. Polybrominated diphenyl ethers in the sediments of the Great Lakes. 1. Lake Superior. Environ. Sci. Technol. 2004, 38, 3286–3293. 43. Song, W. L.; Li, A.; Ford, J. C.; Sturchio, N. C.; Rockne, K. J.; Buckley, D. R.; Mills, W. J. Polybrominated diphenyl ethers in the sediments of the great lakes. 2. Lakes Michigan and Huron. Environ. Sci. Technol. 2005, 39, 3474–3479. 40 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

44. Qiu, X.; Marvin, C. H.; Hites, R. Dechlorane plus and other flame retardants in a sediment core from Lake Ontario. Environ. Sci. Technol. 2007, 41, 6014–6019. 45. Zhu, L. Y.; Hites, R. A. Brominated flame retardants in sediment cores from lakes Michigan and Erie. Environ. Sci. Technol. 2005, 39, 3488–3494. 46. Kimbrough, K. L.; Johnson, W. E.; Lauenstein, G. G.; Christensen, J. D.; Apeti, D. A. 2009. Mussel Watch Program. An Assessment of Polybrominated Diphenyl Ethers (PBDEs) in Sediment and Bivalves of the U.S. Coastal Zone. https://archive.org/stream/ AnAssessmentOfPolybrominatedDiphenylEthersInBivalves/ MWPAnAssessmentofPolybrominatedDiphenylEthersinBivalves_djvu.txt (accessed May 2016). 47. Hites, R. A. 2006. The Handbook of Environmental Chemistry; Springer: Berlin/Heidelberg, Germany, 2006; Vol. 5N, pp 355–390. 48. Toms, L. M. L.; Mortimer, M.; Symons, R. K.; Paepke, O.; Mueller, J. F. Polybrominated diphe- nyl ethers (PBDEs) in sediment by salinity and landuse type from Australia. Environ. Int. 2008, 34, 58–66. 49. Minh, N. H.; Isobe, T.; Ueno, D.; Matsumoto, K.; Mine, M.; Kajiwara, N.; Takahashi, S.; Tanabe, S. Spatial distribution and vertical profile of polybrominated diphenyl ethers and hexabromocyclodo- decanes in sediment core from Tokyo Bay, Japan. Environ. Pollut. 2007, 148, 409–417. 50. Chen, S-J.; Luo, X-J.; Lin, Z.; Luo, Y.; Li, K-C.; Peng, X-Z.; Mai, B-X.; Ran, Y.; Zeng, E. Time trends of polybrominated diphenyl ethers in sediment cores from the Pearl River estuary, South China. Environ. Sci. Technol. 2007, 41, 5595–5600. 51. Sadi, B. B.; Loganathan, B. G. In Global Contamination Trends of Persistent Organic Chemicals; Loganathan, B. G., Lam, P. K. S., Eds.; CRC Press, Taylor & Francis Group: Boca Raton, FL, 2012; pp 481−494. 52. Drage, D.; Mueller, J. F.; Birch, G.; Eaglesham, G.; Hearn, L. K.; Harrad, S. Historical trends of PBDEs and HBCD in sediment cores from Sydney estuary, Australia. Sci. Total Environ. 2015, 512-13, 177–184. 53. La-Guardia, M. J.; Hale, R. C; Harvey, E. Detailed polybrominated diphenyl ether (PBDE) congener composition of the widely used penta-, octa-, and deca-PBDE technical flame-retardant mix-tures. Environ. Sci. Technol. 2006, 40, 6247–6254. 54. Yun, S. H.; Addink, R.; McCabe, J. M.; Ostaszewski, A.; MackenzieTaylor, D.; Taylor, A. B.; Kannan, K. Polybrominated diphenyl ethers and polybrominated biphenyls in sediment and flood-plain soils of the Saginaw River watershed, Michigan, USA. Arch. Environ. Contam. Toxicol. 2008, 55, 1–10. 55. Wang, P.; Zhang, Q. H.; Wang, Y. W.; Wang, T.; Li, X. M.; Li, Y. M.; Ding, L.; Jiang, G. B. Altitude dependence of polychlorinated biphenyls (PCBs) and polybrominated diphenyl ethers (PBDEs) in surface soil from Tibetan Plateau, China. Chemosphere 2009, 76, 1498–1504. 56. Luo, Y.; Luo, X. J.; Lin, Z.; Chen, S. J.; Liu, J.; Mai, B. X.; Yang, Z. Y. Polybrominated diphenyl ethers in road and farmland soils from an e-waste recycling region in Southern China: Concentrations, source profiles, 41 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

57.

58.

59.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

60.

61.

62.

63.

64.

65.

66.

67.

68.

69.

and potential dispersion and deposition. Sci. Total Environ. 2009, 407, 1105–1113. Batterman, S. A.; Chernyak, S.; Jia, C. R.; Godwin, C.; Charles, S. Concentrations and emissions of polybrominated diphenyl ethers from US houses and garages. Environ. Sci. Technol. 2009, 43, 2693–2700. Takigami, H.; Suzuki, G.; Hirai, Y.; Sakai, S. Transfer of brominated flame retardants from components into dust inside television cabinets. Chemosphere 2008, 73, 161–169. Huang, Y. M.; Chen, L. G.; Peng, X. C.; Xu, Z. C.; Ye, Z. X. PBDEs in indoor dust in South- Central China: Characteristics and implications. Chemosphere 2010, 78, 169–174. 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.; Mai, B. X. Occurrence of brominated flame retardants other than polybrominated diphenyl ethers in environmental and biota samples from southern China. Chemosphere 2009, 74, 910–916. Wilford, B. H.; Shoeib, M.; Harner, T.; Zhu, J.; Jones, K. C. Polybrominated diphenyl ethers in indoor dust in Ottawa, Canada: Implications for sources and exposure. Environ. Sci. Technol. 2005, 39, 7027–7035. Harrad, S.; Ibarra, C.; Diamond, M.; Melymuk, L.; Robson, M.; Douwes, J.; Roosens, L.; Dirtu, A. C.; Covaci, A. Polybrominated diphenyl ethers in domestic indoor dust from Canada, New Zealand, United Kingdom and United States. Environ. Int. 2008, 34, 232–238. Suzuki, G.; Takigami, H.; Takahashi, S.; Sakai, S. PBDEs and PBDD/Fs in house and office dust from Japan. Organohalogen Compd. 2006, 68, 1843–1846. Tan, J.; Cheng, S. M.; Loganath, A.; Chong, Y. S.; Obbard, J. P. Polybrominated diphenyl ethers in house dust in Singapore. Chemosphere 2007, 66, 985–992. Sjödin, A.; Päpke, O.; McGahee, E.; Focant, J. F.; Jones, R. S.; Pless-Mulloli, T.; Toms, L. M. L.; et al. Concentration of polybrominated diphenyl ethers (PBDEs) in household dust from various countries. Chemosphere 2008, 73, S131–S136. Johnson-Restrepo, B.; Kannan, K. An assessment of sources and pathways of human exposure to polybrominated diphenyl ethers in the United States. Chemosphere 2009, 76, 542–548. Allen, J. G.; McClean, M. D.; Stapleton, H. M.; Webster, T. F. Critical factors in assessing exposure to PBDEs via house dust. Environ. Int. 2008, 34, 1085–1091. Monirith, I.; Ueno, D.; Takahashi, S.; Nakata, H.; Sudaryanto, A; Subramanian, A.; Karuppiah, S.; Ismail, A.; Muchtar, M.; Zheng, J.; Richardson, B. J.; Prudente, M.; Hue, N. D.; Tana, T. S.; Tkalin, A. V.; Tanabe, S. Asia-Pacific mussel watch: Monitoring contamination of persistent organochlorine compounds in coastal waters of Asian countries. Mar. Pollut. Bull. 2003, 46, 281–300. Ramu, K.; Kajiwara, N.; Sudaryanto, A.; Isobe, T.; Takahashi, S.; Subramanian, A.; Ueno, D.; Zheng, G. J.; Lam, P. K.; Takada, H.; Zakaria, M. P.; Vie, P. H.; Prudente, M.; Tana, T. S.; Tanabe, S. Asian mussel watch 42

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

70.

71.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

72.

73.

74.

75.

76.

77.

78.

79.

80.

81.

82.

program: Contamination status of polybrominated diphenyl ethers and organochlorines in coastal waters of Asian countries. Environ. Sci. Technol. 2007, 41, 4580–4586. Lapviboonsuk, J.; Loganathan, B. G. Polynuclear aromatic hydrocarbons in sediments and mussel tissues from the lower Tennessee River. J. Ky. Acad. Sci. 2007, 68, 186–197. Oros, D. R.; Hoover, D.; Rodigari, F.; Crane, D.; Sericano, J. Levels and distribution of poly-brominated diphenyl ethers in water, surface sediments, and bivalves from the San Francisco Estuary. Environ. Sci. Technol. 2005, 39, 33–41. deBruyn, A. M. H.; Meloche, L. M.; Lowe, C. J. Patterns of bioaccumulation of polybrominated diphenyl ether and polychlorinated biphenyl congeners in marine mussels. Environ. Sci. Technol. 2009, 43, 3700–3704. Tittlemier, S. A.; Forsyth, D.; Breakell, K.; Verigin, V.; Ryan, J. J.; Hayward, S. Polybrominated diphenyl ethers in retail fish and shellfish samples purchased from Canadian markets. J. Agric. Food Chem. 2004, 52, 7740–7745. Law, K.; Halldorson, T.; Danell, R.; Stern, G.; Gewurtz, S.; Alaee, M.; Marvin, C.; Whittle, M.; Tomy, G. Bioaccumulation and trophic transfer of some brominated flame retardants in a Lake Winnipeg (Canada) food web. Environ. Toxicol. Chem. 2006, 25, 2177–2186. Wang, Z.; Ma, X.; Lin, Z.; Na, G.; Yao, Z. Congener specific distributions of polybrominated diphenyl ethers (PBDEs) in sediment and mussel (Mytilus edulis) of the Bo Sea, China. Chemosphere 2009, 74, 896–901. Pan, J.; Yang, Y.-L.; Xu, Q.; Chen, D.-Z.; Xi, D.-L. PCBs, PCNs and PBDEs in sediments and mussels from Qingdao coastal sea in the frame of current circulations and influence of sewage sludge. Chemosphere 2007, 66, 1971–1982. Liu, Y.; Zheng, G. J.; Yu, H.; Martin, M.; Richardson, B. J.; Lam, M. H. W.; Lam, P. K. S. Polybrominated diphenyl ethers (PBDEs) in sediments and mussel tissues from Hong Kong marine waters. Mar. Pollut. Bull. 2005, 50, 1173–1184. Guo, J.-Y.; Zeng, E. Y.; Mai, B.-X.; Luo, X.-J.; Wu, F.-C. Polybrominated diphenyl ethers in seafood products of South China. J. Agric. Food Chem. 2007, 55, 9152–9158. Johnson-Restrepo, B.; Kannan, K.; Addink, R.; Adams, D. H. Polybrominated diphenyl ethers and polychlorinated biphenyls in a marine foodweb of coastal Florida. Environ. Sci. Technol. 2005, 39, 8243–8250. Shaw, S. D.; Berger, M. L.; Brenner, D.; Kannan, K.; Lohmann, N.; Papke, O. Bioaccumulation of polybrominated diphenyl ethers and hexabromocyclododecane in the northwest Atlantic marine food web. Sci. Total Environ. 2009, 407, 3323–3329. Dodder, N. G.; Strandberg, B.; Hites, R. A. Concentrations and spatial variation of polybrominated diphenyl ethers in fish and air from the Northeastern United States. Organohalogen Compd. 2000, 47, 69–72. Braune, B. M.; Outridge, P. M.; Fisk, A. T.; Muir, D. C. G.; Helm, P. A.; Hobbs, K.; Hoekstra, P. F.; Kuzyk, Z. A.; Kwan, M.; Letcher, R. J.; 43

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

83.

84.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

85.

86.

87.

88.

89.

90.

91.

92.

93.

94.

Lockhart, W. L.; Norstrom, R. J.; Stern, G. A.; Stirling, I. Persistent organic pollutants and mercury in marine biota of the Canadian Arctic: An overview of spatial and temporal trends. Sci. Total Environ. 2005, 351–352, 4–56. Asplund, L.; Hornung, M.; Peterson, R. E.; Turesson, K.; Bergman, Å. Levels of polybrominated diphenyl ethers (PBDEs) in fish from the Great Lakes and Baltic Sea. Organohalogen Compd. 1999, 40, 351–354. Holden, A.; She, J. W.; Tanner, M.; Lunder, S.; Sharp, R.; Hooper, K. PBDEs in the San Francisco Bay area: Measurements in fish. Organohalogen Compd. 2003, 61, 255–258. Brown, F. R.; Winkler, J.; Visita, P.; Dhaliwal, J.; Petreas, M. Levels of PBDEs, PCDDs, PCDFs, and coplanar PCBs in edible fish from California coastal waters. Chemosphere 2006, 64, 276–286. Rice, C. P.; Chernyak, S. M.; Begnoche, L.; Quintal, R.; Hickey, J. Comparisons of PBDE composition and concentration in fish collected from the Detroit River, MI and Des Plaines River, IL. Chemosphere 2002, 49, 731–737. Boon, J. P.; Lewis, W. E.; Tjoen-A-Choy, M. R.; Allchin, C. R.; Law, R. J.; de Boer, J.; ten Hallers- Tjabbes, C. C.; Zegers, B. N. Levels of polybrominated diphenyl ether (PBDE) flame retar- dants in animals representing different trophic levels of the North Sea food web. Environ. Sci. Technol. 2002, 36, 4025–4032. Easton, M. D. L.; Luszniak, D.; Von der Geest, E. Preliminary examination of contaminant loadings in framed salmon, wild salmon and commercial salmon feed. Chemosphere 2002, 46, 1053–1074. Manchester-Neesvig, J. B.; Valters, K.; Sonzogni, W. C. Comparison of polybrominated diphenyl ethers (PBDEs) and polychlorinated biphenyls (PCBs) in Lake Michigan salmonids. Environ. Sci. Technol. 2001, 35, 1072–1077. Luross, J. M.; Alaee, M.; Sergeant, D. B.; Cannon, C. M.; Whittle, D. M.; Solomon, K. R.; Muir, D. C. G. Spatial distribution of polybrominated diphenyl ethers and polybrominated biphenyls in lake trout from the Laurentian Great Lakes. Chemosphere 2002, 46, 665–672. Zhu, L. Y.; Hites, R. A. Temporal trends and spatial distributions of brominated flame retardants in archived fishes from the Great Lakes. Environ. Sci. Technol. 2004, 38, 2779–2784. Ikonomou, M. G.; Rayne, S.; Fischer, M.; Fernandez, M. P.; Cretney, W. Occurrence and con- gener profiles of polybrominated diphenyl ethers (PBDEs) in environmental samples from coastal British Columbia, Canada. Chemosphere 2002, 46, 649–663. Ikonomou, M. G.; Fernandez, M. P.; Hickman, Z. L. Spatio-temporal and species-specific variation in PBDE levels/patterns in British Columbia’s coastal waters. Environ. Pollut. 2006, 140, 355–363. Ueno, D.; Kajiwara, N.; Tanaka, H.; Subramanian, A.; Fillmann, G.; Lam, P. K. S.; Zheng, G. J.; Muchitar, M.; Razak, H.; Prudente, M.; Chung, K-H.; Tanabe, S. Global pollution monitoring of polybrominated diphenyl ethers using Skipjack Tuna as a bioindicator. Environ. Sci. Technol. 2004, 38, 2312–2316. 44

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

95. Meng, X.-Z.; Zeng, E. Y.; Yu, L.-P.; Mai, B.-X.; Luo, X.-J.; Ran, Y. Persistent halogenated hydrocarb;ns in consumer fish of China: Regional and global implications for human exposure. Environ. Sci. Technol. 2007, 41, 1821–1827. 96. Xiang, C.-H.; Luo, X.-J.; Chen, S.-J.; Yu, M.; Mai, B.-X.; Zeng, E. Y. Polybrominated diphenyl ethers in the biota and sediments of the Pearl River Estuary, South China. Environ. Toxicol. Chem. 2007, 26, 616–623. 97. Akutsu, K.; Obana, H.; Okihashi, M.; Kitagawa, M.; Nakazawa, H.; Matsuki, Y.; Makino, T.; Oda, H.; Hori, S. GC/MS analysis of polybrominated diphenyl ethers in fish collected from the Inland Sea of Seto, Japan. Chemosphere 2001, 44, 1325–1333. 98. Losada, S.; Roach, A.; Roosens, L.; Santos, F. J.; Galceran, M. T.; Vetter, W.; Neels, H.; Covaci, Biomagnification of anthropogenic and naturally-produced organobrominated compounds in a marine food web from Sydney Harbour, Australia. Environ. Int. 2009, 35, 1142–1149. 99. Hermanussen, S.; Matthews, V.; Paepke, O.; Limpus, C. J.; Gaus, C. Flame retardants (PBDEs) in marine turtles, dugongs and seafood from Queensland, Australia. Mar. Pollut. Bull. 2008, 57, 409–418. 100. Hites, R. A.; Foran, J. A.; Schwager, S. J.; Knuth, B. A.; Hamilton, M. C.; Carpenter, D. O. Global assessment of polybrominated diphenyl ethers in farmed and wild salmon. Environ. Sci. Technol. 2004, 38, 4945–4949. 101. Zennegg, M.; Kohler, M.; Gerecke, A. C.; Schmid, P. Polybrominated diphenyl ethers in white-fish from Swiss lakes and farmed rainbow trout. Chemosphere 2003, 51, 545–553. 102. Shaw, S. D.; Berger, M. L.; Brenner, D.; Carpenter, D. O.; Tao, L.; Hong, C. S.; Kannan, K. Polybrominated diphenyl ethers (PBDEs) in farmed and wild salmon marketed in the Northeastern United States. Chemosphere 2008, 71, 1422–1431. 103. Loganthan, B. G.; Kannan, K. Global organochlorine contamination trends: An overview. AMBIO 1994, 23, 187–191. 104. Luo, Q.; Wong, M. H.; Cai, Z. W. Determination of polybrominated diphenyl ethers in fresh-water fishes from a river polluted by e-wastes. Talanta 2007, 72, 1644–1649. 105. Wu, J.-P.; Luo, X.-J.; Zhang, Y.; Luo, Y.; Chen, S.-J.; Mai, B.-X.; Yang, Z.-Y. Bioaccumulation of polybrominated diphenyl ethers (PBDEs) and polychlorinated biphenyls (PCBs) in wild aquatic species from an electronic waste (e-waste) recycling site in South China. Environ. Int. 2008, 34, 1109–1113. 106. Dodder, N. G.; Strandberg, B.; Hites, R. A. Concentrations and spatial variations of polybro- minated diphenyl ethers and several organochlorine compounds in fishes from the Northeastern United States. Environ. Sci. Technol. 2002, 36, 146–151. 107. She, J. W.; Holden, A.; Adelsbach, T. L.; Tanner, M.; Schwarzbach, S. E.; Yee, J. L.; Hooper, K. Concentrations and time trends of polybrominated diphenyl ethers (PBDEs) and polychlorinated biphenyls (PCBs) in aquatic bird eggs from San Francisco Bay, CA 2000–2003. Chemosphere 2008, 73, S201–S209. 45 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

108. Norstrom, R. J.; Simon, M.; Moisey, J.; Wakeford, B.; Weseloh, D. V. C. Geographical distribution (2000) and temporal trends (1981–2000) of brominated diphenyl ethers in Great Lakes Herring Gull eggs. Environ. Sci. Technol. 2002, 36, 4783–4789. 109. Lam, J. C. W.; Kajiwara, N.; Ramu, K.; Tanabe, S.; Lam, P. K. S. Assessment of polybrominated diphenyl ethers in eggs of waterbirds from South China. Environ. Pollut. 2007, 148, 258–267. 110. Gao, F.; Luo, X.-J.; Yang, Z.-F.; Wang, X.-M.; Mai, B.-X. Brominated flame retardants, poly-chlorinated diphenyls, and organochlorine pesticides in bird eggs from the Yellow River Delta, North China. Environ. Sci. Technol. 2009, 43, 6956–6962. 111. Chen, D.; La Guardia, M. J.; Harvey, E.; Amaral, M.; Wohlfort, K.; Hale, R. C. Polybrominated diphenyl ethers in Peregrine Falcon (Falco peregrinus) eggs from the Northeastern US. Environ. Sci. Technol. 2008, 42, 7594–7600. 112. Park, J. S.; Holden, A.; Chu, V.; Kim, M.; Rhee, A.; Patel, P.; Shi, Y.; Linthicum, L.; Waltson, B. J.; McKeown, K.; Jewell, N. P.; Hooper, K. Time-trends and congener profiles of PBDEs and PCBs in California Peregrine Falcons (Falco peregrinus). Environ. Sci. Technol. 2009, 43, 8744–8751. 113. Chen, D.; Mai, B.; Song, J.; Sun, Q.; Luo, Y.; Luo, X.; Zeng, E. Y.; Hale, R. C. Polybrominated diphenyl ethers in birds of prey from Northern China. Environ. Sci. Technol. 2007, 41, 1828–1833. 114. Elliott, J. E.; Wilson, L. K.; Wakeford, B. Polybrominated diphenyl ether trends in eggs of marine and freshwater birds from British Columbia, Canada, 1979–2002. Environ. Sci. Technol. 2005, 39, 5584–5591. 115. Muir, D. C. G.; Backus, S.; Derocher, A. E.; Dietz, R.; Evans, T. J.; Gabrielsen, G. W.; Nagy, J.; Norstrom, R. J.; Sonne, C.; Stirling, I.; Taylor, M. K.; Letcher, R. J. Brominated flame retardants in polar bears (Ursus maritimus) from Alaska, the Canadian Arctic, East Greenland, and Svalbard. Environ. Sci. Technol. 2006, 40, 449–455. 116. Kannan, K.; Yun, S. H.; Evans, T. J. Chlorinated, brominated, and perfluorinated contaminants in livers of polar bears from Alaska. Environ. Sci. Technol. 2005, 39, 9057–9063. 117. Bentzen, T. W.; Muir, D. C. G.; Amstrup, S. C.; O’Hara, T. M. Organohalogen concentrations in blood and adipose tissue of Southern Beaufort Sea polar bears. Sci. Total Environ. 2008, 406, 352–367. 118. Dietz, R.; Rigét, F. F.; Sonne, C.; Letcher, R. J.; Backus, S.; Born, E. W.; Kirkegaard, M.; Muir, D. C. G. Age and seasonal variability of polybrominated diphenyl ethers in free-ranging East Greenland polar bears (Ursus maritimus). Environ. Pollut. 2007, 146, 166–173. 119. Verreault, J.; Gabrielsen, G. V.; Chu, S. G.; Muir, D. C. G.; Andersen, M.; Hamaed, A.; Letcher, R. J. Flame retardants and methoxylated and hydroxylated polybrominated diphenyl ethers in two Norwegian Arctic top predators: Glaucous gulls and polar bears. Environ. Sci. Technol. 2005, 39, 6021–6028. 120. Kajiwara, N.; Ueno, D.; Takahashi, A.; Baba, N.; Tanabe, S. Polybrominated diphenyl ethers and organochlorines in archived northern fur seal samples 46 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

121.

122.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

123.

124.

125.

126.

127.

128.

129.

130.

131.

from the Pacific coast of Japan, 1972–1998. Environ. Sci. Technol. 2004, 38, 3804–3809. Meng, X.-Z.; Blasius, M. E.; Gossett, R. W.; Maruya, K. A. Polybrominated diphenyl ethers in pinnipeds stranded along the southern California coast. Environ. Pollut. 2009, 157, 2731–2736. Shaw, S. D.; Kannan, K. Polybrominated diphenyl ethers in marine ecosystems of the American continents: Foresight from current knowledge. Rev. Environ. Health 2009, 24, 157–229. Kajiwara, N.; Kamikawa, S.; Ramu, K.; Ueno, D.; Yamada, T. K.; Subramanian, A.; Lam, P. K. S.; Jefferson, T. A.; Prudente, M.; Chung, K-H.; Tanabe, S. Geographical distribution of polybrominated diphenyl ethers (PBDEs) and organochlorines in small cetaceans from Asia waters. Chemosphere 2006, 64, 287–295. Stapleton, H. M.; Dodder, N. G.; Kucklick, J. R.; Reddy, C. M.; Schantz, M. M.; Becker, P. R.; Gulland, F.; Porter, B. J.; Wise, S. A. Determination of HBCD, PBDEs and MeO-BDEs in California sea lions (Zalophus californianus) stranded between 1993 and 2003. Mar. Pollut. Bull. 2006, 52, 522–531. Kannan, K.; Perrotta, E.; Thomas, N. J.; Aldous, K. M. A comparative analysis of polybrominated diphenyl ethers and polychlorinated biphenyls in southern sea otters that died of infectious diseases and noninfectious causes. Arch. Environ. Contam. Toxicol. 2007, 53, 293–302. She, J. W.; Petreas, M.; Winkler, J.; Visita, P.; McKinney, M.; Kopec, D. PBDEs in the San Francisco Bay Area: Measurements in harbor seal blubber and human breast adipose tissue. Chemosphere 2002, 46, 697–707. Ikonomou, M. G.; Rayne, S.; Addison, R. F. Exponential increases of the brominated flame retardants, polybrominated diphenyl ethers, in the Canadian Arctic from 1981 to 2000. Environ. Sci. Technol. 2002, 36, 1886–1892. Lam, J. C. W.; Lau, R. K. F.; Murphy, M. B.; Lam, P. K. S. Temporal trends of hexabromocy- clododecanes (HBCDs) and polybrominated diphenyl ethers (PBDEs) and detection of two novel flame retardants in marine mammals from Hong Kong, South China. Environ. Sci. Technol. 2009, 43, 6944–6949. Lebeuf, M.; Gouteux, B.; Measures, L.; Trottier, S. Levels and temporal trends (1988–1999) of polybrominated diphenyl ethers in Beluga whales (Delphinapterus leucas) from the St. Lawrence estuary, Canada. Environ. Sci. Technol. 2004, 38, 2971–2977. Dorneles, P. R.; Lailson-Brito, J.; Dirtu, A. C.; Weijs, L.; Azevedo, A. F.; Torres, J. P. M.; Malm, O.; Neels, H.; Blust, R.; Das, K.; Covaci, A. Anthropogenic and naturally-produced organobrominated compounds in marine mammals from Brazil. Environ. Int. 2010, 36, 60–67. Isobe, T.; Ochi, Y.; Ramu, K.; Yamamoto, T.; Tajima, Y.; Yamada, T. K.; Amano, M.; Miyazaki, N.; Takahashi, S.; Tanabe, S. Organohalogen contaminants in striped dolphins (Stenella coeruleoalba) from Japan: Present contamination status, body distribution and temporal trends (1978–2003). Mar. Pollut. Bull. 2009, 58, 396–401. 47

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch002

132. Kajiwara, N.; Kamikawa, S.; Amano, M.; Hayano, A.; Yamada, T. K.; Miyazaki, N.; Tanabe, S. Polybrominated diphenyl ethers (PBDEs) and organochlorines in melon-headed whales, Peponocephala electra, mass stranded along the Japanese coasts: Maternal transfer and temporal trend. Environ. Pollut. 2008, 156, 106–114. 133. Ramu, K.; Kajiwara, N.; Lam, P. K. S.; Jefferson, T. A.; Zhou, K.; Tanabe, S. Temporal varia- tion and biomagnification of organohalogen compounds in finless porpoises (Neophocaena phocaenoides) from the South China Sea. Environ. Pollut. 2006, 144, 516–523. 134. Exporting Harm: The High-Tech Trashing of Asia, 2002. The Basel Action Network (BAN) and Silicon Valley Toxic Coalition (SVTC). http://www.ban.org/E-waste/technotrashfinalcomp.pdf (accessed May 2016). 135. Hites, R. A. Polybrominated diphenyl ethers in the environment and in people: A meta-analysis of concentrations. Environ. Sci. Technol. 2004, 38, 945–956. 136. Law, R. J.; Herzke, D.; Harrad, S.; Morris, S.; Bersuder, P.; Allchin, C. R. Levels and trends of HBCD and BDEs in the European and Asian environments, with some information for other BFRs. Chemosphere 2008, 73, 223–241. 137. Akutsu, K.; Kitagawa, M.; Nakazawa, H.; Makino, T.; Iwazaki, K.; Oda, H.; Hori, S. Time-trend (1973–2000) of polybrominated diphenyl ethers in Japanese mother’s milk. Chemosphere 2003, 53, 645–654. 138. Choi, J. W.; Fujimaki, S.; Kitamura, K.; Hashimoto, S.; Ito, H.; Suzuki, N.; Sakai, S.; Morita, M. Polybrominated dibenzo-p-dioxins, dibenzofurans, and diphenyl ethers in Japanese human adipose tissue. Environ. Sci. Technol. 2003, 37, 817–821. 139. Toms, L. M. L.; Harden, F. A.; Symons, R. K.; Burniston, D.; Furst, P.; Muller, J. F. Polybrominated diphenyl ethers (PBDEs) in human milk from Australia. Chemosphere 2007, 68, 797–803. 140. Dye, J. A.; Venier, M.; Zhu, L.; Ward, C. R.; Hites, R. A.; Birnbaum, L. S. Elevated PBDE levels in pet cats: Sentinels for humans? Environ. Sci. Technol. 2007, 41, 6350–6356. 141. Liang, S. X.; Zhao, Q.; Qin, Z. F.; Zhao, X. R.; Yang, Z. Z.; Xu, X. B. Levels and distribution of polybrominated diphenyl ethers in various tissues of foraging hens from an electronic waste recycling area in South China. Environ. Toxicol. Chem. 2008, 27, 1279–1283. 142. Hu, G-C.; Luo, X-J.; Dai, H-Y.; Zhang, X-L.; Wu, H.; Zhang, C-L.; Guo, W.; Xu, M-Q.; Mai, B-X.; Wei, F-W. Brominated flame retardants, polychlorinated biphenyls, and organochlorine pesticides in captive giant panda (Ailuropoda melanoleus) and red panda (Ailurus fulgens) from China. Environ. Sci. Technol. 2008, 42, 4704–4709.

48 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Chapter 3

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Dioxins and Furans: Environmental Contamination and Regulatory Status in Colombia Aída L. Villa,* Alexander Quintero, and Diana Pemberthy Environmental Catalysis Research Group, Chemical Engineering Department, Engineering Faculty, Universidad de Antioquia, Calle 70 No. 52-21, Medellín, Colombia *E-mail: [email protected]

This review deals with the regulatory history and the levels of polychlorinated dibenzo-p-dioxins (PCDDs), dibenzofurans (PCDFs), and non- and mono-orthosubstituted polychlorinated biphenyls (dioxin-like PCBs or dl-PCBs) in stack gas emissions, fly ashes, ambient air, particulate matter and several types of foods from Colombia. Colombia was one of the signatory countries of the Stockholm Convention on Persistent Organic Pollutants. Current regulations allow the maximum permissible level of 0.5 ng TEQ/m3 for stationary sources of industrial activities, 0.1 ng TEQ/m3 for hazardous was incinerators and cement kiln wastes, 4.0 pg WHO-TEQ/g fresh weight for fish and fishery products, and 0.75 – 3.0 pg WHO-TEQ/g fat for oils and fats. Based on available literature, fish oil contained the highest level of PCDD/Fs + dl-PCBs (2.40 pg WHO-TEQ/g of fat) followed by shrimp (1.95 pg WHO-TEQ/g) and butter (1.08 pg WHO-TEQ/g), while vegetable oil presented the lowest level (0.36 pg WHO-TEQ/g). Butter was the only product with PCDD/Fs concentration above the maximum level established by the Colombian legislation. From the review is concluded that even though there is some research in Colombia about levels of PCDD/Fs and dl-PCBs in several matrices, it is necessary to determine the concentration of dioxins and furans

© 2016 American Chemical Society Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

in more Colombian regions and matrices in order to better understand the status of dioxins/furans contamination and to prevent further contamination and also to protect wildlife and human health.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Introduction Persistent Organic Pollutants (POPs) are a group of organic chemicals that, due to their physicochemical properties, are toxic, bioaccumulative, persistent and semivolatile (1). The Stockholm Convention on POPs that was created in 2001 under the auspices of the United Nations Environmental Program (2), identified these compounds to eliminate from production and usage. Originally, there were 12 specific chemicals listed under the Stockholm Convention, called the “Dirty Dozen”; nine new chemicals were added in 2009 and two more in 2011. POPs can be transfered to the atmosphere from soils, plants and water bodies; in the atmosphere, POPs exist either in the gaseous or particle-associated form, both of which can facilitate transport over long distances (3). The POPs can be produced intentionally as organochlorine pesticides (OCPs), polychlorinated biphenyls (PCBs) and polybrominated diphenyl ethers (PBDEs) or unintentionally as by-products during chemical manufacture or incineration processes. Polychlorinated dibenzo-p-dioxins (PCDDs) and polychlorinated dibenzofurans (PCDFs) are POPs formed and released unintentionally as derivatives during the production of industrial chemicals such as PCBs, polychlorinated naphthalenes, chlorinated phenols, and polyvinyl chlorides, and in chlorine bleaching in paper making and metal smelting (4–7), as well as forest fires, combustion engines, and home fireplaces (8). It has been reported that municipal solid waste incineration (MSWI) is the main source of PCDDs/Fs to air, followed by open burning processes and heat and power generation in the United Nations Environment Programme (UNEP) participating countries (9). Dibenzo-p-dioxin was first prepared by Lesimple in the year 1866 from triphenylphosphate and lime, and the structure was determined in 1871 by Hoffmeister (10). In 1872, German chemists Mertz and Weitz reported preparation of the first chlorinated dibenzo-p-dioxin; the most toxic congener, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), was reported in the mid-1950s (11). PCDDs/Fs are widely dispersed in the global environment and their presence has been reported in water, air, soil, sediment, and aquatic and terrestrial organisms including human tissues (12, 13). People in developing countries are exposed to high levels of organochlorines from food and air because the continued use of these chemicals. As many developed nations import food stuffs from developing countries, their populations are also exposed to organochlorines through their food. This was concluded from the study of the global contamination trends of POPs that showed a steady state or very slow decline of organochlorine burden in the populations (14); those results have motivated strict regulations with respect to concentration of POPs in foods. This chapter deals with PCDD/Fs contamination and regulatory status in Colombia. Based on available literature, it presents the regulations regarding 50 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

the allowed concentrations of these pollutants in emissions and in some foods. The Ministry of Environment of Colombia used UNEP standardized toolkit to determine PCCD/Fs emissions, as well as the levels of PCDD/Fs in several matrices that have been reported from academic research. The Colombian territory is characterized by a great diversity of ecosystems, it has three mountain ranges and five main natural regions: Caribbean, Andean, Pacific, Orinoquia and Amazonia. Colombia is located at the northwestern tip of South America, consists of thirty-two political divisions (departments) and has 2,070,408 km2 (1,141,748 km2 continental land mass and 958,660 km2 territorial waters). It is the fourth largest country in South America and the only one with Caribbean and Pacific Coasts (15). According to the 2013 analysis, the population of Colombia was around 48 million people. Colombia is sparsely populated with just 41 people per square kilometer. The largest city and capital of Colombia is Bogotá, which has a population of 7.9 million. Other major cities include Medellín (2.5 million), Cali (2.4 million) and Barranquilla (1.2 million) (16).

Regulatory History The Colombian Constitution states that a healthy environment is a right of all people and the State has the duty to protect the diversity and integrity of the environment (Article 79). Because of that, Law 99 - 1993 established the National Environmental System for the environmental management of the country, and created the Ministry of Environment as its coordinator. This Ministry is responsible for the definition of the environmental policies and regulations in Colombia. Specifically, the permissible levels of emission of dioxins and furans from stationary sources were established by the Ministry of Environment in Resolution 909 – 2008 (17). The Colombian legislation history related with dioxins and furans is described in Figure 1.

Figure 1. Regulatory timeline to control dioxin/furan contamination in Colombia. 51 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

The dioxin and furan legislation in Colombia was influenced by the responsibilities associated with The Stockholm Convention on Persistent Organic Pollutants (2). Colombia was one of the signatory countries of this Convention on May 22nd 2001 in Stockholm, Sweden, but just in 2008 the country ratified the Convention through Law 1196. In October 30th 2001 dioxins and furans appeared for the first time in Colombian legislation, in the Resolution 970 of the Ministry of Environment (18). In this resolution, the requirements, conditions and maximum permissible levels of emission for the elimination of plastics contaminated by pesticides in cement kilns for clinker production were established. Specifically, 0.2 ng TEQ/m3 was the emission limit for the seventeen toxic dioxins and furans. The same limit concentration was fixed for the elimination of soil (or similar materials) contaminated by pesticides (Resolution 458 – 2002) (19) and new and used tires (Resolution 1488 – 2003) (20) in cement kilns for clinker production. The next step in the control of dioxin and furans in Colombia was given in Resolution 0058 - 2002 of the Ministry of Environment (21). This Resolution established the maximum permissible levels of emission for solid and liquid waste incinerators. Initially, 1 ng TEQ/m3 was the emission limit for dioxins and furans, but the concentration was gradually reduced up to 0.1 ng TEQ/m3 nine years after the Resolution was valid. For the first time in Colombian legislation, Resolution 0058 – 2002 (21) demanded the use of specific methods for dioxin and furan sampling and analysis (any of these): VDI 3499 part 2 (from German), EN-1948 parts 2 and 3 (from European Community), or EPA 23, 23A, 8280A, 8290 (from USA). On July 27th 2004, the Ministry of Environment issued Resolution 0886 (22) that changed the maximum permissible levels of emission of dioxins previously established in Resolution 0058 – 2002. This was made because the Ministry evaluated the compliance level and consequences of the Resolution, established that the country had an incipient capacity for dioxin analysis and concluded that it was necessary to increase the initial deadlines and levels of emission until gradually reach the final levels. In that sense, Resolution 0886 determined different levels of emissions from incinerators according to their capacities and state (new or old) for specifics years (2005, 2006, 2009 and 2012). The final goal for new incinerators (any capacity) and old incinerators that processed more than 100 kg/h was 0.1 ng TEQ/Nm3, and 0.2 ng TEQ/Nm3 for old incinerators that processed less than 100 kg/h. However, the maximum permissible levels of emission of dioxins changed again in 2008. Resolution 909 – 2008 of the Ministry of Environment, Housing and Territorial Development (17) repealed the previous Resolutions related to dioxins emission and it is currently the valid regulation in Colombia for dioxins and furans emission control. For stationary sources of industrial activities, this Resolution establishes that the permissible emission standard for dioxins is 0.5 ng TEQ/m3 at 25°C and 760 mmHg and reference oxygen content of 11%. The industrial activities include steel, copper, and zinc foundries, coke production, vegetable material ovens and others. According to Resolution 909 – 2008 (17), for hazardous waste incinerators and cement kilns co-processing waste, the standard for dioxins emission is 0.1 ng TEQ/m3 at 25°C and 760 mmHg and reference oxygen content of 11%, whereas 52 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

for non-hazardous waste incinerators is 0.5 ng TEQ/m3 at the same conditions. In addition, the Resolution refers to a protocol for the sampling methods in stationary sources. Specifically, EPA 23 is the valid method in Colombia for dioxins sampling and analysis. The dioxin regulations described before only apply for emissions from stationary sources. Currently, Colombia does not have a regulation for dioxins in mobile sources or air quality, but in order to protect people’s health from the ingestion of dioxin contaminated foods, the Ministry of Health and Social Protection has issued maximum permissible levels for some products. Because Colombia fishery products are considered foods at risk to public health, Resolution 776 – 2008 (23) established 3.0 pg WHO-TEQ/g fresh weight as the action threshold for dioxins and furans in fish and fishery products (fish and shellfish) except eel. However, Resolution 776 was modified in 2012 by Resolution 122 (24) and 4.0 pg WHO-TEQ/g fresh weight was the new requirement for dioxins content in fish and fishery products. If dl-PCBs are included, the maximum levels are 8.0 pg WHO-TEQ/g fresh weight of fish and fishery products, 12.0 pg WHO-TEQ/g fresh weight for eel and 25.0 pg WHO-TEQ/g fresh weight for fish liver and its derived products (except oil from sea organisms). In the same way, dioxin content limits have been stablished for oils and fats in Colombia. Resolution 2154 – 2012 of the Ministry of Health and Social Protection (25) established the maximum allowed levels of dioxins and furans for fat from cattle and sheep (3.0 pg WHO-TEQ/g fat), fat from pig (1.0 pg WHO-TEQ/g fat), animal fat mixture (2.0 pg WHO-TEQ/g fat) and vegetable oils and fats (0.75 pg WHO-TEQ/g fat). If dl-PCBs are included, the maximum levels are 4.5, 1.5, 3.0 and 1.5 pg WHO-TEQ/g fat, respectively. Table 1 summarizes the valid dioxins regulations in Colombia.

PCDD/Fs Levels in Environmental Matrices in Colombia Emissions Determined Using Toolkits In 2010 a projection and diagnostic of PCDD/Fs emissions in Colombia was obtained using a review and update of the results from the first national inventory of sources and vectors of release of PCDD/Fs in 2002 that was developed following the proposed methodology by the United Nations Environment Programme (UNEP), in the “Standardized Toolkit for Identification and Quantification of Dioxin and Furan Releases” (26). The Toolkit is used for obtaining national inventories of dioxin and furans based on secondary information and emission factors, without the necessity of carrying out expensive analytical procedures. The toolkit includes the following five steps (27): i) generate a matrix to identify the main categories of sources of PCDD/Fs in the country; ii) determine the subcategories for identification of the activities that generate PCDD/Fs; iii) specify information of the processes and activities for characterization in order to classify the identified sources of PCDD/Fs; iv) calculate the release of PCDD/Fs based of the gotten information using the emission factors and the activity in tonne per year; v) summarize the standardized inventory of PCDD/Fs. 53 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Table 1. Valid regulations in Colombia for dioxins maximum permissible levels Maximum permissible levels Matrix

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Resolution

909-2008 (Ministry of Environment, Housing and Territorial Development) (17)

122-2012 (Ministry of Health and Social Protection) (24)

2154-2012 (Ministry of Health and Social Protection) (25)

a

Dioxins and furans

Dioxins, furans and dl-PCBs

Stationary sources of industrial activities

(0.5 ng TEQ/m3)a



Hazardous waste incinerators and cement kilns co-processing waste

(0.1 ng TEQ/m3)a



Non-hazardous waste incinerators

(0.5 ng TEQ/m3)a



Fish and fishery products (fish and shellfish) except eel

4.0 pg WHOTEQ/g fresh weight

8.0 pg WHOTEQ/g fresh weight

Eel

4.0 pg WHOTEQ/g fresh weight

12.0 pg WHOTEQ/g fresh weight



25.0 pg WHOTEQ/g fresh weight

Fish liver and its derived products (except oil from sea organisms) Fat from cattle and sheep

3.0 pg WHOTEQ/g fat

4.5 pg WHOTEQ/g fat

Fat from pig

1.0 pg WHOTEQ/g fat

1.5 pg WHOTEQ/g fat

Animal fat mixture

2.0 pg WHOTEQ/g fat

3.0 pg WHOTEQ/g fat

Vegetable oils and fats

0.75 pg WHOTEQ/g fat

1.5 pg WHOTEQ/g fat

Volume at 25°C and 760 mmHg and reference oxygen content of 11%.

In the reported analysis, the considered categories were: incineration of wastes, production of ferrous and non-ferrous metals, power generation and heating, mineral production, transportation, uncontrolled combustion processes, production and use of chemicals and consumer products, miscellaneous, and waste treatments (27). It was found that the emissions for the country were 945.50 g TEQ/year (gram of equivalent toxic per year) and the order of emissions were: no controlled combustion (46.05%) > incineration of wastes (30.05%) > power generation and heating (7.43%) > production of ferrous and non-ferrous metals (5.01%) > others (5.02%). The uncontrolled combustion activities included the accidental fires and burning of biomass (forest fires), agricultural waste burning and soil preparation. The main vectors of emission were air with 60.67% and 54 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

wastes with 30.32%; the release in soil was 1.93%, in the products 3.51% and in water 2.12%. The authors of the analysis concluded that the emission rate per capita in Colombia (21.02 g TEQ/year) was relatively low compared with countries with similar levels of industrial development (27).

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Reported Concentrations of Several Matrices Table 2 summarizes PCDD/Fs levels reported in some regions of Colombia by several authors in samples of ambient air, particulate matter and samples taken from incinerators (stack gas emissions and fly ashes). Figure 2 shows the localization of the sites where some matrices have been sampled for analysis of PCDD/Fs and dl-PCBs.

Figure 2. Localization map of PCDD/Fs samples analyzed in Colombia (Maps edited from (28, 29)).

55 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Table 2. Concentration of PCDDs, PCDFs, and dl-PCBs in several matrices analysed in Colombia

Stack gas emissions

Year of collection

2003–2005

56

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Matrix

Stack gas emissions

Stack gas emissions

2003–2005

Not reported

Location

No of samples

Concentration

All over Colombia Incinerators of industrial refuse materials and medical residues, without air pollution control devices.

5

All over Colombia Incinerators of industrial refuse materials and medical residues, with air pollution control devices.

7

PCDD/Fs: 0.5 to 39.2 ng I-TEQ/Nm3

Antioquia Hospital batch operated single-chamber incinerators

4

PCCD/Fs: 13.0 - 263.8 ng I-TEQ/Nm3 0.15 - 1.32 ng I-TEQ/Nm3.kg waste

Antioquia. Hospital waste incinerators, batch operated, double-chamber pyrolytic incinerator

4

PCCD/Fs: 27.5 - 708.5 ng I-TEQ/Nm3 0.39 - 8.86 ng I-TEQ/Nm3.kg waste

Antioquia. Hospital waste incinerators, batch operated, double-chamber excess air incinerators

4

PCCD/Fs: 7.2 - 557.8 ng I-TEQ/Nm3 0.19 - 14.12 ng I-TEQ/Nm3.kg waste

Medellín’s Metropolitan Area Incinetator of medical and industrial wastes

4

PCCD/Fs: 1 to 30 ng I-TEQ/Nm3

Ref.

PCDD/Fs: 6.9 to 343.8 ng I-TEQ/Nm3 (30)

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

(31)

(32)

Stack gas emissions

Fly ashes

Fly ashes

Year of collection

Location

Not reported

Tabio. Co-firing process of municipal solid waste in a Hoffmann-type brick kiln in six measuring points

1

PCDD/Fs: 0.35 and 1.29 ng I-TEQ/Nm3

(33)

2003–2005

All over Colombia Incinerators of industrial refuse materials and medical residues, with air pollution control devices.

7

PCDD/Fs: 8.5–67.5 ng I-TEQ/g

(30)

Medellín Collected from the bag filter of a hazardous waste incinerator of mixed medical and industrial residues.a

1

Total PCDD: 57,064.3 pg WHO-TEQ/g Total PCDF: 12,4471.5 pg WHO-TEQ/g Total dl-PCBs: 1481.7 pg WHO-TEQ/g

(34)

Liceo (Manizales)

6

PCDD/Fs: 19 - 52 fg WHO-TEQ/m3 (PM10: 42-54 μg/m3).

Nubia (Manizales)

2

PCDD/Fs: 1 - 3 fg WHO-TEQ/m3 (PM10: 23-26 μg/m3)

Palogrande (Manizales)

6

PCDD/Fs: 4-23 fg WHO-TEQ/m3. (PM10: 27-34 μg/m3).

Not reported

57

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Matrix

Particulate matter, PM10

September 2009 and July 2010

No of samples

Concentration

Ref.

(35)

Continued on next page.

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Matrix

Year of collection

Particulate matter

September 2009 to June 2012

Location

Manizales

Liceo (Manizales)

Sena (Manizales)

58

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Table 2. (Continued). Concentration of PCDDs, PCDFs, and dl-PCBs in several matrices analysed in Colombia

Ambient air

No of samples

22

Total PCDD/Fs: 23 fg WHO-TEQ2005/m3day Total dl-PCB: 0.99 fg WHO-TEQ2005/m3day

7

Total (PCDD/Fs + dl-PCBs): 8.04-12.24 fg WHO2005-TEQ/m3 PCDD/Fs: 7.0 fg WHO2005-TEQ/m3

3

Total (PCDD/Fs + dl-PCBs): 11.23–13.55 fg WHO2005-TEQ/m3 PCDD/Fs: 8.3 fg WHO2005-TEQ/m3

7

Total (PCDD/Fs + dl-PCBs): 4.54–19.54 fg WHO2005-TEQ/m3 PCDD/Fs: 7.7 fg WHO2005-TEQ/m3

7

Total (PCDD/Fs + dl-PCBs): 3.46–8.45 fg WHO2005-TEQ/m3 PCDD/Fs: 3.6 fg WHO2005-TEQ/m3

From June 2012 to November 2014 Nubia (Manizales)

Palogrande (Manizales)

Concentration

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Ref.

(36)

(37)

Location

Year of collection

Ambient air

1

1

Total PCDD/Fs: 32 fg WHO-TEQ2005/m3-day Total dl-PCB: 11 fg WHO-TEQ2005/m3-day

1

Total PCDD/Fs: 39 fg WHO-TEQ2005/m3-day Total dl-PCB: 25 fg WHO-TEQ2005/m3-day

Palogrande (Manizales)

1

Total PCDD/Fs: 18 fg WHO-TEQ2005/m3-day, Total dl-PCB: 7 fg WHO-TEQ2005/m3-day

Manizales

4

PCDD/Fs: 3.54-7.37 fg I-TEQ/m3

Sena (Manizales)

La Nubia (Manizales)

Ambient air

January 2011 and December 2012.

Concentration Total PCDD/Fs: 27 fg WHO-TEQ2005/m3-day Total dl-PCB: 12 fg WHO-TEQ2005/m3-day

Liceo (Manizales)

JuneOctober 2012 (between 101 and 106 days)

No of samples

59

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Matrix

Ref.

(36)

(38) Continued on next page.

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Matrix

Year of collection

Location

Ambient air

From December 2013 to November 2014

Bogotá

Ambient air

January 2011 and December 2012

Arauca

No of samples

Concentration

Ref.

3

Total (PCDD/Fs + dl-PCBs): 30.75–43.42 fg WHO2005-TEQ/m3 PCDD/Fs: 26.5 fg WHO2005-TEQ/m3

(37)

4

PCDD/Fs: 2.36-3.46 fg I-TEQ/m3

(38)

a

The incinerator was equipped with a gas cooling heat exchanger, a wet acid scrubber, a cyclone and a bag filter for particulate collection. Finally, the system was equipped with a fixed bed of activated carbon for dioxin adsorption.

60

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Table 2. (Continued). Concentration of PCDDs, PCDFs, and dl-PCBs in several matrices analysed in Colombia

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Emissions from the Incineration Sector in Colombia In Colombia the hazardous waste from industrial and medical activities is mainly thermally treated, whereas most MSW are commonly landfill disposed. The number of incinerators in Colombia in 2007 was about 170, 57% were located in hospitals, 37% in private companies for burning their own waste and 11% corresponded to commercial incinerators servicing third parties (39). In most of the incinerators, there was not a characterization of waste streams and the industrial and medical wastes were incinerated together. Most plants operating in Colombia had capacities below 100 kg/h and the estimated total installed capacity was 18,000 ton/year (30, 32). If incinerators are not properly designed and operated, when medical wastes are incinerated, several pollutants are emitted: particulate matter, acidic gases, trace metals, products of incomplete combustion and polynuclear organic matter as PCDD/Fs that are emitted in exhaust gases and incineration ashes (40, 41).

Gases Monitoring of PCDD/Fs emissions from stack gas and fly ash samples of twelve plants from Antioquia province (located in Colombian northwestern zone) were sampled during a two-year period, 2003-2005 (31). In the incinerators that did not use air pollution control devices (capacities lower than 100 kg/h), dioxin and furan concentrations varied significantly from 6.9 to 343.8 ng I-TEQ/Nm3. For the incinerators equipped with at least one conventional air pollutant control system, dioxin emissions ranged from 0.5 to 39.2 ng I-TEQ/Nm3, the concentration was still high because the control systems used (electrostatic precipitator, cyclone or scrubber) were more suitable for controlling dust and acidic gases than PCDD/Fs emissions; 8.5–67.5 ng I-TEQ/g were measured in fly ash samples (30). The highest PCDD/Fs emissions were found for those incinerators that burned medical waste mixed with industrial/chemical waste (30). Dioxin concentrations from monitored medical waste incinerators were in the range from about 7 to 700 ng I-TEQ/Nm3; dioxin concentrations from most incinerators increased with emissions of total suspended particulate (TSP) (31). It was reported that the average dioxin emissions depended on the incinerator type; the PCDD/Fs emissions increased in the following order: double-chamber pyrolytic (79.0 ng I-TEQ/Nm3) < double-chamber excess air (187.1 ng I-TEQ/Nm3) < single-chamber incinerators (235.8 ng I-TEQ/Nm3); the authors concluded that the obtained sequence might be due to improperly operated incinerators (31). The lowest dioxin concentration, around 1 ng I-TEQ/Nm3, was obtained for a plant equipped with air pollution control devices (heat exchanger, scrubber, cyclone, bag filter and active carbon bed). The measured dioxin concentrations from several waste incinerators with capacities of 345, 1037, 1653 and 345 ton/year with volumetric flows of 2378, 4907, 5645 and 6912 Nm3/year, respectively, were lower than the values found using the Toolkit (2.0 × 10–6 vs 3.45 × 10–3, 0.8 × 10–4 vs 3.11, 1 × 10–4 vs 5, 2.1 × 10–4 vs 13.8 g TEQ/year) (32).

61 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Fly Ashes In Colombia, fly ash from waste incinerators has been mixed with household waste and landfille without either previous treatment or use of appropriate human protection (34). Fly ash contains high concentrations of dioxins, heavy metals and soluble salts (42); and the content of dioxins is large for particulate matter that is formed in long residence time devices like bag filters and electrostatic precipitators (43, 44). Fly ash samples from bag filter from a hazardous (industrial and medical) waste incinerator located in Medellín were analysed. It was concluded that the high concentration of PCDD/Fs (181,535.8 pg WHO-TEQ/g) was owing to the inefficient combustion operation (batch process, outdated furnace design, slow gas cooling system) and the composition of the waste input (medical–industrial waste mixtures) (34). Fly ash samples from several control systems of plants which incinerated industrial refuse materials and medical residues were also analysed. PCDD/Fs content of fly ashes from bag filter, cyclone and electrostatic precipitator was 67.5, 27.0 and 8.5 ng I-TEQ/g, respectively (30).

Emissions from Co-Firing Process The stack emissions analysis during the co-firing process of 2 tonne municipal solid waste (MSW) from the town of Tabio, Colombia, in a Hoffmann-type brick kiln that processed about 18.5 tonnes of clay was reported. 15 samples of collected MSW, between February 2004 and June 2005, contained organic matter (53.5%), plastics & rubber (12.5%), textiles (6.5%), paper and cardboard (24.5%), metals (0.5%), glass and ceramics (2.5%). The concentration of PCDD/Fs in six measuring points of the Hoffmann-type brick kiln, in three different tests, varied between 0.26 and 1.35 ng I-TEQ /Nm3 with a mean value of 0.74 ng I-TEQ /Nm3 (33). At the date of the publication, the emission limit values for incineration plants in Colombia and European Union (22, 45) were, respectively, 10 and 0.1 ng I-TEQ/Nm3. As it was found that the studied process fulfilled the environmental standards established in Colombia for the incineration systems through Resolution 0886 of the Ministry of Environment Housing and Territorial Development for PCDD/Fs and other contaminants, then the co-fired incineration process implemented in a Hoffmann-type brick kiln was suggested as a MSW disposal process in Colombia (22).

Emissions Determined from Stations in Several Cities of Colombia Manizales It is located on the western slopes of the Colombian central mountain range (part of the longest continental mountain range of the Andes) at 2150 m above sea level. Its urban density is approximately 6800 inhabitants/km2 (37). Manizales 62 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

is impacted by emissions from an industrial zone that includes a coal-fired metal foundry recycling plant, food processing plants, plastic processing industries and a municipal solid waste incinerator. There are areas of high vehicular density (310 vehicles per 1000 inhabitants) (46), located mainly downtown; furthermore, the area around Manizales includes the active Nevado del Ruiz volcano, 28 km to the southeast, whose daily emissions influence the atmospheric chemistry of the city and neighboring towns (37). PCDD/Fs were measured in areas with different vehicular density and in residential zones from different stations in Manizales. Higher concentrations were found in high vehicular density areas (Liceo station: total mean concentration of 151 fg/m3 and 7.0 fg WHO2005-TEQ/m3), than in residential areas (Palogrande station: total mean concentration of 64 fg/m3 and 3.6 fg WHO2005-TEQ/m3). Intermediate concentration was observed in the industrial influence areas (Nubia and SENA stations: total mean concentration of 100 fg/m3) (37)

Bogotá It is a megacity located in the plateau of eastern Cordillera of the Andes at 2600 m above sea level. With a population of 9.56 million, it is one of the principal cities in Latin America (47). The ambient air of this region is characterized by constant high vehicular activity (294 vehicles per 1000 inhabitants (46)) and the presence of a wide variety of industrial processes in different areas of the city. The Fontibón station, at the west of Bogotá, was located as an urban/industrial-commercial zone with high vehicular traffic and manufacturing industrial activities (37). The total mean concentration of PCDD/Fs was 373 fg/m3 and 26.5 fg WHO2005-TEQ/m3 with a total concentration of PCDD/Fs and dl-PCBs ranging from 30.75 to 43.42 fg WHO2005-TEQ/m3. High dioxin concentrations were found in Bogotá samples throughout the range of congeners (37). PCDD/Fs levels obtained in Bogotá, Manizales and Arauca were compared with reports from the GAPS network in the GRULAC zone and other zones in Europe (38). Bogotá showed results higher than the urban area of Quito, Ecuador (mean Σ4-8PCDD/Fs = 223 fg/m3, where Σ4-8PCDD/Fs is the sum of all homologue groups). Bogotá (mean Σ4-8PCDD/Fs = 373 fg/m3), Manizales (mean Σ4-8PCDD/ Fs = 151 fg/m3 in the zone of high vehicular influence) and Arauca (mean Σ48PCDD/Fs = 167 fg/m3) reported levels much lower than the urban zones of Sao Paulo, Brazil (meanΣ4-8PCDD/Fs = 1580 fg/m3), and São Luis, Brazil (i.e., 2560 fg/m3) and the agricultural region of Sonora in México (mean Σ4-8PCDD/Fs = 1310 fg/m3), but higher that Tapanti (Costa Rica, BA) (10.8 fg/m3) (38).

PCDD/Fs Occurrence in Some Foods Consumed in Colombia The three most important pathways for human contamination by dioxins are: a) deposition of vapors and particles in plants that are consumed by animals or humans; b) pollution of the soil with the chemicals, transfer to plants by root 63 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

uptake or by volatilization from soil and deposition in plants that are consumed by animals or humans; and c) ingestion of the contaminated soil by animals (48). However, human exposure to dioxin and furans is mainly (90-98%) through daily consumption of contaminated foods with PCDD/Fs and related compounds. In recent years, a number of studies have focused on determining the levels of PCDD/Fs and related compounds in various countries around the world. For instance, in Catalonia, Spain the levels of PCDD/Fs in foodstuffs of high consumption were determined in 2000 (49) and 2008 (50) in order to estimate the dietary intake of dioxins by the population. The results indicated that the main contribution to the dietary intake of PCDD/Fs is given by fish and seafood followed by dairy products, cereals, meat and oils/fats, while the lowest contributions were from vegetables and fruits. As an example of this, the highest contribution to the dietary intake of PCDD/Fs for a standard male adult of 70 kg body weight from Catalonia corresponded to fish and seafood (28.0%), followed by dairy products (15.4%), while the lowest contributions corresponded to tubers (1.1%) followed by vegetables (2.7%) and fruits (3.7%) (50). A similar study was developed in 2005 in Sweden, where food products with high fat content and of animal origin were analysed. The calculated contributions of the different food groups to the per capita intake of PCDD/F and dl-PCB showed that fish was a major contributor (49%) followed by dairy products (22%), meat/meat products (15%) and fats/oils (13%) (51). In Colombia there are few studies related with the PCDD/Fs content in food. To our knowledge, just one study about the content of those compounds in foods produced and consumed in Colombia has been reported (52). Recently, soybean oil, butter and shrimp have shown an increase in Colombian consumption. In 2001, an average of 100,000 tons of refined soybean oil and 130,000 tons of fat were consumed in Colombia and the butter exhibited an increase in population consumption of around 6% during 2008-2012 (52, 53). Also, during 2011 and 2012 the shrimp industry reported a significant increase (37%) in Colombia (54). The important consumption and the high contribution in dietary intake of PCDD/Fs in these kind of food, makes them an interesting food group for studying the levels of PCDD/Fs. Besides, as it was mentioned in the section of regulatory history, the maximum levels for dioxins in oils, fats and fishery products are regulated in Colombia. The levels of PCDD/Fs and dl-PCBs in soybean oil, fish oil, butter and shrimp produced and consumed in Colombia were reported (52, 55–57). In most oil samples, 2,3,7,8-substituted congeners with a high-chlorination degree were predominant, especially OCDD followed by 1234678-HpCDF, 1234789-HpCDF and OCDF (Figure 3). For soybean oil, butter and shrimp samples, the OCDD was the predominant congener compared to the others compounds and those results agree with the trend that has been reported for the contribution of PCDD/Fs in foods (49, 50, 52). Shrimp samples showed the highest levels of PCDD/Fs (44.8 pg/g) followed by butter (12.90 pg/g) and fish oil (7.72 pg/g), whereas soybean oil sample showed the lowest concentration (4.71 pg/g) (Table 3) (52).

64 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Table 3. Total concentrations and WHO-TEQ values for PCDD/Fs and dl-PCBs for food samples produced in colombia and commercialized in Medellín-Colombia. Adapted from reference (52), Copyright (2016), with permission from Elsevier. Compound Σ PCDD/Fsa WHO-TEQ Σ

PCBsa

WHO-TEQ Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

(PCDD/Fs)b

(PCBs)b

WHO-TEQ ( PCDD/Fs +

PCBs)b

Soybean oil

Fish oil

Butter

Shrimp

4.71

7.72

12.90

44.80

0.24

1.40

1.03

1.71

109.50

5493.40

92.30

193.50

0.12

1.00

0.05

0.24

2.40

1.08

1.95

0.36

a

pg/g of fat for oil/fat and dry weight for shrimp dry weight for shrimp

b

pg WHO-TEQ/g of fat for oil/fat and

Figure 3. Individual concentrations of PCDD/Fs congeners in samples (pg/g of fat for oil/fat or dry weigh for shrimp). (data from (52)).

65 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Shrimp and fish oil presented the highest PCDD/Fs WHO-TEQ values (1.71 and 1.40 pg WHO-TEQ/g, respectively) followed by butter (1.03 pg WHO-TEQ/g), while soybean oil showed the lowest levels (0.24 pg WHO-TEQ/g). Similarly, fish oil (2.40 pg WHO-TEQ/g) and shrimp (1.95 pg WHO-TEQ/g) samples exhibited the highest levels of total WHO-TEQ (PCDD/Fs and dl-PCBs) and vegetable oil presented the lowest levels (0.36 pg WHO-TEQ/g). Butter sample exhibited intermediate values (1.08 pg WHO-TEQ/g) (52). Figure 4 shows the contribution of PCDFs and PCDDs to the total concentration for each sample. Butter (76.7%) and shrimp (82.2%) samples exhibited the highest PCDD contribution to the total content. Regarding to PCDF contribution, soybean oil and fish oil showed the highest values corresponding to 61.9 and 66.9 %, respectively (52).

Figure 4. PCDD and PCDF contribution to the total concentration of foods produced and consumed in Colombia. (Adapted from reference (52), Copyright (2016), with permission from Elsevier)

Furthermore, the food samples study (52) allowed to assess whether the levels of PCDD/Fs and dl-PCBs in oils, fat and shrimp exceeded the maximum values permitted for Colombian regulations (Table 1). For soybean oil and butter samples PCDD/Fs + dl-PCBs levels were below the limits established by the Colombian legislation (0.75 pg WHO-TEQ/g for PCDD/Fs and 1.5 pg WHO-TEQ/g for PCDD/Fs and dl-PCBs) (25), whereas PCDD/Fs content in butter (1.03 pg WHO-TEQ/g) showed concentrations above the maximum levels established. Because Colombian legislation does not have stablished maximum levels for fish oils, the results for this kind of products were compared with the European regulation (1.75 pg WHO-TEQ/g for PCDD/Fs and 6 pg WHO-TEQ/g for PCDD/Fs and dl-PCBs) (58) and it was found that fish oil 66 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

total concentration was below the maximum concentration levels established. In addition, the PCDD/Fs levels found for soybean oil were below threshold established in European regulation for vegetable oils and fats (vegetable oil: 0.75 pg WHO-TEQ/g for PCDD/Fs) (58). On the other hand, shrimp sample exhibited values below the maximum concentration levels established by the EU legislation (3.5 pg WHO-TEQ/g of dry weight for PCDD/Fs and 6.5 pg WHO-TEQ/g of dry weight for PCDD/Fs and dl-PCBs) (58) and Colombian regulation (4 pg WHO-TEQ/g of dry weight for PCDD/F and 8 pg WHO-TEQ/g of dry weight for PCDD/Fs and dl-PCBs) (25).

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

Conclusions In Colombia, the regulations related with the emissions and content of dioxin and furans in some matrices is new, just since 2001 the government is regulating those items. It is necessary that all the activities that generate large amounts of PCDD/Fs are monitored and regulated. In order to reduce the PCDD/Fs generation and emissions, incinerators in Colombia must improve their processes, classifying the wastes to incinerate, controlling the inceneration conditions and installing or adequating air pollution control devices for the treatment of combustion gases. As large differences are obtained between values obtained with emission factors and the measured concentrations, and only some locations in Colombia have been tested; the number of studies for measuring the concentration of PCDD/Fs in several matrices and in several places from Colombia must increase for getting a better understanding of the PCDD/Fs problem in the country. The main cities that have been characterized are just Medellín and Manizales, and in some extent Bogotá and Arauca. The results from the preliminary research of PCDD/Fs and dl-PCBs levels in several types of foods that are consumed in Colombia indicated that there are not possible health risks owing to PCDD/Fs. However, it is important to implement monitoring programs or encourage related research to assess dioxin and furan contents in other foods consumed in Colombia and to increase the number of analysis of the products currently studied.

Acknowledgments The authors thank finantial support from Universidad de Antioquia, UdeA.

References 1. 2.

Wania, F.; MacKay, D. Peer Reviewed: Tracking the Distribution of Persistent Organic Pollutants. Environ. Sci. Technol. 1996, 30, 390A–396A. UNEP. Stockholm Convention on Persistent Organic Pollutants (POPs); Stockholm Convention Website, 2009 [Online]. http://chm.pops.int/ Portals/0/download.aspx?d=UNEP-POPS-COP-CONVTEXT-2009.En.pdf (accessed August 30, 2016). 67

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

3.

4.

5.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

6.

7.

8.

9.

10.

11.

12. 13.

14.

15.

Iwata, H.; Tanabe, S.; Sakai, N.; Tatsukawa, R. Distribution of Persistent Organochlorines in the Oceanic Air and Surface Seawater and the Role of Ocean on Their Global Transport and Fate. Environ. Sci. Technol. 1993, 27, 1080–1098. Hutzinger, O.; Choudhry, G. G.; Chittim, B. G.; Johnston, L. E. Formation of Polychlorinated Dibenzofurans and Dioxins during Combustion, Electrical Equipment Fires and PCB Incineration. Environ. Health Perspect. 1985, 60, 3–9. Hryhorczuk, D. O.; Orris, P.; Kominsky, J. Ŕ.; Melius, J.; Burton, W.; Hinkamp, D. L. PCB, PCDF, PCDD Exposure Following a Transformer fire—Chicago. Chemosphere 1986, 15, 1297–1303. Agency for Toxic Substances and Disease Registry (ATSDR). Toxicological Report for Pentachlorophenol (Update). Publication No. PB/2001/109106/ AS; U.S. Department of Health and Human Services: Atlanta, Georgia, 2001. Masunaga, S.; Takasuga, T.; Nakanishi, J. Dioxin and Dioxin like Impurities in Some Japanese Agrochemical Formulations. Chemosphere 2001, 44, 873–885. Loganathan, B. G.; Kannan, K.; Sajwan, K. S.; Chetty, C. S.; Giesy, J. P.; Owen, D. A. Polychlorinated Dibenzo-P-Dioxins, Dibenzofurans and Polychlorinated Biphenyls in Street Dusts and Soil Samples from Savannah, Georgia. Organohalogen Compd. 1997, 32, 192–197. Fiedler, H. National PCDD/PCDF Release Inventories under the Stockholm Convention on Persistent Organic Pollutants. Chemosphere 2007, 67, S96–S108. Ullmann’s Encyclopedia of Industrial Chemistry, 5th ed.; Evers, B., Hawkins, S., Ravenscroft, M., Rounsaville, J. F., Schultz, G., Eds.; VCH: Weinheim, Germany, 1989. Rappe, C. In Dioxin: Toxicological and Chemical Aspects; Cattabeni, F., Cavallaro, A., Galli, G., Eds.; SP Medical and Scientific Books: New York, 1978; pp 9–11. Loganathan, B. G.; Kannan, K. Global Organochlorine Contamination Trends: An Overview. Ambio 1994, 23 (3), 187–191. Safe, S. Polychlorinated Biphenyls (PCBs), Dibenzo-P-Dioxins (PCDDs),dibenzofurans (PCDFs) and Related Compounds: Environmental and Mechanistic Considerations Which Support the Development of Toxic Equivalency Factors (TEFs). Crit. Rev. Toxicol. 1990, 21, 51–88. Loganathan, B. G. In Global Contamination Trends of Persistent Organic Chemicals; Loganathan, B. G., Lam, P. K.-S., Eds.; CRC Press (Taylor and Francis Group): Boca Raton, FL, 2012. Colombian Institute of Hydrology Meteorology and Environmental Studies (IDEAM); Colombian Ministry of Environment; SINA. Colombia. First National Communication to the United Nations Framework Convention on Climate Change [Online]; United Nations Framework Convention on Climate Change. http://http://unfccc.int/cop8/se/se_pres/presnc01col.pdf (accessed August 30, 2016).

68 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

16. World Population Review. Colombia Population 2016; http://worldpopulationreview.com/countries/colombia-population/ (accessed June 25, 2016). 17. Colombian Ministry of Environment Housing and Territorial Development. Resolution 909 of June 5th 2008 (in Spanish); Bogotá, D.C, Colombia, 2008; p 36. http://www.minambiente.gov.co/index.php/normativa/resoluciones (accessed September 6, 2016) 18. Colombian Ministry of Environment. Resolution 970 of October 30th 2001 (in Spanish); Bogotá, D.C, Colombia, 2001; p 7. http://www.icbf.gov.co/ cargues/avance/docs/resolucion_minambiente_rma97001.htm (accessed September 6, 2016). 19. Colombian Ministry of Environment. Resolution 458 of May 27th 2002 (in Spanish); Bogotá, D.C, Colombia, 2002; p 8. http://www.icbf.gov.co/ cargues/avance/docs/resolucion_minambiente_0458_2002.htm (accessed September 6, 2016). 20. Colombian Ministry of Environment Housing and Territorial Development. Resolution 1488 of December 19th 2003 (in Spanish); Bogotá, D.C, Colombia, 2003; p 6. http://web2006.minambiente.gov.co:8091/ fichaArchivo.aspx?id=95 (accessed September 6, 2016). 21. Colombian Ministry of Environment. Resolution 0058 of January 21st 2002 (in Spanish); Bogotá, D.C, Colombia, 2002; p 20. http:// www.minambiente.gov.co/index.php/normativa/resoluciones (accessed September 6, 2016). 22. Colombian Ministry of Environment Housing and Territorial Development. Resolution 0886 of July 27th 2004 (in Spanish); Bogotá, D.C, Colombia, 2004; p 16. http://www.minambiente.gov.co/index.php/normativa/ resoluciones (accessed September 6, 2016). 23. Colombian Ministry of Social Protection. Resolution 776 of March 6th 2008 (in Spanish); Bogotá, D.C, Colombia, 2008; p 9. https://www.invima.gov.co/ images/stories/resoluciones/resolucion_776_2008.pdf (accessed September 6, 2016). 24. Colombian Ministry of Health and Social Protection. Resolution 122 of January 26th 2012 (in Spanish); Bogotá, D.C, Colombia, 2012; p 8. https://www.invima.gov.co/images/pdf/normatividad/alimentos/ resoluciones/resoluciones/2012/4R_122_de_2012_Pesca.pdf (accessed September 6, 2016). 25. Colombian Ministry of Health and Social Protection. Resolution 2154 of August 2nd 2012 (in Spanish); Bogotá, D.C, Colombia, 2012; p 32. https://www.invima.gov.co/images/pdf/normatividad/alimentos/ resoluciones/resoluciones/2012/r 2154 de 2012 a y g pdf.pdf (accessed September 6, 2016). 26. United Nations Environment Programme (UNEP). Standardized Toolkit for Identification and Quantification of Dioxin and Furan Releases; UNEP: Geneva, Switzerland, 2003. 27. García Ubaque, C. A.; García Ubaque, J. C.; Vaca Bohórquez, M. L. Dioxins and Furans Emissions in Colombia: Assessment and Diagnosis (in Spanish). Tecnura 2012, 16, 194–206. 69 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

28. Map of Colombia. http://d-maps.com/carte.php?num_car=22481&lang=es (accessed Jul 6, 2016). 29. Map of America. http://d-maps.com/carte.php?num_car=1425&lang=es (accessed Jul 6, 2016). 30. Aristizábal, B.; Cobo, M.; Hoyos, A.; Montes de Correa, C.; Abalos, M.; Martínez, K.; Abad, E.; Rivera, J. Baseline Levels of Dioxin and Furan Emissions from Waste Thermal Treatment in Colombia. Chemosphere 2008, 73, S171–S175. 31. Hoyos, A.; Cobo, M.; Aristizábal, B.; Córdoba, F.; Montes de Correa, C. Total Suspended Particulate (TSP), Polychlorinated Dibenzodioxin (PCDD) and Polychlorinated Dibenzofuran (PCDF) Emissions from Medical Waste Incinerators in Antioquia, Colombia. Chemosphere 2008, 73, S137–S142. 32. Aristizábal, B.; Cobo, M.; Montes de Correa, C.; Martínez, K.; Abad, E.; Rivera, J. Dioxin Emissions from Thermal Waste Management in Medellín, Colombia: Present Regulation Status and Preliminary Results. Waste Manage. 2007, 27, 1603–1610. 33. García Ubaque, C. A.; Gonzales Hässig, A.; Acosta Mendoza, C. Stack Emissions Tests in a Brick Manufacturing Hoffmann Kiln: Firing of Municipal Solid Waste. Waste Manag. Res. 2010, 28, 596–608. 34. Cobo, M.; Gálvez, A.; Conesa, J. A.; Montes de Correa, C. Characterization of FLy Ash from a Hazardous Waste Incinerator in Medellin, Colombi. J. Hazard. Mater. 2009, 168, 1223–1232. 35. Aristizábal, B. H.; Gonzalez, C. M.; Morales, L.; Abalos, M.; Abad, E. Polychlorinated Dibenzo-P-Dioxin and Dibenzofuran in Urban Air of an Andean City. Chemosphere 2011, 85, 170–178. 36. Cortés, J.; González, C.; Morales, L.; Abalos, M.; Abad, E.; Aristizábal, B. Levels of PCDD/PCDFs and Dl-PCBs in Ambient Air of Manizales Using Passive and Active Samplers. Organohalogen Compd. 2013, 75, 787–791. 37. Cortés, J.; Cobo, M.; González, C. M.; Gómez, C. D.; Abalos, M.; Aristizábal, B. H. Environmental Variation of PCDD/Fs and Dl-PCBs in Two Tropical Andean Colombian Cities Using Passive Samplers. Sci. Total Environ. 2016, 568, 614–623. 38. Schuster, J. K.; Harner, T.; Fillmann, G.; Ahrens, L.; Altamirano, J. C.; Aristizábal, B.; Bastos, W.; Castillo, L. E.; Cortés, J.; Fentanes, O.; Gusev, A.; Hernandez, M.; Ibarra, M. V.; Lana, N. B.; Lee, S. C.; Martínez, A. P.; Miglioranza, K. S. B.; Puerta, A. P.; Segovia, F.; Siu, M.; Tominaga, M. Y. Assessing Polychlorinated Dibenzo- P -Dioxins and Polychlorinated Dibenzofurans in Air across Latin American Countries Using Polyurethane Foam Disk Passive Air Samplers. Environ. Sci. Technol. 2015, 49, 3680–3686. 39. Colombian Ministry of Environment Housing and Territorial Development. Environmental Policy for the Integral Management of Waste or Hazardous Waste. Draft Document for Public Consultation (in Spanish); Bogotá, Colombia, 2005. http://www.ingenieroambiental.com/4014/politicaamb.pdf (accessed September 6, 2016). 40. Colombian Ministry of Environment. Decree 948 of June 5th 1995 (in Spanish); Bogotá, D.C, Colombia, 1995; p 57. 70 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

41.

42.

43.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

44. 45.

46. 47.

48.

49.

50.

51.

52.

53.

http://www.minambiente.gov.co/images/normativa/app/decretos/54dec_0948_1995.pdf (accessed September 6, 2016). Colombian Ministry of Environment; Colombian Ministry of Health. Decree 1669 of August 2nd 2002 (in Spanish); Bogotá, D.C, Colombia, 2002; p 4. http://www.bogotaturismo.gov.co/sites/intranet.bogotaturismo.gov.co/files/ DECRETO 1669 DE 2002.pdf (accessed September 6, 2016). Zhao, L.; Zhang, F.-S.; Wang, K.; Zhu, J. Chemical Properties of Heavy Metals in Typical Hospital Waste Incinerator Ashes in China. Waste Manage. 2009, 29, 1114–1121. Stanmore, B. R. The Formation of Dioxins in Combustion Systems. Combust. Flame 2004, 136, 398–427. Chang, M.-B.; Chung, Y.-T. Dioxin Contents in Fly Ashes of MSW Incineration in Taiwan. Chemosphere 1998, 36, 1959–1968. The European Parliament and the Council of the European Union. Directive 2000/76/EC of 4 December 2000 on the Incineration of Waste. Annex V. Off. J. Eur. Communities 2000, L332, 109–110. Manizales Cómo Vamos. Quality of Life Report 2013 (in Spanish); http:// manizalescomovamos.org/?page_id=2278 (accessed Jun 6, 2016). United Nations. Department of Economic and Social Affairs. Population Division. World Urbanization Prospects. The 2014 revision http:// esa.un.org/unpd/wup/DataQuery/ (accessed Jun 15, 2016). Fries, G. F. A Review of the Significance of Animal Food Products as Potential Pathways of Human Exposures to Dioxins. J. Anim. Sci. 1995, 73, 1639–1650. Llobet, J. M.; Domingo, J. L.; Bocio, A.; Casas, C.; Teixidó, A.; Müller, L. Human Exposure to Dioxins through the Diet in Catalonia, Spain: Carcinogenic and Non-Carcinogenic Risk. Chemosphere 2003, 50, 1193–1200. Perelló, G.; Gómez-Catalán, J.; Castell, V.; Llobet, J. M.; Domingo, J. L. Assessment of the Temporal Trend of the Dietary Exposure to PCDD/Fs and PCBs in Catalonia, over Spain: Health Risks. Food Chem. Toxicol. 2012, 50, 399–408. Törnkvist, A.; Glynn, A.; Aune, M.; Darnerud, P. O.; Ankarberg, E. H. PCDD/F, PCB, PBDE, HBCD and Chlorinated Pesticides in a Swedish Market Basket from 2005 – Levels and Dietary Intake Estimations. Chemosphere 2011, 83, 193–199. Pemberthy, D.; Quintero, A.; Martrat, M. G.; Parera, J.; Ábalos, M.; Abad, E.; Villa, A. L. Polychlorinated Dibenzo-P-Dioxins, Dibenzofurans and Dioxin-like PCBs in Commercialized Food Products from Colombia. Sci. Total Environ. 2016, 568, 1185–1191. Espinal G, C. F.; Martínez C, H. J.; Soler, M. S.; Barrios Urrutia, C. A. Ministry of Agriculture and Rural Development Observatory Competitiveness- Agro-Chains. The Chain of Oil in Colombia (in Spanish); 2001 [Online] http://www.siame.gov.co/siame/documentos/documentacion/ mdl/03_VF_Bibliografia/Biodiesel/Aceites%20de%20palma.Col_0203.pdf (accessed September 6, 2016). 71

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch003

54. Instituto Colombiano Agropecuario. The Colombian Shrimp Culture Sector: Evolution and Admissibility (in Spanish); Cartagena, Colombia 2012; pp 1–34. 55. Pemberthy, D.; Quintero, A.; Mg, M.; Parera, J.; Ábalos, M.; Abad, E.; Montes De Correa, C. Dioxins and Furans in Vegetable Oils Sold in Colombia. Organohalogen Compd. 2011, 73, 813–816. 56. Pemberthy, D.; Quintero, A.; Martrat, M. G.; Ábalos, M.; Abad, E.; Villa, A. Levels of PCDD/PCDFs and Dl-PCBs in Food Commercial Samples Quantified by HRGC-HRMS. Organohalogen Compd. 2014, 76, 1525–1528. 57. Pemberthy, D.; Quintero, A.; Martrat, M. G.; Parera, J.; Abad, E.; Villa, A. Levels of Polychlorinated Dibenzo-P-Dioxins, Polychlorinated Dibenzofurans and Dioxin-like PCBs in Oils Commercialized in Colombia. Organohalogen Compd. 2015, 77, 736–739. 58. European Commission. Regulation (EU) No 1259/2011 of 2 December 2011. Off. J. Eur. Union 2011, L 320, 18–23.

72 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Chapter 4

Continuous Monitoring of Persistent Organic Pollutants in China for the Effectiveness Evaluation of the Stockholm Convention: 2007–2014 Lirong Gao,1 Minghui Zheng,*,1 Yibing Lv,2 Qiang Fu,2 Li Tan,2 and Qingqing Zhu1 1Research

Center For Eco-Environmental Sciences, Chinese Academy of Sciences, No. 18 Shuangqing Road, Haidian District, Beijing 100085, China 2National Environmental Monitoring Center of China, No. 8 Anwai Dayangfang Road, Chaoyang District, Beijing 100012, China *E-mail: [email protected]

Article 16 of the Stockholm Convention requires the Conference of the Parties to periodically evaluate whether the Convention is an effective tool for achieving the objective of protecting human health and the environment from persistent organic pollutants (POPs). A global monitoring plan for POPs, which has been put in place under the Convention, is a key component of the effectiveness evaluation that provides a harmonized framework for identifying changes in POP concentrations over time and information on the regional and global transport of POPs in the environment. To meet the requirements of the effectiveness evaluation of the Stockholm Convention, we monitored concentrations of polychlorinated dibenzo-p-dioxins and furans (PCDD/Fs) and polychlorinated biphenyls (PCBs), including dioxin-like (DL-PCBs) and indicator PCBs, in air in China from 2007 and 2011. We also monitored perfluorooctane sulfonate (PFOS) and perfluorooctanoic acid (PFOA) in water in 2013. We collected air from 11 remote sites in China, 3 © 2016 American Chemical Society Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

urban sites, and 3 rural sites using high volume samplers with a size-selective inlet that only accepted particles with diameters less than 10 µm. The remote air sampling sites were referred to as background sites because the PCDD/F and DL-PCB concentrations were typical of background levels and ranged from 2.64 to 101.7 World Health Organization toxic equivalents (WHO-TEQ) fg/m3 (average 21.7 WHO-TEQ fg/m3). The concentrations of PCDD/F were highest in eastern China, where there is intensive economic activity. The PCDD/F and DL-PCB concentrations in the air samples from the 3 urban and 3 rural sites ranged from 98.1 to 212.2 WHO-TEQ fg/m3 and from 11.2 to 46.4 WHO-TEQ fg/m3, respectively. The average PCDD/Fs concentrations were much higher at urban sites than at the background and rural sites. The indicator PCB concentrations in the air samples from the 11 background sites, 3 urban sites, and 3 rural sites ranged from 5.2 to 44.9 pg/m3 (average 18.5 pg/m3), from 55.1 to 90.6 pg/m3 (average 71.6 pg/m3), and from 4.62 to 10.3 pg/m3 (average 7.32 pg/m3), respectively. The levels of PCDD/Fs and PCBs in the air samples from the background sites were lower than, or comparable to, those detected in the air at other sites worldwide. The concentrations of PFOA and PFOS were monitored in two lakes and two coastal marine areas. The average concentration of PFOS and PFOA in Taihu Lake was 32 ng/L. The concentrations of PFOA and PFOA at the other sites were below the detection limit. There were no consistent temporal trends in the POP concentrations in the air samples from the background sites.

Introduction Background and Objectives Persistent organic pollutants (POPs) are chemicals that remain intact in the environment for long periods of time. They are mobile and can be transported over long distances in the global environment. They bioaccumulate in fatty tissues of living organisms, and are harmful for humans and wildlife. The Stockholm Convention on POPs, hereafter referred to as the Stockholm Convention, is an international treaty that aims to protect human health and the environment from the potential toxic effects of POPs. Twenty-six POPs are listed in annexes to the Stockholm Convention. The initial list included polychlorinated dibenzo-p-dioxins and dibenzofurans (PCDD/Fs) and polychlorinated biphenyls (PCBs). As unintentionally produced compounds, PCDD/Fs form mainly through processes such as combustion, production of chlorinated compounds, metal smelting, paper and pulp production, and petroleum refining. Polychlorinated biphenyls are thermally and chemically stable, and have electrical insulating properties. These properties mean that they have been used extensively in industrial applications, such as in the dielectric fluids of transformers and 74 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

capacitors, and in heat-transfer and hydraulic fluids. As well as having industrial applications, PCBs also form unintentionally in industrial processes. Both PCDD/Fs and PCBs are highly toxic to humans and in ecosystems, and are able to persist in various environmental compartments. Atmospheric PCDD/F and PCB concentrations have been examined recently in urban areas in China to evaluate their impact on local residents. Few studies have established background concentrations of PCDD/Fs and PCBs in remote areas in China. Perfluorooctane sulfonyl fluoride substances (PFASs) were added to the Stockholm Convention in 2009. Perfluoroalkyl substances (PFASs) are persistent global pollutants that have been used in numerous industrial processes and consumer products for more than 60 years. PFASs can be used to provide stain-resistant coatings on surfaces; they are also used in firefighting foams, cosmetics, pesticides, and cleaners. Some PFASs, such as perfluorooctanoic acid (PFOA) and perfluorooctane sulfonate (PFOS), have been detected in a wide range of environmental compartments, wildlife species, and human populations worldwide. Given their long half-lives and ability to cross the placental barrier, there is considerable concern about the effects of these substances on fetal growth and development. Article 16 of the Stockholm Convention requires the Conference of the Parties to periodically evaluate whether the Convention is an appropriate method for achieving the objective of protecting human health and the environment from POPs. A global monitoring plan (1) for POPs has been put in place under the Convention. This is a key component of the effectiveness evaluation, and provides a harmonized framework for identifying changes in POP concentrations over time, as well as information on regional and global transport of POPs in the environment. For the first effective evaluation, various initiatives were implemented to ensure broad application of the Convention worldwide, and to obtain at least core representative data from all regions. This evaluation was seen as an opportunity to establish a global-scale set of baseline data for POP concentrations in the environment. The first monitoring reports from five United Nations Regions were submitted at the fourth meeting of the Conference of the Parties in May 2009. Also at this meeting (SC4/32), the Conference agreed that a 6-year period was appropriate for effectiveness evaluations. The second monitoring reports were therefore submitted at the seventh meeting of the Conference of the Parties in May 2015. The aim of this study was to monitor PCDD/Fs and PCBs in air and PFOS and PFOA in water, in line with the requirements of the effectiveness evaluation of the Stockholm Convention in China. Monitoring of POPs in China for the effectiveness evaluation commenced in 2007, and was the first time that POPs had been monitored at locations all over China. Eleven remote sites, referred to as background sites, were selected to determine the background concentrations. Sites corresponded with national air monitoring sites, and were distributed through different geographical regions of China. The monitoring of POPs under the Stockholm Convention in China was supported and organized by the Foreign Economic Cooperation Office of the Ministry of Environmental Protection. The 11 POPs, PCDD/Fs, PCBs, aldrin, chlordane, dichlorodiphenyltrichloroethane, dieldrin, endrin, heptachlor, hexachlorobezene and mirex were determined in air samples collected from the 11 remote sites in China through 3 different time 75 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

periods, namely 2007–2008, 2008–2009, and 2010–2011. The sampling and analytical methods followed those specified in the global monitoring plan (1) for POPs. Concentrations of PFOS and PFOA were determined in water samples from two lakes and two coastal marine areas in 2013 by the China National Environmental Monitoring Center. Description of China

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Country Profile China is in eastern Asia, to the west of the Pacific Ocean. It has a land area of about 9.6 million km2, and a sea area of about 4.73 million km2. China stretches 5,500 km from north to south, from the center of the Heilongjiang River (53°31′N) to the Zengmu Reef (3°52′N) at the southernmost tip of the Nansha Islands in the South China Sea. The distance from the easternmost point of China, at the confluence of the Heilongjiang and Wusuli Rivers in Heilongjiang Province (135°5′E), to the westernmost point on the Pamir Plateau (73°40′E) in the Xinjiang Uygur autonomous region, is 5,200 km. Latitudes, longitudes, and altitudes in China span large ranges, meaning that the climate is diverse with temperate, subtropical, and tropical climate zones. China has more than 22,000 km of land borders, which it shares with 14 other countries. To the east, China is bordered by various seas, including the Bohai Sea, East China Sea, Yellow Sea, and South China Sea, and its total territorial sea area is 4.7 million km2. There are about 7,600 islands in China’s territorial waters. The coastline of mainland China is 18,000 km long, to which the coastlines of the Chinese islands add a further 14,000 km, such that China’s total coastline extends to 32,000 km. China also has a large number of rivers and lakes, and, because of its topographical features, most of the rivers flow east or south into the ocean. The gross domestic product (GDP) of China was 1,098,283,000,000,000 US dollars in 2015. The GDP value of China represents 17.71% of the world economy. Economic development among the different regions of China is unbalanced. The coastal areas in eastern China are comparatively well-developed, and the GDPs of only 5 provinces (municipalities) in the southeastern coastal area (Guangdong, Jiangsu, Shandong, Zhejiang, and Shanghai) account for about 40% of the GDP of the whole country. By comparison, the economies in the middle and western areas lag behind those in eastern provinces, and there are considerable disparities between the technical levels, enterprise scales, and environmental awareness in the eastern and western areas.

Emission and Use Inventories of POPs in China As with all unintentionally produced POPs, any estimation of PCDD/Fs emissions in China is difficult and complex because of the vast range of industries involved. Using the guidelines in the Standardized Toolkit for Identification and Quantification of Dioxin and Furan Releases of the UNEP Chemicals Branch, the potential emission sources of PCDD/Fs in China in 2004 were estimated. 76 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

The total releases of PCDD/Fs from all types of sources in China was 10.2 kg toxic equivalents (TEQ), 5.0, 0.04, 0.17, and 5.0 kg TEQ of which were released to air, released to water, found in products, and found in residues, respectively. Out of all the industries, most PCDD/Fs were released from metal production, including iron, steel, and other metals, and accounted for 45.6% of the total. This was followed by power and heat generation, and waste incineration. These three sources accounted for 81% of the total PCDD/Fs released (2). In China, approximately 10,000 t of PCBs were produced from 1965 to 1974, 9000 t as trichlorobiphenyl and 1000 t as pentachlorobiphenyl. The production and use of PCBs were banned in China in 1974. An earlier investigation found that the most common PCB-containing electrical devices in use in China were capacitors, which are typically used by various large enterprises and industries in the nonpower sector. Because of the wide range of sectors involved, the large number of enterprises, management weaknesses, and the time period required, there are a number of difficulties associated with investigating PCB-containing capacitors in the non-power sector in China. Investigators concluded that there were about 554 PCB-containing capacitors in use in the non-power sector in Liaoning Province, China . For many decades, PFOS and PFOA and their corresponding derivatives were the most widely-used perfluoroalkyl and polyfluoroalkyl substances worldwide. Because of the diverse sources of PFOS, including industrial and consumer uses of PFOS-containing products, it is difficult to establish an emission inventory of PFOS in China; China, however, does have experience in establishing emission inventories for other POPs. Internationally, excellent progress has been made towards establishing emission inventories for global PFOS and PFOA, and existing studies have provided important results that can be applied to China. Two global emission estimations indicated that the production and emissions of PFOS were highest in Asian countries, and, in particular, in China (3, 4). Recently, identified industrial sources of PFOS, including PFOS manufacturing, textile treatment, metal plating, fire-fighting, and semiconductors, were used to compile a national PFOS emission inventory in China (5), which showed that, in general, the emission rates, emission densities, and emission intensities of PFOS were higher in eastern China than in other regions of China.

Materials and Methods Sampling Sampling Sites for Air Samples Monitoring of POPs in ambient air at 11 background sites (B1–B11) in China between 2008 and 2014 was led by the China National Environmental Monitoring Center. The 11 background sites were located in the provinces of Qinghai, Hubei, Liaoning, Heilongjiang, Fujian, Anhui, Hebei, Yunnan, and Shandong, the Tibet Autonomous Region, and the city of Chongqing. These background areas were selected to correspond with national air monitoring sites in different geographical regions of China. Three urban sites (U1–U3) were chosen in Nanjing, Wuhan, 77 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

and Chongqing, and three rural sites (R1–R3) were located in the county-level divisions of Rizhao, Luan, and Yangshuo. Details of the sampling sites are provided in Table 1.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Table 1. Locations of the ambient air sampling sites The codes of the sampling sites

Sampling sites

Longitude

Latitude

B1

Qingyuan

124° 56′ 16″ E

41° 51′ 08″ N

B2

Changdao

120° 41′ 44″ E

37° 59′ 23″ N

B3

Wuyishan

117° 43′ 48″ E

27° 35′ 12″ N

B4

Luan

116° 09′ 36″ E

31° 33′ 05″ N

B5

Lasa

90° 44′ 32″ E

29° 21′ 13″ N

B6

Lijiang

100° 14′ 60″ E

26° 52′ 54″ N

B7

Shennongjia

110° 16′ 16″ E

31° 27′ 26″ N

B8

Daxinganling

121° 14′ 59″ E

50° 52′ 51″ N

B9

Wulong

107° 44′ 47″ E

29° 30′ 39″ N

B10

Chengde

116° 29′ 40″ E

41° 07′ 11″ N

B11

Qinghaihu

100° 29′ 36″ E

36° 35′ 02″ N

R1

Rizhao

119° 18′ 52″ E

35° 41′ 37″ N

R2

Yangshuo

110° 30′ 36″ E

24° 47′ 33″ N

R3

Luan

116° 22′ 15″ E

31° 29′ 09″ N

U1

Chongqing

106° 33′ 43″ E

29° 38′ 44″ N

U2

Wuhan

114° 09′ 36″ E

29° 58′ 20″ N

U3

Nanjing

118° 44′ 44″ E

32° 02′ 35″ N

The criteria for selecting the 11 remote air sampling sites at which to establish background levels followed the guidelines in The Global Monitoring Plan for Persistent Organic Pollutants, as follows: 1.

2.

Regional representation–the location should not be influenced by local sources of POPs and other pollutants, and the air being sampled should be representative of a large region around the site; Minimal meso-scale meteorological circulation influences–the location should be free from strong systematic diurnal variations in local meteorological circulation caused by topography (e.g., upslope/downslope mountain winds, coastal land breezes, or lake breezes);

78 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

3.

4.

5.

Long-term stability–the parameters relating to many aspects of the environment and environmental management should be stable, and these aspects should include the infrastructure, institutional commitment, and land development in the surrounding area; Ancillary measurements–other atmospheric composition and meteorological measurements (wind speed, temperature, humidity, and boundary layer stability) should be taken at the sampling sites; Appropriate infrastructure and utilities–the site should have electrical power, buildings, platforms, towers, and roads, and should be accessible.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Sampling Sites for Water Samples In 2013, the Chinese National Environmental Monitoring Center monitored PFOS and PFOA concentrations in water from 16 sampling sites at 4 locations, as follows: 3 sites in Qinghaihu Lake, 3 sites in Taihu Lake, 5 sites in the Bohai Sea, and 5 sites in the Huanghai Sea. Details of these sampling sites are provided in Table 2.

Air Sampling Methodology Samples of ambient air to be analyzed for POP concentrations were collected at the 11 background sites from 2007 to 2008, from 2008 to 2009, and from 2010 to 2011. Air samples were also collected from the three rural sites and three urban sites in 2012. In line with the Guidance for the Global Monitoring Plan for Persistent Organic Pollutants published by the Secretariat of the Stockholm Convention on POPs in April 2007, and the updated version issued at COP-6 in 2013 (UNEP/POPS/COP.6/INF/31), the air samples were collected at each site by high volume air samplers (Echo Hi Vol, Tecora, Italy) with quartz fiber and PUF filters and size-selective inlets that only accepted particles with diameters of less than 10 µm (PM10). Two samples and a blank sample were collected at each sampling site. Each sample was collected over a period of more than 3 days. The maximum flow rate of the high volume air sampler was 250 mL/min, and each sample had a total volume of approximately 1000 m3.

79 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

80

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Table 2. Information about the water sampling sites in China Sampling sites

C1–C5

C6–C10

W1

W2

Sampling areas

Huanghai Sea

Bohai Sea

Qinghaihu Lake

Taihu Lake

Sampling date

26–27 Apr 2013

17–18 Jun 2013

28 Jun 2013

29 Aug 2013

The codes of the sampling sites

Longitude

Latitude

LN0214

38° 32′ 27.60′′ N

121° 21′ 50.40′′ E

LN0212

38° 40′ 44.40′′ N

121° 54′ 21.60′′ E

LN0211

38° 53′ 56.40′′ N

122° 23′ 52.80′′ E

LN0206

39° 16′ 52.32′′ N

123° 08′ 24.00′′ E

LN0220

39° 07′ 59.88′′ N

122° 46′ 59.88′′ E

LN0201

40° 09′ 36.00′′ N

121° 04′ 55.20′′ E

LN0208

39° 12′ 39.60′′ N

121° 05′ 31.20′′ E

TJ13

38° 39′ 56.55′′ N

118° 04′ 29.82′′ E

TJ17

38° 36′ 53.86′′ N

118° 28′ 03.81′′ E

TJ14

38° 38′ 59.79′′ N

118° 53′ 53.24′′ E

QH-B

36° 53′ 32.59′′ N

100° 03′ 33.99′′ E

QH-D

36° 47′ 51.39′′ N

100° 16′ 22.36′′ E

QH-E

36° 44′ 32.16′′ N

100° 24′ 15.99′′ E

XSX

31° 08′ 04.92′′ N

120° 11′ 22.92′′ E

PTS

31° 13′ 32.88′′ N

120° 06′ 11.88′′ E

JSN

31° 17′ 02.51′′ N

120° 06′ 39.26′′ E

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

POP Analysis Methods

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Analysis of PCDD/Fs The entire analytical process was carried out as outlined in method 1613B of the United States Environmental Protection Agency (US EPA). The quartz fiber filter and the PUF filter were combined during sample preparation and spiked with 13C-labeled compound solution (Wellington Laboratories, Guelph, Canada). The combined sample was then extracted with dichloromethane and hexane (1:1, v/v) using accelerated solvent extraction (ASE 300, Dionex, U.S.A.). After concentration, an automated sample preparation system (Power-Prep™, Fluid Management System, U.S.A.) was used for sample cleanup. This cleanup process comprised, in the following order, a multilayer acid/base/neutral silica column, a basic alumina column, and a carbon column. The final eluate was concentrated under a gentle stream of purified nitrogen, and the residue was dissolved in nonane (10 μL) in a mini vial. To evaluate the recovery, the sample extract was spiked with a 13C-labeled internal standard (Wellington Laboratories) immediately before instrumental analysis. PCDD/Fs were determined by a gas chromatograph (Agilent 6890, Agilent Technologies, U.S.A.) coupled with a high-resolution mass spectrometer (Waters Autospec Ultima, Waters Corp., U.S.A.) by tracing the M+, (M+2)+, or the most intense ions of the isotope cluster. PCDD/F congeners were determined on a DB 5 MS column (60 m × 0.25 mm i.d., 0.25 µm film thickness, Agilent Technologies). Helium, at a flow rate of 1.2 mL/min, was the carrier gas. The injection volume was 1 µl in splitless mode with a splitless period of 60 s. The MS was operated at a resolution of >10 000 in electron ionization mode (35 eV) with selected ion monitoring. Calibration standard solutions were analyzed with each batch of samples to check the instrument stability and variance in the relative response factor. For quality control, the retention times of the analytes in a sample had to be within 2 s of the retention times of the internal standards. Isotope ratios of the two monitored ions for each compound had to be within 15% of the theoretical chlorine values. The limit of detection for PCDD/Fs in a given sample was defined as a signal-to-noise ratio greater than three times the average baseline variation.

Analysis of DL-PCBs Analysis of the samples for DL-PCBs followed US EPA method 1668A. Stable isotope labeled analogs of the DL-PCB congeners were quantitatively spiked into the PUF and quartz fiber filter samples, and then the samples were Soxhlet-extracted using toluene or extracted using ASE with hexane and dichloromethane. The extracts were cleaned up by column chromatography using acidic silica gel, multi-layer silica gel, activated alumina, and dual-layer carbon columns. The extracts were then evaporated to near dryness, and stable isotope labeled PCB congeners were added as recovery standards before HRGC-HRMS analysis. 81 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Analysis of Marker PCBs The samples were analyzed for marker PCBs by HRGC-HRMS. Stable isotope labeled analogs of the marker PCB congeners were quantitatively spiked into the PUF and quartz fiber filter samples, and then the samples were extracted using ASE with hexane and dichloromethane. The extracts were cleaned up by column chromatography using acidic silica gel, multi-layer silica gel, and activated alumina columns. The extracts were then evaporated to near dryness, and stable isotope labeled PCB congeners were added as recovery standards before HRGC-HRMS analysis. Stable isotope labeled analogs of the marker PCB congeners (PCBs 28, 52, 101, 118, 153, and 180) were quantitatively spiked into the PUF and quartz fiber filter samples, and then the samples were Soxhlet-extracted using toluene. The extracts were cleaned up using an acid–base partitioning method and column chromatography comprising acidic silica gel, multi-layer silica gel, and activated alumina columns. The extracts were then evaporated to near dryness, and stable isotope labeled PCBs 70, 111, 138, and 170 were added as recovery standards before GC-MS analysis.

Analysis of PFOS and PFOA The water samples were analyzed for PFOS and PFOA concentrations following US EPA method 537. Each water sample was filtered through a 0.22-µm filter, and then extracted and cleaned up using an Oasis WAX solid phase extraction cartridge. Before use, each column was activated using ammonia in methanol followed by methanol. After adding a sample to a cartridge, it was eluted with an acetic acid buffer solution, methanol, and ammonia in methanol. The extracts were concentrated and PFOS and PFOA were determined by liquid chromatography-electrospray ionization-tandem mass spectrometry. Quality Assurance and Quality Control Quality Assurance and Quality Control for the PCDD/F and PCB Analyses All of the laboratories that analyzed the samples for PCDD/Fs and PCBs were ISO/IEC 17025 accredited. To ensure that high quality data were collected in this study, procedures were carefully followed during collection of field and lab blanks, duplicates from collocated samplers, and breakthrough samples. Field and laboratory blanks were collected with each set of samples and processed following the same methods. The method detection limits were determined using the background concentrations in these blanks rather than the instrumental detection limit. None of the less-chlorinated congeners were detected in the blanks. Octachlorodibenzodioxin (OCDD) was the most prevalent congener in the blanks, followed by 1,2,3,4,7,8-hexachlorodibenzodioxin. Together, these two congeners made up 90% of the total PCDDs and PCDFs detected in the blanks. However, their concentrations in the blanks corresponded to less than 8% 82 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

of the concentrations found in the air samples. The concentrations in duplicate samples obtained at the collocated sites agreed well with each other. Among the 17 congeners, 2,3,7,8-tetrachlorodibenzodioxin was the most difficult to measure because its concentrations were extremely low. Recoveries of each chemical during the clean-up procedure were calculated and determined separately from surrogates. The average recoveries of the 13C12-PCDD/Fs ranged from 56% to 122%. DL-PCB congeners were not detected in the blank samples. The most prevalent congener in the blanks was CB28, followed by CB52. Together, these two congeners made up 96% of the total indicator PCBs detected in the blanks. However, their concentrations in the blanks corresponded to less than 5% of the concentrations detected in the air samples. The average recoveries of the 13C12-DL-PCBs and indicator PCBs ranged from 40% to 118%. The laboratories participated in several global POPs interlaboratory assessments, and the analytical results agreed well with the median values from all of the participating laboratories worldwide.

Quality Assurance and Quality Control for PFOA and PFOS Analyses Quality Assurance and Quality Control for the PFOS and analyses followed US EPA method 537. The recoveries of PFOS ranged from 70% to 130%, and the analytical replicates were within 25%.

Results Levels of PCDD/Fs and PCBs in Ambient Air The PCDD/F concentrations detected in the background air samples between 2007 and 2011 are shown in Table 3, and the PCDD/F concentrations detected in the urban and rural air samples in 2012 are shown in Table 4. The World Health Organization (WHO)-TEQ is the TEQ calculated using WHO toxic equivalency factors (TEF). The PCDD/F and DL-PCB concentrations in the background air samples ranged from 2.64 to 101.7 WHO-TEQ fg/m3 (with an average of 21.7 WHO-TEQ fg/m3). The PCDD/F concentrations in the background air samples ranged from 0.84 to 98.05 WHO-TEQ fg/m3 (with an average of 20.03 WHOTEQ fg/m3). The concentrations were highest at site B3, in Southeast China. The concentrations were lowest at site B10, in North China. The concentrations of PCDD/Fs were higher in eastern China than in western China, and may reflect the more intensive economic development in eastern regions relative to that in western regions. The PCDD/F concentrations in the air samples from the three urban and three rural sites were between 91.3 and 202 WHO-TEQ fg/m3 (average of 164 WHO-TEQ fg/m3) and between 9.45 and 42.7 WHO-TEQ fg/m3 (average of 22.2 WHO-TEQ fg/m3). The average concentrations of PCDD/Fs were much higher at the urban sites than at the background and rural sites. Comparison showed that the concentrations of PCDD/Fs were lower than the annual average concentrations of PCDD/Fs in ambient air in Porto, Portugal (range 29.4–298.3 fg/m3) (6). Chen et al. reported PCDD/F concentrations of between 2.6 and 120 pg/m3 (0.04–1.93 83 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

pg I-TEQ /m3) at various sampling sites in China (7), which were higher than the levels of PCDD/Fs detected in this study. In addition, the concentrations at the background, rural, and urban sites were all below the ambient air standard of 600 fg TEQ/ m3 proposed by the government of Japan (8). The DL-PCB and marker PCB concentrations in the background air samples between 2007 and 2011 are shown in Table 5, while the DL-PCB and marker PCB concentrations found in the air samples from the three urban and three rural sites in mainland China in 2012 are shown in Table 6. The DL-PCB concentrations in the background, urban, and rural air samples ranged from 0.10 to18.6 WHOTEQ fg/m3, from 6.84 to 14.2 WHO-TEQ fg/m3 (average 10.2 WHO-TEQ fg/ m3), and from 0.93 to 3.73 WHO-TEQ fg/m3 (average 2.14 WHO-TEQ fg/m3), respectively. DL-PCB levels in South Korea were similar, and ranged from not detected to 16 WHO-TEQ fg /m3 (average of 8 WHO-TEQ fg/m3) (9). The marker PCB concentrations in the air samples from the 11 background sites, 3 urban sites, and 3 rural sites were between 4.70 and 44.9 pg/m3 (average 18.5 pg/m3), 55.1 and 90.6 pg/m3 (average 71.6 pg/m3), and 4.62 and 10.3 pg/m3 (average 7.32 pg/m3), respectively. The annual average of the sum of the seven marker PCBs in the atmosphere in Nordic regions in 2011 was 12.5 pg/m3 (10), which was similar to the levels of indicator PCBs at the background sites in the present study.

Congener Profiles of PCDD/Fs and PCBs The major contributors to atmospheric 2,3,7,8-PCDD/Fs measured at the background sites during the sampling periods of this study were 1,2,3,4,6,7,8-heptachlorodibenzofuran, octachlorodibenzofuran, and OCDD (Figure 1), and accounted for between 54% and 92% of the Σ2,3,7,8-PCDD/Fs. As with other studies (11, 12), OCDD was the major contributor to the total annual average PCDD/Fs congener concentration, with a concentration and a percentage contribution to the total of 223 fg/m3 and 27%, respectively. In the ambient air samples, PCDD/Fs tended to have a higher degree of chlorination. Most of the 2,3,7,8-PCDD/Fs congeners were detected, except for 2,3,7,8-trichlorodibenzodioxin and 2,3,7,8-tetrachlorodibenzodioxin. PCDFs accounted for between 11% and 47% of total PCDD/Fs. These results are consistent with those reported in samples collected for national dioxin inventories in ambient air in both the United Kingdom (13), and in Catalonia, Spain (14). The dominant DL-PCB congeners were CB118, CB105, and CB77, as shown in Figure 2. The congener profiles of indicator PCBs were dominated by CB28 and CB52, which are tri- and tetra-PCBs, respectively. The percentage contributions of the two less-chlorinated congeners (CB28 and CB52) to the Σ7PCBs concentrations were between 54% and 98% (Figure 3). Generally, less-chlorinated PCBs have lower vapor pressure than highly chlorinated PCBs, and the lighter congeners are expected to be more readily transferred to air. The DL-PCB concentrations were much lower than those of indicator PCBs, but they were still detected because of the relatively low method detection limit. The most abundant congeners were CB118 (mono-ortho), CB77 (non-ortho), and CB105. Because of its high TEF, CB126 accounted for more than 85% of the total TEQ 84 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

for all samples. These results agree with those reported by Die et al. (15) for samples collected from industrial sites in Shanghai, China.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Levels of PFOS and PFOA in Water The PFOS and PFOA concentrations were measured in water samples collected from the Bohai Sea and Huanghai Sea (Table 7) and Qinghai Lake and Taihu Lake (Table 8). The average PFOS concentration in the water samples from Taihu Lake was 32 ng/L, while the levels in all the water samples from the Bohai Sea, Huanghai Sea, and Qinghai Lake were below the detection limit. Comparison showed that our results were lower than, or comparable with, the ranges for the concentrations of PFOA (4–93 ng/L) and PFOS (3–29 ng/L) in surface water from Tamil Nadu, India (16). PFOA and PFOS levels in this study were below the permissible limits set by the US EPA (17).

85 Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

86

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Table 3. Concentrations (fg/m3) of 2,3,7,8-PCDD/F in air samples collected at background (B) sites in China B1

B1

B1

B2

B2

B2

B3

B3

B4

B4

B4

20072008

20082009

20102011

20072008

20082009

20102011

20072008

20102011

20072008

20082009

20102011

2378-TCDF

23.6

4.6

4.41

32.1

18.9

18.4

36.50

12.6

3.7

21.8

9.5

12378-PeCDF

19.7

5.2

3.72

18.45

21.5

13.3

49.25

22.3

2.85

33.3

14

23478-PeCDF

35.3

7.1

5.63

24.45

29.6

17.1

82.00

23.1

3.35

44.8

29

123478-HxCDF

66.65

7.7

5.77

66.3

32.9

14.3

125.95

64.7

5.8

81.4

47.3

123678-HxCDF

46.5

7.5

4.44

46.2

32.5

13.3

91.60

33.1

5.65

79.1

64.8

234678-HxCDF

47.75

8.4

5.77

56.55

37.4

10.2

115.30

33.7

5.7

69.2

50

123789-HxCDF

14.3

1.7

3.1

14.25

9.2

2.86

39.70

18

1.25

14.8

14

1234678-HpCDF

289.45

27.1

70

222.2

124

31.4

449.25

203

28.4

400

392

1234789-HpCDF

37.05

3.7

3.86

11.6

16.7

2.54

68.30

39.6

3.35

50.4

57.3

OCDF

531.9

22.5

87.4

358.4

106

26.4

744.55

420

145.7

458

431

2378-TCDD

2.6

0.4

0.69

3.8

0.7

0.32

3.60

0.98

0.75

0.8

0.54

12378-PeCDD

7.95

1.2

2.58

3.85

4.3

2.44

2.80

5.73

0.9

5.2

3.76

123478-HxCDD

4.75

0.8

2.15

5.15

3.2

0.63

7.85

5.07

0.7

3.9

4.84

6

2.22

14.60

9.99

0.9

6.5

6.81

123678-HxCDD

5.3

1.7

1.76

5

123789-HxCDD

7.7

1.1

2.15

4.9

4.7

0.63

15.10

6.55

1.1

6.1

18.6

1234678-HpCDD

95.55

8.6

14.1

54.2

36.7

17.8

145.05

129

19.6

43.6

67.7

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

87

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

B1

B1

B1

B2

B2

B2

B3

B3

B4

B4

B4

20072008

20082009

20102011

20072008

20082009

20102011

20072008

20102011

20072008

20082009

20102011

OCDD

445.65

24.5

103

272.95

80.1

82.5

589.70

466

334.7

91.6

219

WHO-TEQ

49.6

9.1

7.23

45.45

37.2

14.9

98.05

37.4

5.6

63.3

40.1

B5

B5

B5

B6

B6

B6

B7

B7

B8

B8

B8

20072008

20082009

20102011

20072008

20082009

20102011

20072008

20102011

20072008

20082009

20102011

2378-TCDF

3.7

2.7

0.17

5.75

11.5

9.1

66.8

2.22

25.6

8.8

9.86

12378-PeCDF

1.5

2

0.75

2.45

11.2

9.17

81.9

0.66

5.25

3.2

2.57

23478-PeCDF

6.3

5

0.2

4

13.8

13.7

96.9

0.87

10.8

6

4.91

123478-HxCDF

3.7

11.8

1.7

4.3

11.7

11.7

267.3

3.52

22.35

3.7

1.76

123678-HxCDF

4.1

14

0.78

5.05

11.5

11.8

261.65

1.95

14

3.9

3.51

234678-HxCDF

5.2

13.1

1.12

5.2

10

11.3

285.3

2.44

16.4

4.1

3.87

123789-HxCDF

4.3

2.9

0.66

3.25

2.8

3.4

61.6

0.68

2.25

0.7

0.41

1234678-HpCDF

111.2

106

6.95

58.45

21.7

32.3

2098.7

15.6

56.15

10.2

11.7

1234789-HpCDF

5.2

11.7

1.06

2.05

3.2

4.98

255.05

1.53

6.05

1

1.67

OCDF

186

121

29

140.7

9.5

24

2867.85

38.1

77.4

13.2

32.1

2378-TCDD

4.4

ND

ND

2.65

1

0.69

4.8

0.83

2.35

ND

1.02

12378-PeCDD

4.8

1.1

0.37

2.75

2

1.93

8.2

1.3

1.9

ND

1.15

Continued on next page.

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

B5

B5

B5

B6

B6

B6

B7

B7

B8

B8

B8

20072008

20082009

20102011

20072008

20082009

20102011

20072008

20102011

20072008

20082009

20102011

123478-HxCDD

3.8

1.1

0.37

2.45

2.2

1.77

5.55

0.84

1.3

0.5

0.72

123678-HxCDD

3.7

1.7

1.47

2.1

3.7

2.45

35.85

0.68

2.15

1.2

0.8

123789-HxCDD

2.6

1.9

1.84

2.45

2.8

3.04

29.8

0.44

1.45

0.4

1.02

1234678-HpCDD

35

9.5

10.6

26.8

27.6

22.9

208.95

12.6

18.4

9.1

12.4

439.2

35.8

67.4

285.05

178

59.9

735.2

83.2

168.25

174

66.6

10.2

10.3

2.35

8.35

16.6

12.7

186.25

3.01

16.25

5.7

5.7

OCDD WHO-TEQ

88

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Table 3. (Continued). Concentrations (fg/m3) of 2,3,7,8-PCDD/F in air samples collected at background (B) sites in China

B9

B9

B9

B10

B10

B10

B11

B11

B11

2007-2008

2008-2009

2010-2011

2007-2008

2008-2009

2010-2011

2007-2008

2008-2009

2010-2011

2378-TCDF

6.95

6.7

3.35

17.55

5.8

0.1

2.50

3.8

7.85

12378-PeCDF

5.4

9.3

1.39

13.75

6.6

0.2

2.50

2.1

11.8

23478-PeCDF

5.7

10.3

4.74

19.3

4.4

0.1

1.00

4.2

7.85

123478-HxCDF

14

19.8

6.13

22.15

13.9

0.8

3.40

4.9

9.97

123678-HxCDF

10.05

25.2

5.58

21.4

14.2

0.5

6.00

4

9.36

234678-HxCDF

12.95

18.2

5.58

16.05

15.8

0.9

6.30

4.1

11.2

123789-HxCDF

4.15

2.1

1.12

9.6

4.2

0.4

0.50

0.8

0.30

1234678-HpCDF

159.95

110

37.1

76.05

60

6

54.60

18.9

36.50

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

89

B9

B9

B9

B10

B10

B10

B11

B11

B11

2007-2008

2008-2009

2010-2011

2007-2008

2008-2009

2010-2011

2007-2008

2008-2009

2010-2011

1234789-HpCDF

12.85

11.9

4.74

11.2

8.6

0.8

6.15

1.6

3.93

OCDF

249.75

69.4

62.2

114.55

68.7

14.3

136.00

44.5

71.60

2378-TCDD

3.55

ND

0.24

10.95

0.5

0.06

1.55

ND

0.60

12378-PeCDD

2.8

0.4

0.84

9.85

0.4

0.04

5.20

ND

4.83

123478-HxCDD

3

1.2

0.28

3.95

0.7

0.2

1.70

1.3

1.51

123678-HxCDD

1.2

1.6

1.67

1.65

1.5

0.6

1.35

2.8

3.62

123789-HxCDD

2.15

1.6

1.39

4.2

1.2

0.2

1.15

1.4

4.23

1234678-HpCDD

23.05

19.9

26.2

23.3

18.4

5.8

19.10

25.4

22.60

OCDD

232.35

120

55.8

330.6

194

17.7

197.10

555.00

65.2

WHO-TEQ

12.3

15.6

7.17

28.6

10

0.84

6.80

6.80

13.9

Toxic equivalent (TEQ) values were calculated using World Health Organization- toxic equivalency factors (TEF2005) was used to calculate the TEQ when the concentration was below the limit of detection (LOD).

ND = not detected

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

A value of 0

Table 4. PCDD/F concentrations (fg/m3) in urban (U) and rural (R) air samples collected in China

90

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Urban

Rural

sampling sites

U3

U2

U1

U2

R1

R3

sampling date

Oct 2012

Nov 2012

Nov 2012

Jun 2012

Nov 2012

Dec 2012

2378-TCDF

66.6

107

108

10.4

20.3

4.40

0.04

12378-PeCDF

35.4

113

111

9.80

ND

ND

0.2

23478-PeCDF

99.2

169

182

8.75

46.8

10.2

0.2

123478-HxCDF

116

168

187

9.80

56.6

16.6

0.1

123678-HxCDF

88.3

146

164

9.90

43.8

13.7

0.1

234678-HxCDF

106

176

205

6.60

48.2

21.2

0.1

123789-HxCDF

29.8

44.2

47.0

ND

12.0

4.60

0.1

1234678-HpCDF

387

553

639

33.3

190

94.4

0.1

1234789-HpCDF

51.4

74.8

78.4

3.75

22.4

12.0

0.1

OCDF

409

364

380

26.2

221

124

0.1

2378-TCDD

ND

12.2

9.50

ND

ND

ND

0.1

12378-PeCDD

ND

15.8

32.8

ND

ND

ND

0.1

123478-HxCDD

8.20

20.2

27.7

ND

5.30

ND

0.2

123678-HxCDD

21.1

40.4

65.0

2.20

8.70

3.80

0.2

123789-HxCDD

14.2

30.4

54.6

2.10

6.80

3.20

0.2

1234678-HpCDD

160

222

482

32.1

70.4

44.4

0.1

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

LOD

Rural

sampling sites

U3

U2

U1

U2

R1

R3

sampling date

Oct 2012

Nov 2012

Nov 2012

Jun 2012

Nov 2012

Dec 2012

194

198

134

0.1

9.45

42.7

14.5

--

OCDD

376

411

1.18×103

WHO-TEQ

91.3

202

198

Toxic equivalent (TEQ) values were calculated using World Health Organization-toxic equivalency factors (TEF2005) was used to calculate the TEQ when the concentration was below the limit of detection (LOD).

ND = not detected

91

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Urban

Loganathan et al.; Persistent Organic Chemicals in the Environment: Status and Trends in the Pacific Basin Countries II ... ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

LOD

A value of 0

92

Publication Date (Web): December 7, 2016 | doi: 10.1021/bk-2016-1244.ch004

Table 5. Concentrations of DL-PCBs (fg/m3) and marker PCBs (pg/m3) in air samples collected at background (B) sites in China B1

B1

B1

B2

B2

B2

B3

B3

B4

B4

B4

20072008

20082009

20102011

20072008

20082009

20102011

20072008

20102011

20072008

20082009

20102011

CB77

4.9

25.3

8.98

94.4

272

737

222

30.1

289

96.9

41

CB81

3.4

5.6

4.24

19.2

11.7

15.4

21.7

12

22.6

14.7

13.3

CB105

65.2

38.1

15.1

288

98.6

44.5

128

38.9

66.7

76.3

34.2

CB114

10.3

9.7

5.57

144

23.7

17.6

22.7

12.5

33.6

24.9

13.3

CB118

211

122

64

606

278

101

434

119

416

228

91.2

CB123