• No results found

Tracing and apportioning sources of dioxins using multivariate pattern recognition techniques

N/A
N/A
Protected

Academic year: 2021

Share "Tracing and apportioning sources of dioxins using multivariate pattern recognition techniques"

Copied!
65
0
0

Loading.... (view fulltext now)

Full text

(1)

Tracing and apportioning sources

of dioxins using multivariate

pattern recognition techniques

Anteneh Assefa

A dissertation submitted to Umeå University for the degree of Doctor of Philosophy in the Faculty of Science and Technology

Faculty opponent: Professor Shigeki Masunaga Department of Risk Management and Environmental Sciences

Graduate School of Environment and Information Sciences Yokohama National University

(2)

Copyright © Anteneh Assefa ISBN: 978-91-7601-285-7

Cover picture: Diavolessa, https://eu.fotolia.com Back picture: Iculig, https://eu.fotolia.com Printed by: KBC Service Centre

(3)

Dedicated to my beloved mother Yeshi and sisters Mihret, Mesay and Selam and my sweet daughter Melkam

(4)
(5)

Tracing and apportioning sources of dioxins using

multivariate pattern recognition techniques

Anteneh Assefa, Department of Chemistry, Umeå University, SE-901 87 Umeå, Sweden.

Abstract

High levels of polychlorinated p-dioxins and polychlorinated dibenzo-furans (PCDD/Fs) in edible fish in the Baltic Sea have raised health concerns in the Baltic region and the rest of Europe. Thus, there are urgent needs to characterize sources in order to formulate effective mitigation strategies. The aim of this thesis is to contribute to a better understanding of past and present sources of PCDD/Fs in the Baltic Sea environment by exploring chemical fingerprints in sediments, air, and biota. The spatial and temporal patterns of PCDD/F distributions in the Baltic Sea during the 20thcentury were studied in

Swedish coastal and offshore sediment cores. The results showed that PCDD/F levels peaked in 1975 (± 7 years) in coastal and 1991 (± 5 years) in offshore areas. The time trends of PCDD/Fs in the sediment cores also showed that environmental half-lives of these pollutants have been shorter in coastal than in offshore areas (15 ± 5 and 29 ± 14 years, respectively). Consequently, there have been remarkable recoveries in coastal areas, but slower recovery in offshore areas with 81 ± 12% and 38 ± 11% reductions from peak levels, respectively.

Source-to-receptor multivariate modeling by Positive Matrix Factorization (PMF) showed that six types of PCDD/F sources are and have been important for the Baltic Sea environment: PCDD/Fs related to i) atmospheric back-ground, ii) thermal processes, iii) manufacture and use of tetra-chlorophenol (TCP) and iv) penta-chlorophenol (PCP), v) industrial use of elementary chlo-rine and the chloralkali-process (Chl), and vi) hexa-CDD sources. The results showed that diffuse sources (i and ii) have consistently contributed >80% of the total amounts in the Southern Baltic Sea. In the Northern Baltic Sea, where the biota is most heavily contaminated, impacts of local sources (TCP, PCP and Chl) have been higher, contributing ca. 50% of total amounts. Among the six sources, only Thermal and chlorophenols (ii-iv) have had major impacts on biota. The impact of thermal sources has, however, been declining as shown from source apportioned time-trend data of PCDD/Fs in Baltic herring. In contrast, impacts of chlorophenol-associated sources generally increased, remained at steady-state or slowly decreased during 1990-2010, suggesting that

(6)

these sources have substantially contributed to the persistently high levels of PCDD/Fs in Baltic biota.

Atmospheric sources of PCDD/Fs for the Baltic region (Northern Europe) were also investigated, and specifically whether the inclusion of parallel mea-surements of metals in the analysis of air would help back-tracking sources. PCDD/Fs and metals in high-volume air samples from a rural field station near the shore of the central Baltic Sea were measured. The study focused on the winter season and air from the S and E sectors, as these samples showed elevated levels of PCDD/Fs, particularly PCDFs. Several metals were found to correlate significantly with the PCDFs. The wide range of candidate metals as source markers for PCDD/F emissions, and the lack of an up-to-date extensive compilation of source characteristics for metal emission from various sources, limited the use of the metals as source markers. The study was not able to pin-point primary PCDD/F sources for Baltic air, but it demonstrated a new promising approach for source tracing of air emissions. The best leads for back-tracking primary sources of atmospheric PCDD/Fs in Baltic air were seasonal trends and PCDD/F congener patterns, pointing at non-industrial related thermal sources related to heating. The non-localized natures of the sources raise challenges for managing the emissions and thus societal efforts are required to better control atmospheric emissions of PCDD/Fs.

Keywords: polychlorinated dibenzo-p-dioxin, polychlorinated dibenzofuran, positive matrix factorization, PMF, principal component analysis, PCA, Baltic Sea, sediment core, PCDD/F, sources, marine, fish, environmental half-live, peak year, peak level, temporal trend, spatial variation, coastal, offshore, chem-ical fingerprint

(7)

Contents

Page Abstract v List of Abbreviations 1 Publications 3 1 INTRODUCTION 5 1.1 Background . . . 5 1.2 Objective . . . 7

2 MATERIALS AND METHODS 9

2.1 Sampling and analysis of sediment cores . . . 9 2.2 Air sampling and analysis . . . 11 2.3 Baltic herring . . . 12

3 Data Analysis 13

3.1 Positive matrix factorization (PMF) . . . 13 3.2 Principal component analysis (PCA) . . . 14 3.3 Spline smoothing . . . 15

4 PCDD/F dynamics in the Baltic Sea 17

4.1 Spatial and temporal trends of PCDD/Fs in the Baltic Sea by the Swedish coast . . . 17 4.2 Spatial and temporal trends of PCDD/Fs in offshore Baltic Sea

areas . . . 18 4.3 Environmental half-lives . . . 19

5 Sources of PCDD/Fs in the Baltic Sea 23

5.1 Historical sources in sediments . . . 23 5.1.1 Source identification . . . 23

(8)

5.1.2 Source apportionment . . . 24

5.2 Historical sources of PCCD/Fs in biota . . . 27

5.2.1 Transformation indices (TI) . . . 27

5.2.2 Sources and temporal changes in their impacts . . . 28

5.3 Atmospheric sources . . . 31

5.3.1 PCDD/F source regions, congener patterns and seasonality 31 5.3.2 The potential to use metals as source markers . . . 33

6 Conclusions and future work 39

Acknowledgments 43

(9)

List of Abbreviations

AB Atmospheric background BB Bothnian Bay BC Black carbon BP Baltic Proper BS Bothnian Sea Chl Chlorine dw Dry weight E East fg Femtogram GC Gas chromatography HCB Hexachlorobenzene

HELCOM Baltic Marine Environment Protection Commission (Helsinki Commission) HL Half-life HpCDF Heptachlorinated dibenzofuran HpD 1,2,3,4,6,7,8-Heptachlorodibenzo-p-dioxin HpF1 1,2,3,4,6,7,8-Heptachlorodibenzofuran HpF2 1,2,3,4,7,8,9-Heptachlorodibenzofuran

HRMS High-resolution mass spectrometry

HxCDD Hexachlorinated dibenzo-p-dioxin HxD1 1,2,3,4,7,8-Hexachlorodibenzo-p-dioxin HxD2 1,2,3,6,7,8-Hexachlorodibenzo-p-dioxin HxD3 1,2,3,7,8,9-Hexachlorodibenzo-p-dioxin HxF1 1,2,3,4,7,8-Hexachlorodibenzofuran HxF2 1,2,3,6,7,8-Hexachlorodibenzofuran HxF3 1,2,3,7,8,9-Hexachlorodibenzofuran HxF4 2,3,4,6,7,8-Hexachlorodibenzofuran

(10)

ICP-AES Inductively coupled plasma atomic emission spectroscopy ICP-SFMS Inductively coupled plasma-sector field mass spectrometry

LOD Limit of detection

lw Lipid weight

MSWI Municipal solis waste incinerator

N North

OCDD / OD Octachlorodibenzodioxin

OCDF / OF Octachlorodibenzofuran

PAH Polycyclic aromatic hydrocarbon

PC Principal component

PCA Principal component analysis

PCB Polychlorinated biphenyl PCDD Polychlorinated dibenzo-p-dioxin PCDF Polychlorinated dibenzofuran PCP Pentachlorophenol PD 1,2,3,7,8-Pentachlorodibenzo-p-dioxin PF1 1,2,3,7,8-Pentachlorodibenzofuran PF2 2,3,4,7,8-Pentachlorodibenzofuran pg Picogram

PMF Positive matrix factorization

POP Persistent organic pollutants

PUF Polyurethane Foam

S/N Signal to noise ratio

TCDD Tetrachlorinated dibenzo-p-dioxin

TCP Tetrachlorophenol

TD 2,3,7,8-Tetrachlorodibenzo-p-dioxin

TEF Toxicity equivalent factor

TEQ Toxic equivalent

TF 2,3,7,8-Tetrachlorodibenzofuran

TI Transformation index

USEPA United States Environmental Protection Agency

(11)

Publications

This thesis is based on the following papers, which in the text will be referred to by their Roman numerals.

I Assefa, A. T.; Sobek, A.; Sundqvist, K.; Cato, I.; Jonsson, P.; Tysklind, M.; Wiberg, K. Temporal trends of PCDD/Fs in Baltic Sea sediment cores covering the 20th century. Environ. Sci. Technol. 2014, 48 (2), 947−953.

II Assefa, A. T.; Tysklind, M.; Sobek, A.; Sundqvist, K.; Geladi, P.; Wiberg, K. Assessment of PCDD/F source contributions in Baltic Sea sediment core records. Environ. Sci. Technol. 2014, 48 (2), 9531−9539.

III Anteneh Assefa; Mats Tysklind; Anders Bignert; Karin Wiberg. Source-apportionment shows why dioxin levels have not declined in Baltic fish. (Submitted for publication)

IV Assefa AT.; Tysklind M.; Wiberg K. Airborne dioxins and other POPs in the Baltic Sea environment: the potential of using metals as source markers. (Manuscript)

(12)
(13)

Chapter 1

INTRODUCTION

1.1 Background

Polychlorinated dibenzo-p-dioxins (PCDDs) and polychlorinated dibenzofu-rans (PCDFs), commonly referred to as dioxins, are classes of persistent or-ganic pollutants (POPs) that are toxic and bioaccumulative [1, 2]. PCDD/Fs are distributed globally via long-range transport [3–5]. Seventeen (2,3,7,8-substituted) CDD/F congeners are known to be particularly toxic, poorly metabolized and highly bioaccumulative in higher organisms. Of the 17 2,3,7,8-substituted congeners, TCDD is regarded as the most toxic congener. The total potential toxicity of a mixture of the 2,3,7,8-substituted PCDD/Fs can be expressed in concentrations of Toxic Equivalents (TEQs). Several sets of Toxicity Equivalent Factor (TEFs), which describe toxicities of the congeners relative to TCDD, have been proposed over the years (see Table 1.1) [6– 9]. WHO2005 is the latest evaluation of TEFs used for calculating TEQs in different matrices. Various health problems associated with chronic exposure to PCDD/Fs have been reported, including impairment of immune, nervous, endocrine and reproductive systems [10, 11]. Indications that PCDD/Fs are carcinogenic are also emerging [12, 13].

PCDD/Fs are produced predominantly unintentionally as by-products dur-ing industrial processes such as production of organochlorine pesticides [14–16] and bleaching of pulp and paper using elementary chlorine [8,17]. Other tradi-tional sources of PCDD/Fs include incineration of various materials, including municipal solid waste (MSW) and fuels such as coal and wood. Considerable anthropogenic emission of PCDD/Fs started in the 1920s following

(14)

industrial-CHAPTER 1. INTRODUCTION

ization [18]. At that time the hazardous effects of PCDD/Fs were not known, but as awareness grew from the end of the 1970s and onwards, a number of goals were formulated, measures were taken, and programs were launched to reduce the primary production and emissions of PCDD/Fs [14, 19–27]. For example: use of pentachlorophenol (PCP) as a pesticide and wood preservative was banned in increasing numbers of countries from the late 1970s [28,29]; the fifth EU Action Program included a goal to reduce industrial emissions of PCDD/Fs by 90% before 2005 [14]; the UNEP Stockholm Convention on Persistent Or-ganic Pollutants imposed various restrictions [30]; and the HELCOM Baltic Sea Action Plan was launched with ambitious aims (inter alia) to restore the abundance of hazardous substances in the Baltic Sea to close to natural levels [31]. During the 1980’s, in Sweden, moratorium on constructions of municipal solid waste incinerators (MSWI), immediately followed by guidelines, resulted in strict emission control of PCDD/Fs [32]. Consumers preference of dioxin-free materials and elementary chlorine dioxin-free (ECF) processes in pulp-and paper industries, significantly reducing the emission PCDD/Fs [33]. Consequently, the environmental levels and health risks associated with exposure to PCDD/Fs have been declining (Paper I).

The Baltic Sea is a polluted semi-enclosed sea surrounded by several in-dustrialized countries in Northern Europe, and all together almost 86 million people are living in 16 countries in the drainage area. Due to high possible inflows of PCDD/Fs and slow exchange of water with the North Sea (and thus long residence times for contaminants), the Baltic Sea is one of the most polluted natural marine systems in the world [34]. Several emission inventories and environmental fate models have shown that levels of these pollutants in the Baltic area are declining [34, 35]. Similar declines have also been detected in dated sediments from the Baltic Sea [36, 37]. However, PCDD/Fs still raise concern in the Baltic Sea region due to the high levels measured in fatty fish, which occasionally exceed the maximum limits for food and feed assigned by the European Commission [38–40]. Thus, there is an urgent continuous need to identify sources that are contributing PCDD/Fs to the Baltic Sea system. Various aspects of PCDD/F pollution in the Baltic Sea are getting attention in action programs, e.g. the HELCOM Baltic Sea Action Plan (BSAP) [31]. The studies this thesis is based upon have also contributed to efforts to eluci-date pollution in the Baltic Sea by increasing our understanding of sources of PCDD/Fs and suggesting management strategies for further reducing its levels of PCDD/Fs.

(15)

CHAPTER 1. INTRODUCTION

Table 1.1 – Toxicity equivalent factor (TEF) values for 2,3,7,8-substituted CDD/F congeners.

Congener EPA1987a EPA1989a WHO1998b WHO2005b I-TEF1989c Nordic1987c

PCDDs TD 1 1 1 1 1 1 PeD 0.5 0.5 1 1 0.5 0.5 HxD1 0.04 0.1 0.1 0.1 0.1 0.1 HxD2 0.04 0.1 0.1 0.1 0.1 0.1 HxD3 0.04 0.1 0.1 0.1 0.1 0.1 HpD 0.001 0.01 0.01 0.001 0.01 0.01 OD 0 0.001 0.0001 0.0003 0.001 0.001 PCDFs TF 0.1 0.1 0.1 0.1 0.1 0.1 PeF1 0.1 0.05 0.05 0.03 0.05 0.01 PeF2 0.1 0.5 0.5 0.3 0.5 0.5 HxF1 0.01 0.1 0.1 0.1 0.1 0.1 HxF2 0.01 0.1 0.1 0.1 0.1 0.1 HxF3 0.01 0.1 0.1 0.1 0.1 0.1 HxF4 0.01 0.1 0.1 0.1 0.1 0.1 HpF1 0.001 0.01 0.01 0.01 0.01 0.01 HpF2 0.001 0.01 0.01 0.01 0.01 0.01 OF 0 0.001 0.0001 0.0003 0.001 0.001 References: a, [9]; b, [6, 7]; c, [8] TF: 2,3,7,8-TCDF. PF1: 1,2,3,7,8-PeCDF. PF2: 2,3,4,7,8-PeCDF. HxF1: 1,2,3,4,7,8-HxCDF. HxF2: 1,2,3,6,7,8-HxCDF. HxF3: 1,2,3,7,8,9-HxCDF. HxF4: 2,3,4,6,7,8-HxCDF. HpF1: 1,2,3,4,6,7,8-HpCDF. HpF2: 1,2,3,4,7,8,9-HpCDF. OF: OCDF. TD: 2,3,7,8-TCDD. PD: 1,2,3,7,8-PeCDD. HxD1:

1,2,3,4,7,8-HxCDD. HxD2: 1,2,3,6,7,8-HxCDD. HxD3: 1,2,3,7,8,9-HxCDD. HpD: 1,2,3,4,6,7,8-HpCDD. OD: OCDD.

1.2 Objective

The overall aim of the studies this thesis is based upon was to identify historical and current sources of PCDD/Fs in the Baltic Sea by examining sediment, biotic and air matrices. The thesis includes the following four papers, which are referred to in the text by the corresponding Roman numerals.

• Paper I: This aim of this paper was to increase understanding of spatial and temporal patterns of PCDD/F distributions in the Baltic Sea during the 20th century obtained from analyses of sediment cores.

• Paper II: The aims of this study were to identify the main historical sources of PCDD/Fs in the Baltic Sea and to characterize spatiotemporal

(16)

CHAPTER 1. INTRODUCTION

changes in their impacts.

• Paper III: The key aims here were to identify the main sources of PCDD/Fs in fish (Baltic herring) and to detect probable reasons why PCDD/F levels in Baltic biota have not declined significantly since the 1990s.

• Paper IV: The aim of this study was to investigate the potential of using metals as markers in tracing atmospheric PCDD/Fs sources in Baltic air.

(17)

Chapter 2

MATERIALS AND METHODS

2.1 Sampling and analysis of sediment cores

Sediment cores were collected from 15 Baltic Sea sampling stations (six offshore and nine coastal, designated O and C, respectively; Figure 2.1 and Paper I). The offshore stations are located along a north-south transect of the Baltic Sea and belong (among others) to the National Swedish Status and Trends Monitoring Program (NSSTMP) [41]. All the coastal stations are located on the Swedish coast of the Baltic Sea, where there is a long history of industrial activities related to forestry and cellulose processing (except at the coastal reference station C2). Geological Survey of Sweden (SGU) staff collected the offshore cores from the survey vessel S/V Ocean Surveyor (Figure 2.2) in 2008 (O1-O4 and O6) and 2010 (O5). The coastal sediment cores were sampled from the research vessel R/V Sunbeam during May and June 2010, after examining the condition of local surface sediment using an under-water video camera. The coastal sediment cores were sampled from R/V Sunbeam during May and June 2010, after checking that the surface had not been recently disturbed. Undisturbed coring sites deemed to be representative of large seabed areas were selected and characterized using hydro-acoustic instruments, including a side scan sonar (chirp) system, sediment sub-bottom profiler (3.5/7.5 kHz) and echo-sounder (35/200 kHz).

At all six offshore sampling stations (each of which is circular, with a 50 m radius), a master site in the center and another six coring sites were randomly chosen [41]. At each coastal station, only one site was selected for coring. Sediment cores 80 cm long were collected using a Gemini Corer, a twin-barrel

(18)

CHAPTER 2. MATERIALS AND METHODS

Figure 2.1 – Locations of the coastal (C1-C9) and offshore (O1-O6) sediment sampling stations in the Baltic Sea.

Niemistö Corer [42], then immediately x-rayed on board the research vessel in an upright position. One of each pair of cores collected from the twin barrel was dissected into two parts from top to bottom and photographs of the cuts were visually examined. The other core of each pair was cut horizontally into 1 cm thick segments for dating and 2-3 cm thick segments, which were stored immediately at -18oC until further (PCCD/F) analysis.

The sediment cores were dated by a commercial laboratory (Flett Research Ltd. Canada) and by SGU experts using Cs-137 and Pb-210 methods as well as by studying laminations of the cores. The Cs-137 method involves calibration using the Cs-peak generated by the Chernobyl accident in 1986 [43], and the assumption that the rate of sedimentation has been constant for each core. Lamination was also used to support the dating obtained using the Cs-137 method. For a few cores, (C1-C3) the Cs-137 method could not be used because the Chernobyl peak could not be identified with sufficient certainty. Consequently, these cores were dated using Pb-210 and the Constant Rate of Supply (CRS) model [44,45]. In the clean-up procedure prior to the PCDD/F analyses, the sediment samples were Soxhlet-extracted and cleaned using mul-tilayer silica and activated carbon columns [46]. All tetra-octa substituted congeners were then identified [47] and quantified by gas chromatography-high resolution mass spectrometry (GC-HRMS) using 60 m DB-5 columns.

(19)

CHAPTER 2. MATERIALS AND METHODS

Figure 2.2 – S/V Ocean Surveyor, a survey vessel owned by the Geological Survey of Sweden (SGU).

In order to ensure robust identification of the individual tetra-ortho-CDD/Fs, a signal/noise (S/N) threshold of 3 was applied. PCCD/F levels in blanks were generally <10% of levels in the sediment samples, so no blank corrections were applied. For concentrations that were below the limit of detection (<LOD), half of the corresponding detection limits (LOD/2) were used in subsequent analyses.

2.2 Air sampling and analysis

The air samples for PCDD/Fs analysis were collected at Aspvreten (a national air monitoring station situated in a coastal, remote area in central Sweden), using continuously operating high-volume samplers during August-September in 2010 and November-February in 2010/2011. The samplers were equipped with glass fiber filters to collect PCDD/Fs associated with atmospheric particles and PUFs to collect gaseous PCDD/Fs on a 24 h basis. Parallel air samples were also collected for metal analysis using a Derenda PNS 16T (Comde-Derenda GmbH, Germany) sequential sampling system, listed as a reference device for collecting PM10 and PM2.5 samples in the European Guidelines CEN 12341 and CEN 14907, respectively [48]. In total, 45 summer and 58 winter samples were collected, from which 18 and 22 samples, respectively, associated with stable air mass back trajectories at 20 m, 100 m and 500 m heights (according

(20)

CHAPTER 2. MATERIALS AND METHODS

to the NOAA HYSPLIT model [49]) were selected and analyzed for PCDD/Fs by Ökometric GmbH (Bayreuth, Germany) using GC-HRMS. Metals (As, Cd, Co, Cr, Cu, Fe, Hg, K, Mn, Ni, Pb, Sb, Se, Tl, V, Zn) were analyzed by ALS Scandinavia AB (Luleå, Sweden) according to USEPA 200.7 (ICP-AES) and 200.8 (ICP-SFMS) methods [50, 51].

2.3 Baltic herring

Levels of 2,3,7,8-CDD/Fs in Baltic herring (Clupea harengus) in the Bothnian Sea during the period 1980-2010, and in the Bothnian Bay and Baltic Proper during the period 1990-2010, were obtained from a previous study [38]. The herring were collected with the Swedish Monitoring Program administrated by the Swedish Museum of Natural History [52, 53] and were analyzed for PCDD/Fs at Umeå University. Pooled samples from 10-12 herring speci-men were used to calculate annual mean concentrations. In order to esti-mate sediment-to-herring Transformation Indices (TI; Paper III) for individual 2,3,7,8-CDD/Fs, different sets of data on their levels in herring [54] and surface sediments [46] were used.

(21)

Chapter 3

Data Analysis

3.1 Positive matrix factorization (PMF)

Positive matrix factorization (PMF) is a modeling technique developed during recent years (mid 1990s) and used to study and apportion different pollutant sources to receptors, for example, air, sediments or water bodies, as illustrated in Figure 3.1 [55, 56]. PMF was developed by Paatero and colleagues in the 1990s [57], and the key source-receptor relationship can be summarized as follows: xij = p X k=1 gikfkj+ eij

where p is the number of contributing sources; xij is the concentration of

the jth species in the ithsample; gikis mass concentration from the

kth source contributing to the jthsample; f

kjis the mass fraction of

the kth source to the jth sample; and e

ijis a model residual of xij.

For a dataset of n samples (objects) and m species (variables), the above equation can be generalized as follows:

X = GF + E

where X is a dataset of measurements of m species in n samples, G is a matrix of source contributions (n x p), F is a matrix of source compositions (p x m) and E is the residual matrix of X (n x m).

(22)

CHAPTER 3. DATA ANALYSIS

Figure 3.1 – Illustration of apportionment of pollutant sources in a marine system using receptor modeling. A, B and C represent various chemical species, e.g., congeners of a group of compounds such as dioxins.

The object function Q is then defined as follows:

n X i=1 m X j=1 x ij Ppk=1gikfkj uij 2

where uij is the uncertainty associated with each data point in matrix X.

PMF is thus a weighted least square problem that minimizes the object function Q with respect to G and F under the constraint that all elements of G and F must be non-negative. The imposition of non-negative constraints on elements of G and F reduces the rotational freedom, and ensures that the solutions are physically meaningful, as no source contributes negative masses to a receptor or emits non-positive fractions of species. PMF is not data-sensitive due to the scaling of each data point by its measurement uncertainty. Therefore, data points that are below LOD, missing or outliers can be included by reducing their weights (increasing their uncertainties). In Paper II, concentrations that were <LOD were replaced by half-LOD (LOD/2), and uncertainties corre-sponding to the measurement data were calculated as follows [58]:

• U = 5

6×LOD (if concentration is  LOD)

• U = LOD + 0.1×Concentration (if concentration is > LOD)

3.2 Principal component analysis (PCA)

Principal component analysis (PCA) is a technique widely applied to reduce dimensionality of a large multivariate dataset [59,60] and used for explanatory

(23)

CHAPTER 3. DATA ANALYSIS

data analysis such as to observe patterns, relationships, trends, and outliers in the data. PCA projects a dataset to a new coordinate system by determin-ing the eigenvectors and eigenvalues of a matrix, thus PCA is an orthogonal linear transformation where all basis vectors are an orthonormal matrix. This involves determining covariance matrix first to maximize the variance. The covariance of two random variables is a quantitative measure of their tendency to vary together. Principal components (PCs) that are linear combinations of the original variables are the axes of the new coordinate system. First, the covariance matrix, and the eigenvectors and eigenvalues of the covariance matrix are determined. A covariance of two variables can be calculated as:

cov(X, Y ) = n X i=1 (xi x)(y¯ i y)¯ n

where xiis the ithelement of X, yiis the ithelement of Y, ¯x is the mean of

X, ¯y is the mean of Y and n is the number of random measurements. After the eigenvectors (unit vectors) and the eigenvalues are calculated from covariance matrix, the eigenvalues are sorted in order of significance, i.e. the eigenvector with the highest eigenvalue to smallest. Eigenvectors with signifi-cant eigenvalues will be retained so that the new data matrix becomes a reduced version of the original dataset with fewer dimensions while retaining most of the variation. Principal components are calculated by multiplying each row of the eigenvectors with the sorted eigenvalues. Singular value decomposition (SVD), explained elsewhere [59–62], is commonly used for finding eigenvectors and eigenvalues.

3.3 Spline smoothing

A ‘spline’ is a function that is constructed piece-wise to allow a user to design smooth shapes that closely follows a data series. Spline smoothing is achieved by using a different polynomial curve between pieces of data points allowing continuity of the polynomial segments at their joints [63]. Cubic polynomial (polynomial function of order 3) is the function that is most typically chosen for constructing smooth curves. In this thesis, spline smoothing was used in Paper I for estimating trends in time series that could not approximated with other polynomial functions as illustrated in Figure 3.2.

(24)

CHAPTER 3. DATA ANALYSIS

Figure 3.2 – Illustration of data fitting using second polynomial (green line), cubic polynomial (black line) and spline function (blue line). The data series are from station O6 (total levels of PCDD/Fs in sediments).

(25)

Chapter 4

PCDD/F dynamics in the Baltic

Sea

4.1 Spatial and temporal trends of PCDD/Fs in the Baltic

Sea by the Swedish coast

Emissions of PCDD/Fs can be traced back as far as we have measured (1920s) [18], but first became strong during the 1940s and the post-war economic boom [64]. Sediments act as sinks and reservoirs of organic pollutants (such as PCDD/Fs and PCBs) in marine environments [65–67]. The current knowledge of PCDD/Fs time trends in Baltic Sea is limited. A single sediment core from an offshore area in the central sub-basin of the Baltic Proper covering 1882-1985 were studied about a couple of decades ago [37]. It was found that the level peaked during late 1970s. Similarly, subsurface peaks of PCDD/Fs and declining trends have also been observed in sediment cores from the Gulf of Finland [34,36]. In Paper-I, the temporal variations of PCDD/Fs in the Baltic Sea using sediment cores as historical records were studied, and spline curves were used to model the temporal trends at each of the stations. In accordance with expectations, the sediment records showed that PCCD/Fs began to rise in the Baltic Sea during the 1940s-1960s (Figure 4.1). They subsequently peaked during 1966-1986, with an overall peak in 1975 (± 7 years). The peak at the coastal reference station (C2) occurred later (1983-1988). These findings are consistent with previously reported peak years in sediments in both Europe (1960s-1980s [68–70]) and the USA (1960s-1970s [28, 71, 72]).

(26)

CHAPTER 4. PCDD/F DYNAMICS IN THE BALTIC SEA

Figure 4.1 – Spline lines of temporal changes in PCDD/F concentrations (Ptetra octa P CDD/F s (pg g 1dw)) in coastal sediment cores (stations C1 to C9) and confidence interval at significance level of 0.05.

In the coastal samples, the peak concentrations were generally in the range 1,400 - 14,000 pg g-1 dry weight (dw). However, at two coastal Bothnian Sea

sites (C8 and C9), the peak levels had been extremely high (up to 47,000 and 17,000 pg g-1dw, respectively). Such hotspot areas may be important for

long-term remobilization of PCDD/Fs in the Baltic system. High concentrations of PCDD/Fs (43,000 pg g-1dw) in surface sediments from the Bothnian Sea have

been previously reported [46]. The magnitude of the concentrations strongly suggests that the PCDD/Fs originated from point sources, and detailed source-tracing analyses are clearly required to elucidate their origins.

4.2 Spatial and temporal trends of PCDD/Fs in offshore

Baltic Sea areas

PCCD/F levels in all of the monitored offshore areas generally followed similar temporal trends, peaking 10 to 20 years later than in the coastal areas, during the period 1985-1994, except in Bornholm Bay (O5; where levels peaked in

(27)

CHAPTER 4. PCDD/F DYNAMICS IN THE BALTIC SEA

Figure 4.2 – Spline lines of temporal changes in PCDD/F concentrations (Ptetra octa P CDD/F s (pg g 1dw)) in offshore sediment cores (stations O1 to O6) and confidence interval at significance level of 0.05.

2001-2006, Figure 4.2). There were no differences in PCDD/F levels between stations in the Northern Baltic Sea (Bothnian Bay and Bothnian Sea) and Southern Baltic Sea (Baltic Proper and Arkona Basin). The overall peak for the open-sea sediments was in 1991 (± 5 years). The time lag between peaks in the coastal and offshore areas appears to have been due to slow transportation of PCDD/Fs to offshore regions. The peak levels were higher in the Southern than in the Northern Baltic Sea (1,200-2,700 pg g-1 dw and

590-1,100 pg g-1dw, respectively). Assuming that atmospheric transportation

is the main contributor of PCDD/Fs to the Baltic Sea, as concluded by various authors [34, 35, 39, 73], the higher levels in the Southern Baltic Sea can be explained by its proximity to Central Europe (the most polluted region in Europe).

4.3 Environmental half-lives

As summarized above, PCDD/F levels in Baltic Sea sediments have declined from peak concentrations several decades ago. The environmental half-lives were estimated from the spline curves, and found to be 6-23 years in the

(28)

CHAPTER 4. PCDD/F DYNAMICS IN THE BALTIC SEA

coastal areas if excluding C2, the reference site with no known point sources in the near area (Table 4.1). The observed reductions are attributable to local and international programs and measures imposed since the 1970s to eliminate, or at least minimize, emissions of environmental pollutants (e.g. the ban on use of PCP in the late 1970s and early 1980s [29], HELCOM [31], and the Stockholm Convention [30]). The short environmental half-lives in coastal regions presumably reflect rapid responses following the closure of point sources. The coastal lives are also comparable to previously reported half-lives in air and sediments affected by urban and industrial emissions (9 years in sediments from an urban lake [74], and 3-11 years in urban air [75, 76]). The overall half-life in the coastal regions was found to be 15 ± 5 years. Consequently, PCDD/F levels have declined by 81 ± 12% from peak levels to levels observed in the most recent layers, corresponding to 2009/2010. However, environmental half-lives have been longer in the open-sea sediments (13-50 years; overall half-life 29 ± 14 years), and there has only been a 38 ± 11% decline from peak levels. No differences in environmental half-lives between the Northern and Southern Baltic Sea were detected, and the apparently slow recovery of the open-sea system shows that more effort is needed to further reduce levels of PCDD/Fs in the Baltic Sea.

(29)

CHAPTER 4. PCDD/F DYNAMICS IN THE BALTIC SEA

Table 4.1 – Summary of peak concentrations of PCCD/Fs (unit), peak years, environmental half-lives (units), and percentage decline from peak year.

station period peak year peak level reduction (%) half-life

Coastal (pg g-1 dw) (years) C1 1971−2010 1984-1985 3900 94 6 C2 1902−2010 1983-1988 630 31 41 C3 1866−2010 1962-1969 5000 77 20 C4 1969−2008 1980-1984 1400 57 23 C5 1961−2009 1975-1978 1700 75 17 C6 1960−2009 1969-1972 1700 89 11 C7 1925−2009 1966-1969 14000 88 15 C8 1970−2010 1970-1972 47000 84 15 C9 1973−2009 1978-1980 17000 86 12 Mean ± SD 1975 (± 7 years) 81±12 15±5 Offshore O1 1943−2007 1992-1996 1100 52 13 O2 1941−2008 1984-1986 860 47 22 O3 1956−2008 1992-1995 590 27 32 O4 1971−2002 1985-1987 2700 36 26 O5 1941−2010 2001-2006 1900 39 NA O6 1921−2007 1982-1987 1200 24 50 Mean ± SD 1991 (± 5 years) 38±11 29±14

(30)
(31)

Chapter 5

Sources of PCDD/Fs in the Baltic

Sea

5.1 Historical sources in sediments

5.1.1 Source identification

PCDD/Fs are predominantly produced unintentionally as byproducts of ther-mal processes (such as MSWI and metal smelter processes) and other processes (such as pulp and paper bleaching) [15, 30, 71, 77–92]. The relative impor-tance of these sources varies in different environments. Thus, identifying and understanding the main sources in a particular environment are important elements of emission management. In Paper II, historical sources of PCDD/Fs in the Baltic Sea were investigated using PCDD/F congener patterns (also known as source fingerprints [93–98]) in sediment cores and the PMF modelling technique. This modelling technique has previously been applied for tracing PCDD/F sources in surface sediments from the Baltic Sea [73], surface sedi-ments from Ichihara Anchorage in northeastern Tokyo Bay, Japan [97], in biota from Dalian, China [99] and PAHs in sediments from Taihu Lake, China [100]. In the Baltic Sea sediment core study (Paper II), five PCDD/F congener patterns representing emission sources were extracted, compared with known source fingerprints, and six sources were identified. These sources are atmo-spheric background (AB), thermal processes (Thermal), use and production of tetra-chlorophenol (TCP), use and production of penta-chlorophenol (PCP),

(32)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

hexa-CDD-related sources (HxCDD) and elementary chlorine-related sources (Chl).

The PCDD/F pattern of AB, which is almost completely dominated by OCDD [73, 101–103], is generally similar to the PCP pattern (Figure 5.1). In addition to OCDD, the PCP pattern includes considerable fractions of 1,2,3,4,6,7,8-HpCDF, 1,2,3,4,6,8,9-HpCDF and OCDF [73, 104–106], but the differences in AB and PCP patterns have been too small for PMF to distinguish between them. Before being banned in the 1970s, pentachlorophenol technical products (e.g. Dowicide) had been used in sawmills and forestry industries for wood preservation [107–109]. Tetrachlorophenol technical products (e.g. Ky-5) had also been used as wood preservatives [110–113]. PCDD/F con-tamination from TCP is characteristically dominated by 1,2,3,4,6,7,8-HpCDF, 1,2,3,4,6,8,9-HpCDF and OCDF congeners and low or no contribution from PCDDs [73, 96, 114].

The Chl- and HxCDD-related sources are associated with cellulose indus-tries. The PCDD/F pattern associated with bleaching of pulp using elementary chlorine is generally dominated by lightly-chlorinated PCDFs [17, 73, 96, 115]. The HxCDD source type pattern can be linked to cellulose industries after use of chlorine gas was abandoned in the 1980s [73]. Unsurprisingly, thermal processes (example MSWI, smelters) have also been identified as major his-torical sources of PCDD/Fs in the Baltic Sea. The associated pattern such as of has been well documented and is characterized by high ratios of PCDD/Fs to PCDDs [73, 81, 85, 116, 117]. The source identification process based on the PCDD/F fingerprints of the sources and marker congeners is discussed in more detail in Paper II and [73].

5.1.2 Source apportionment

Both absolute and relative impacts of each of the identified sources were stud-ied. The combined impacts of the two source types linked to atmospheric emissions, i.e. Thermal and AB/PCP, are and have been most important in the Southern Baltic Sea (Baltic Proper, Bornholm Basin and Arkona Basin; O4-O6), where they appear to have contributed >80% of the total amount of PCDD/Fs (Figure 5.2 red and pink colors). The proximity of the Southern Baltic Sea to the highly populated regions of Central Europe may explain the strength of atmospheric sources relative to local sources. Local sources (TCP, Chl, HxCDD; blue, brown, green) made a much stronger contribution in the

(33)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Figure 5.1 – Model PCDD/F fingerprints that have been identified and assigned after comparison to known sources. Data shown are proportions of total PCDD/F concentrations (0-1).

Northern Baltic Sea (Bothnian Bay and Bothnian Sea; O1-O3), accounting for on average ca. 50% of the total PCCD/F load, compared to <18% in the Southern Baltic. Among the local sources, Chl had minimal impacts (2.0 ± 1.8%) whereas HxCDD had pronounced effects in the Bothnian Bay (15 ± 12%). Interestingly, TCP appeared to be the strongest source in the Bothnian Sea (26 ± 14%), where levels of PCDD/Fs in herring and salmon are occasionally high. This supports a possible link between the local PCDD/F sources and biota, as contamination of biota is pronounced in the Northern Baltic Sea [54, 118, 119].

As discussed in previous sections, PCCD/F patterns at coastal stations generally reflect the impact of local sources in the vicinity of the sediment core sampling points (Figure 5.3). For example, the apparent dominance of Thermal sources at station C3 can be explained by the proximity of a large

(34)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Figure 5.2 – Concentrations (pg g-1dw) of the five identified source types and residuals in dated sediment cores from the open sea system of the Baltic Sea.

metal production plant (Rönnskär Smelter) that has been operating since 1930 [120, 121]. PCP sources probably account for most of the dominance of AB/PCP sources in the Southern Bothnian Sea (specifically at stations C8 and C9), because AB is unlikely to have contributed strongly to the extremely high PCDD/F levels (up to 8,100 pg g-1dw) detected at these sites. Generally, use

of chlorophenol (TCP and PCP) has been the two most important sources of PCDD/Fs in both coastal and offshore areas of the Bothnian Sea and Bothnian Bay. In conclusion, the analyses showed that atmospheric emission contributed most of the PCDD/Fs in the Southern Baltic Sea sediments, whereas local sources (mainly chlorophenols) have been important in the Northern Baltic Sea. In Sweden, the use of chlorophenols had been banned for four decades, and the activity at most of these sites was ceased long before the sampling period. Thus, detailed assessment of previously contaminated sites is required to understand their possible links with the contamination of Baltic biota.

(35)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Figure 5.3 – Concentrations of the identified PCDD/F source types in dated coastal sediment cores in the Baltic Sea.

5.2 Historical sources of PCCD/Fs in biota

5.2.1 Transformation indices (TI)

As previously mentioned, PCDD/F levels in edible fatty fish from the Baltic Sea (e.g. herring and salmon) have raised health concerns [39]. Average concen-trations in Baltic herring caught in the Bothnian Bay, Bothnian Sea and Baltic Proper during the period 1990-2010 were 38, 95 and 48 WHO2005-TEQ pg g-1

lipid weight (lw) respectively [38]. The PCDD/F levels measured in sediments (section 4.2) suggest that biota should be most strongly contaminated in the Baltic Proper. However, the contamination is more pronounced in Bothnian Sea biota, indicating that specific local sources might be tightly linked to contamination in biota. Hence, in the study reported in Paper III, sources of PCDD/Fs in biota were examined using data on temporal changes in their levels in Baltic herring (collected by the Swedish Museum of Natural History

(36)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

herring as part of the Swedish Monitoring Program [52, 54]) and the PMF technique.

Identifying sources of PCDD/Fs in biota using chemical fingerprints of known sources is challenging. This is mainly due to differences in relative bioavailability and metabolism among PCDD/F congeners, which result in changes in their patterns in biota relative to the pattern in the surrounding environment [122–126]. Hence, direct comparison of PCDD/F patterns in biota and potential sources is not informative. Thus, to compensate for the transformations, TI values (Table 5.1) were estimated for the individual 2,3,7,8-substituted congeners by calculating transformation factors as ratios of con-centrations in sediments to concon-centrations in herring using spatially matched sediment and herring samples [46, 52]. Previously, Hebert and co-workers applied the TI concept by assuming particulate matters as starting point for entry into the food web and calculating bioavailability factors from sediment to fish and biomagnification factors (small fish, i.e. Smelt, to herring) [126]. TI values were then estimated by normalizing the bioavailability factors and biomagnification factors to the corresponding values of TCDD and multiplying the two normalized factors together. Similarly, in Paper-III TI values were then obtained by normalizing the transformation factors (Table 5.1) to that of TCDD. In addition, to minimize effects of ecological differences between the sub-basins that may cause variations in behavior of PCDD/Fs [127–129], specific TI values were estimated for each sub-basin. However, the values for the Bothnian Bay, Bothnian Sea and Baltic Proper were found to be generally similar (Table 5.1).

5.2.2 Sources and temporal changes in their impacts

Three model factors representing candidate PCDD/F sources in biota were extracted by PMF. Then, the expected (corresponding) patterns in sediments were calculated using TI values (Figure 5.4) and compared to patterns of known sources to identify likely sources (Figure 5.5). Thermal, TCP and PCP (three of the five main PCDD/F sources identified as impacting the Baltic system; Paper II) were identified as important sources for biota. The source apportioned herring time trend data showed that the relative importance of Thermal sources in biota has been declining, with half-lives of 8-10 years (Figure 5.6) in all three of the sub-basins (i.e. Bothnian Bay, Bothnian Sea, and Baltic Proper). This agrees well with the half-life of decline in atmospheric PCDD/F levels,

(37)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Table 5.1 – Transformation factors and transformation indices (TI) for indicated PCCD/Fs in the Bothnian Bay, Bothnian Sea and Baltic Proper.

congener Transformation factors Transformation index

BB BS BP BB BS BP TF 0.52 0.27 0.08 4.6 2.0 2.2 PF1 0.48 0.31 0.10 4.2 2.2 2.5 PF2 2.3 1.6 0.41 20 12 11 HxF1 0.04 0.04 0.01 0.35 0.29 0.30 HxF2 0.15 0.14 0.03 1.3 1.0 0.86 HxF3 0.12 0.12 0.03 1.1 0.87 0.86 HxF4 0.08 0.05 0.03 0.73 0.38 0.70 HpF1 0.00 0.00 0.00 0.02 0.02 0.03 HpF1 0.03 0.00 0.01 0.22 0.03 0.14 OF 0.00 0.00 0.00 0.02 0.01 0.02 TD 0.11 0.14 0.04 1.00 1.00 1.00 PD 1.1 0.81 0.27 9.9 5.9 7.2 HxD1 0.00 0.00 0.00 0.00 0.00 0.00 HxD2 0.22 0.20 0.05 1.9 1.4 1.3 HxD3 0.02 0.03 0.01 0.20 0.19 0.22 HpD 0.01 0.00 0.00 0.06 0.03 0.04 OD 0.02 0.01 0.00 0.13 0.05 0.06 TF: 2,3,7,8-TCDF. PF1: 1,2,3,7,8-PeCDF. PF2: 2,3,4,7,8-PeCDF. HxF1: 1,2,3,4,7,8-HxCDF. HxF2: 1,2,3,6,7,8-HxCDF. HxF3: 1,2,3,7,8,9-HxCDF. HxF4: 2,3,4,6,7,8-HxCDF. HpF1: 1,2,3,4,6,7,8-HpCDF. HpF2: 1,2,3,4,7,8,9-HpCDF. OF: OCDF. TD: 2,3,7,8-TCDD. PD: 1,2,3,7,8-PeCDD. HxD1: 1,2,3,4,7,8-HxCDD. HxD2: 1,2,3,6,7,8-HxCDD. HxD3: 1,2,3,7,8,9-HxCDD. HpD: 1,2,3,4,6,7,8-HpCDD. OD: OCDD.

3-11 years (Paper II), providing good indications that biota are responding in accordance with the decline in atmospheric levels of PCDD/Fs. However, the relative importance of TCP and PCP was shown to have been slowly declining or even increasing, in all sub-basins, except for PCP in the Bothnian Sea (Figure 5.6). These temporal trends in contributions of TCP and PCP sources may account for the steady-state or increasing levels of PCDD/Fs in Baltic herring [38, 39].

The observed temporal trends suggest that the contamination of Baltic biota might be tightly linked to a few local sources. We suggest that source-specific bioavailability could be a possible explanation for differences in sources’ relative importance to marine biota as the extent of bioavailability depends mainly on the amount of associated black carbon (BC, soot) in the emission

(38)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Figure 5.4 – Normalized PCDD/F patterns. Left panels show modeled (PMF) congener patterns of factors 1-3. Middle panels show transformed factors, i.e. expected patterns in sediments obtained using calculated Transformation Indices. Right panels show source PCDD/F patterns. These include the sediment thermal pattern from Paper-II; TCP patterns — TCP pattern 1 from contaminated sediment (Paper II), TCP pattern 2 from Ky-5 (ca. 80% TCP) [73], and TCP pattern 3 from Ky-5 [130]; and PCP source patterns — PCP pattern 1 from contaminated soil [105], PCP pattern 2 from contaminated river sediments [105], PCP pattern 3 from contaminated coastal sediments [105], and PCP patterns 4 & 5 from commercial PCP [73].

[131–135]. For example, PCCD/Fs from Thermal sources are likely to be more associated with BC than PCCD/Fs from chlorophenol-related sources. As described above, in both Sweden and internationally the manufacture and use of chlorophenols stopped in the early 1980s. Therefore, these findings call for detailed study on between-source variations in bioavailability of pollutants and the importance of abandoned contaminated sites for the marine environment.

(39)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Figure 5.5 – PMF Modeled 2,3,7,8-PCDD/F source patterns in sediments reported in Paper II.

5.3 Atmospheric sources

5.3.1 PCDD/F source regions, congener patterns and seasonality

Long-range atmospheric transport of POPs has played a key role in the global distribution of pollutants including PCDD/Fs [2, 5, 136, 137]. Consequently, POPs have been detected in remote and pristine environments. Atmospheric inputs have also been the leading contributors of PCDD/Fs in certain polluted environments, including the Baltic Sea [Paper II, [35, 73, 138]]]. Atmospheric deposition accounted for >80% and ca. 50% of the PCDD/Fs detected in the Southern and Northern Baltic Sea sediments, respectively (Paper II). The source regions for atmospheric emissions of PCDD/Fs in Northern Europe have been previously studied [101,138,139]. Stable back trajectories (as illustrated in Figure 5.7) can be used to identify the origins of air masses. Areas to the south

(40)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA Fi gu re 5. 6 –S ou rc ea pp or ti on ed( P M F-m od el ed )t em po ra lt re nd so f P C D D /Fc on ce nt ra ti on si nB al ti ch er rin g (pg WHO 2005 -T EQ g -1 lw ) or iginat ing fr om the thr ee ident ified P C DD/F sour ces (T her m al, T C P, and P C P ) and tot al concent rat ions.

(41)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

and south-east (Central and Eastern Europe, respectively) have been reported as the main current source regions for atmospheric PCDD/Fs in the Baltic Sea [138,139]. This appears to be a shift from the situation three decades ago, when the main source areas were to the west and south-west [101]. Moreover, relative contributions of PCDFs to total PCDD/F loads have also increased, e.g. the PCDF/PCDD ratio has increased from 0.3-1.0 to 1.0-1.4.

The atmospheric levels of PCDD/Fs in Baltic air have been shown to exhibit marked seasonal variation, with dramatic increases during the winter season as a result of increased levels of PCDFs [138]. This was also concluded in Paper IV, with a winter mean concentration of 5.1 fg WHO2005-TEQ m-3, and

a summer mean level of 0.26 fg WHO2005-TEQ m-3. The winter atmospheric

levels were most pronounced in the SSE-E sector (14 fg WHO2005-TEQ m-3).

High levels during winter indicate that domestic heating and production of power for heating likely are important sources. The PCDD/F patterns of the four compass sectors (i.e. S, E, N and SW) during winter and summer seasons are shown in Figure 5.8. The high levels in the air were associated with an increased fraction of PCDFs, in accordance with the previous study [138], and this pattern matches Thermal sources previously identified in Paper II. It thus appears that elevated emissions of PCDD/Fs should be linked to thermal sources.

5.3.2 The potential to use metals as source markers

PCDD/F congener patterns alone will not help to understand the dioxin sources. Hence, the potential for using metals as source markers was explored. A correlation matrix study analysis was conducted, in which all elements and PCDD/F congeners were included except those that had a high fraction of missing data (>40% below LOD). The correlation matrixes showed that the total sum of PCDD/Fs was highly correlated to PCDFs (r>0.86). This shows that the furans (PCDFs) should receive special attention in the source tracing. Most of the detected metals were positively correlated with the PCDFs. The exceptions were Cr (r<0.22, p>0.24), Ni (r<0.55, p>0.01), and V (r<0.59, p>0.01). Metals that show significant correlation with the PCDFs (average r is 0.85; p<0.05) included Cd, Co, Cu, Fe, K, Mn, Pb, Sb, and Zn. A strong correlation between PCDFs and metals indicates a co-emission relation. However, the metals that correlated well represent a wide range of metals, which all together do not point at a single source type, not even when focusing

(42)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA Fi gu re 5. 7 – Ex am pl es of st ab le ai r-m as s tr aj ec to rie s ca lc ul at ed fr om of th e A sp vr et en ai r sa m pl in g du rin g w in te r se as on 20 10 /2 01 1.

(43)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Figure 5.8 – PCDD/F congener patterns of 2,3,7,8-substituted CDD/Fs in air from S, W, E and N sectors during summer and winter seasons. A color represents a single sample. on thermal sources only. A limiting factor for the back-tracking is the lack of an exhaustive, up-to-date compilation of metal emissions from various emission sources, not the least from thermal sources. In Table 5.2, some links between metal and source types have been listed. It has for example been reported that K, Zn, Pb and Cu are markers for biomass burning [140–143], and Zn and Pb for non-ferrous smelters [144], while corresponding information for emissions from coal combustion and municipal solid waste incineration (MSWI) are not easily accessible.

In conclusion, the introduction of parallel measurements of metals is a promising complement to seasonal trends and PCDD/F congener pattern anal-ysis. Several metals were found to be significantly correlated to the PCDFs, the congeners of interest in the current study. However, the wide range of candidate metals as source markers for PCDD/F emissions, and the lack of an up-to-date extensive compilation of source characteristics for metal emission from

(44)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Table 5.2 – A summary of marker metals for various source types.

Marker metals References

Medical Waste Incinerators Hg Wu et. al. 2011 [145] Ferrous industries Fe, Mn Morawska et. al. 2002 [146] Vehicle emission Pb Morawska et. al. 2002 [146] Oil burning V Morawska et. al. 2002 [146]

Biomass burning K Silva et. al. 1999 [143] and Hays et. al. 2005 [141] Non-ferrous smelters Zn, Pb Liu et. al. 2003 [144]

various sources, limits the use of the metals as source markers in the current study. Instead, the best leads for back-tracking primary sources of atmospheric PCDD/Fs in Baltic air are still the seasonal trends and the PCDD/F patterns. Although Paper IV was not able to pin-point primary PCDD/F sources for Baltic air, it has demonstrated a new promising approach for source tracing in air. Extensive PCDD/F-metal data sets (from field measurements) along with comprehensive, up-to-date data compilations of metal emissions from thermal sources will increase success potential, as then uncertainties will be lower and advanced data evaluation tools, such as multivariate statistics, could assist in revealing underlying patterns.

(45)

CHAPTER 5. SOURCES OF PCDD/FS IN THE BALTIC SEA

Figure 5.9 – Correlations between metals and PCDD/F congeners in samples during the winter.

(46)
(47)

Chapter 6

Conclusions and future work

Despite being one of the most polluted marine systems in the world, we have limited knowledge of PCDD/F dynamics in the Baltic Sea. It is known that several factors, such as ongoing direct and indirect releases of PCDD/Fs and long residence times (due to the slow exchange with open water), contribute to the accumulation of pollutants in the Baltic Sea. However, several important questions concerning the main sources of pollutants, the recovery of the Baltic system (abiotic and biotic), and the optimal emission management practices for easing the pollution loads require further attention.

Studies on sediment cores, which this thesis is partly based upon, revealed that PCDD/F concentrations in the Baltic Sea rose to high levels from the 1940s, following sharp rises in anthropogenic emissions. After peaks (in the 1970s and 1990s in coastal and offshore areas, respectively) there have been remarkable recoveries by the Swedish coast (ca. 80% reductions in levels) but slower recoveries (ca. 40% reductions) in offshore areas of the Baltic Sea. Generally, PCDD/F levels are higher in the Southern Baltic Sea than in the Northern Baltic Sea, indicating possible differences in relative contributions of sources in different regions of the sea. The results of the studies reported here show that both diffuse sources (atmospheric background and thermal) and local sources (use of tetrachlorophenol, pentachlorophenol and elementary chlorine) have substantially affected the Baltic Sea. Generally, diffuse sources have impacted the Southern more than the Northern Baltic Sea (accounting for >80% and ca. 50% of total loads, respectively). The local sources, which have strong links with the cellulose and forestry industries, have been more important in the Northern Baltic Sea, especially the Bothnian Sea.

(48)

CHAPTER 6. CONCLUSIONS AND FUTURE WORK

Source apportionment of PCDD/Fs in Baltic herring showed that, in con-trast to sediments, only thermal and chlorophenols (use of TCP and PCP) have been important source types for biota. The impact of thermal sources has declined at a similar rate (with a half-life of 8-10 years) to the decline observed in ambient air. However, contributions of TCP- and PCP-related sources have not been declining (or declining very slowly). Thus, they have contributed to the observed stability of PCDD/F levels in Baltic biota. Since the use of chlorophenols began to stop in the early 1980s, detailed studies are required to understand effects of land contaminated with these substances on the marine environment.

The possibility of back-tracing atmospheric sources of PCDD/Fs was also studied, with the main aim to investigate the potential of using metals as markers of source types as a complement to PCDD/F patterns and seasonal trends. Several metals were found to be significantly correlated to the PCDFs, the congeners of interest in the current study. However, the wide range of candidate metals as source markers for PCDD/F emissions, and the lack of an up-to-date extensive compilation of source characteristics for metal emission from various sources, limited the use of the metals as source markers Instead, the best leads for back-tracking primary sources of atmospheric PCDD/Fs in Baltic air were still the seasonal trends and the PCDD/F patterns. Although the air study could not pin-point primary PCDD/F sources for Baltic air, it demonstrated a new promising approach for source tracing in air.

(49)

Acknowledgments

A doctoral thesis is often described as a solitary endeavor; however the long list that follows definitely proves otherwise.

First and foremost, I am deeply grateful for the continuous support, insight and patience of my supervisors Karin Wiberg, Mats Tysklind, Paul Geladi and Patrik Andersson.

Karin, you offered me the opportunity in the first place to join and partic-ipate in your research group. Being here gave me the time and conditions to work on several issues and I cannot thank you enough for that and for helping me in every single step of this long educational process.

Mats, you are a wonderful person from whom I have learned a lot. I always looked forward to show you my new results and have discussions around them because it would guarantee that I am on the right track. There would always be new ideas or directions for development after each meeting.

Paul, I also want to thank you because you helped me to understand multivariate statistics. Patrik, I enjoyed our journal club and gained significant knowledge about in-silco tools.

I also would like to thank Terry Bidleman. You are a wonderful person I enjoyed discussions with. My gratitude also extends to Ingemar Cato, Michael McLachlan, Ulla Sellström, Anna Sobek, Kristina Sundqvist, Per Jonsson and several other people for collaborations and contributions.

Maria Hjelt and Per Liljelind (for the assistance in the analytical work), Stina Jonsson and other researchers and PhD students at Environmental Chem-istry group are also thanked.

Solomon, Kifle, Ewnetu, and Esubalew are also thanked for friendship. I gratefully acknowledge the following funding sources: the Swedish Envi-ronmental Protection Agency for funding BalticPOPs and a project adminis-trated by the County Board of Ga¨vleborg, and the Ecochange research pro-gram, funded by the Swedish Research Council for Environment, Agricultural Sciences and Spatial Planning (Formas) for general support to marine science at Umeå University.

(50)
(51)

Bibliography

[1] Arnold Schecter, Linda Birnbaum, John J Ryan, and John D Constable. Dioxins: an overview. Environmental research, 101(3):419–428, 2006. [2] Kevin C Jones and P De Voogt. Persistent organic pollutants (pops):

state of the science. Environmental Pollution, 100(1):209–221, 1999. [3] Ronald A Hites. Dioxins: An overview and history. Environmental

science & technology, 45(1):16–20, 2010.

[4] Frank Wania and Donald Mackay. Peer reviewed: tracking the distribution of persistent organic pollutants. Environmental Science & Technology, 30(9):390A–396A, 1996.

[5] Frank Wania. Assessing the potential of persistent organic chemicals for long-range transport and accumulation in polar regions. Environmental science & technology, 37(7):1344–1351, 2003.

[6] Martin Van den Berg, Linda S Birnbaum, Michael Denison, Mike De Vito, William Farland, Mark Feeley, Heidelore Fiedler, Helen Hakansson, Annika Hanberg, Laurie Haws, et al. The 2005 world health organization reevaluation of human and mammalian toxic equivalency factors for dioxins and dioxin-like compounds. Toxicological sciences, 93(2):223–241, 2006.

[7] Martin Van den Berg, Linda Birnbaum, AT Bosveld, Björn Brunström, Philip Cook, Mark Feeley, John P Giesy, Annika Hanberg, Ryuichi Hasegawa, Sean W Kennedy, et al. Toxic equivalency factors (tefs) for pcbs, pcdds, pcdfs for humans and wildlife. Environmental health perspectives, 106(12):775, 1998.

[8] Christoffer Rappe. Dioxin, patterns and source identification. Fresenius’ journal of analytical chemistry, 348(1-2):63–75, 1994.

[9] Linda S Birnbaum. Tefs: A practical approach to a real-world problem. Human and Ecological Risk Assessment: An International Journal,

(52)

CHAPTER 6. CONCLUSIONS AND FUTURE WORK

5(1):13–24, 1999.

[10] Richard E Peterson, H Michael Theobald, and Gary L Kimmel. Developmental and reproductive toxicity of dioxins and related compounds: cross-species comparisons. CRC Critical Reviews in Toxicology, 23(3):283–335, 1993.

[11] Mary K Walker and Richard E Peterson. Aquatic toxicity of dioxins and related chemicals. In Dioxins and Health, pages 347–387. Springer, 1994. [12] Kyle Steenland, Laurie Piacitelli, James Deddens, Marilyn Fingerhut, and Lih Ing Chang. Cancer, heart disease, and diabetes in workers exposed to 2, 3, 7, 8-tetrachlorodibenzo-p-dioxin. Journal of the National Cancer Institute, 91(9):779–786, 1999.

[13] Marilyn A Fingerhut, William E Halperin, David A Marlow, Laurie A Piacitelli, Patricia A Honchar, Marie H Sweeney, Alice L Greife, Patricia A Dill, Kyle Steenland, and Anthony J Suruda. Cancer mortality in workers exposed to 2, 3, 7, 8-tetrachlorodibenzo-p-dioxin. New England journal of medicine, 324(4):212–218, 1991.

[14] Ulrich Quaß, Michael Fermann, and Günter Bröker. The european dioxin air emission inventory project – final results. Chemosphere, 54(9):1319– 1327, 2004.

[15] Heidelore Fiedler. Sources of pcdd/pcdf and impact on the environment. Chemosphere, 32(1):55–64, 1996.

[16] Knut Breivik, Ruth Alcock, Yi-Fan Li, Robert E Bailey, Heidelore Fiedler, and Jozef M Pacyna. Primary sources of selected pops: regional and global scale emission inventories. Environmental Pollution, 128(1):3– 16, 2004.

[17] Christoffer Rappe, Lars-Owe Kjeller, Sten-Erik Kulp, Cynthia de Wit, Ingrid Hasselsten, and Ola Palm. Levels, profile and pattern of pcdds and pcdfs in samples related to the production and use of chlorine. Chemosphere, 23(11):1629–1636, 1991.

[18] Roland Weber, Mats Tysklind, and Caroline Gaus. Dioxin-contemporary and future challenges of historical legacies. Environmental Science and Pollution Research, 15(2):96–100, 2008.

[19] Ruth E Alcock, Andy J Sweetman, and Kevin C Jones. A congener-specific pcdd/f emissions inventory for the uk: do current estimates account for the measured atmospheric burden? Chemosphere, 43(2):183– 194, 2001.

(53)

CHAPTER 6. CONCLUSIONS AND FUTURE WORK

[20] Simon J Buckland, Howard K Ellis, and Patrick Dyke. New Zealand inventory of dioxin emissions to air, land and water, and reservoir sources. Organochlorines Programme, Ministry for the Environment, 2000.

[21] Jozef M Pacyna, Knut Breivik, Jörg Münch, and Janina Fudala. European atmospheric emissions of selected persistent organic pollutants, 1970–1995. Atmospheric Environment, 37:119–131, 2003.

[22] Hugo Denier van der Gon, Maarten van het Bolscher, Antoon Visschedijk, and Peter Zandveld. Emissions of persistent organic pollutants and eight candidate pops from unece–europe in 2000, 2010 and 2020 and the emission reduction resulting from the implementation of the unece pop protocol. Atmospheric Environment, 41(40):9245–9261, 2007. [23] Ulrich Quaß and Michael Fermann. Identification of Relevant

Industrial Sources of Dioxins and Furans in Europe:(the European Dioxin Inventory); Final Report. North Rhine-Westphalia State Environment Agency, 1997.

[24] M Wenborn, K King, D Buckley-Golder, and J Gascon. Releases of dioxins and furans to land and water in europe. Final Report. Report produced for Landesumwaltamt Nordrhein-Westfalen, Germany on behalf of European Commission DG Environment, 1999.

[25] UNEP Chemicals. Dioxin and furan inventories: national and regional emissions of PCDD/PCDF. UNEP, 1999.

[26] Ulrich Quaß, Michael W Fermann, and Günter Bröker. Steps towards a european dioxin emission inventory. Chemosphere, 40(9):1125–1129, 2000.

[27] V Vestreng and H Klein. Emission data reported to unece/emep. Quality assurance and trend analysis and Presentation of WebDab, MSC-W Status Report, 2002.

[28] Ruth E Alcock and Kevin C Jones. Dioxins in the environment: a review of trend data. Environmental Science & Technology, 30(11):3133–3143, 1996.

[29] Weiwei Zheng, Xia Wang, Huan Yu, Xuguang Tao, Ying Zhou, and Weidong Qu. Global trends and diversity in pentachlorophenol levels in the environment and in humans: a meta-analysis. Environmental science & technology, 45(11):4668–4675, 2011.

(54)

CHAPTER 6. CONCLUSIONS AND FUTURE WORK

[30] Heidelore Fiedler. National pcdd/pcdf release inventories under the stockholm convention on persistent organic pollutants. Chemosphere, 67(9):S96–S108, 2007.

[31] Hermanni Backer, Juha-Markku Leppänen, Anne Christine Brusendorff, Kaj Forsius, Monika Stankiewicz, Jukka Mehtonen, Minna Pyhälä, Maria Laamanen, Hanna Paulomäki, Nikolay Vlasov, et al. Helcom baltic sea action plan - a regional programme of measures for the marine environment based on the ecosystem approach. Marine pollution bulletin, 60(5):642–649, 2010.

[32] Mats Tysklind, Stellan Marklund, and Christoffer Rappe. Experiences from the swedish msw moratorium. Chemosphere, 16(8):1937–1942, 1987. [33] SE Swanson. Dioxins in the bleach plant. ThesisUniversity of Umeå,

1988.

[34] Matti Verta, Simo Salo, Markku Korhonen, Timo Assmuth, Hannu Kiviranta, Jaana Koistinen, Päivi Ruokojärvi, Pirjo Isosaari, Per-Anders Bergqvist, Mats Tysklind, et al. Dioxin concentrations in sediments of the baltic sea - a survey of existing data. Chemosphere, 67(9):1762–1775, 2007.

[35] James M Armitage, Michael S McLachlan, Karin Wiberg, and Per Jonsson. A model assessment of polychlorinated dibenzo-p-dioxin and dibenzofuran sources and fate in the baltic sea. Science of the total environment, 407(12):3784–3792, 2009.

[36] Pirjo Isosaari, Harri Kankaanpää, Jukka Mattila, Hannu Kiviranta, Matti Verta, Simo Salo, and Terttu Vartiainen. Spatial distribution and tem-poral accumulation of polychlorinated dibenzo-p-dioxins, dibenzofurans, and biphenyls in the gulf of finland. Environmental science & technology, 36(12):2560–2565, 2002.

[37] Lars-Owe Kjeller and Christoffer Rappe. Time trends in levels, patterns, and profiles for polychlorinated dibenzo-p-dioxins, dibenzofurans, and biphenyls in a sediment core from the baltic proper. Environmental science & technology, 29(2):346–355, 1995.

[38] Aroha Miller, Jenny E Hedman, Elisabeth Nyberg, Peter Haglund, Ian T Cousins, Karin Wiberg, and Anders Bignert. Temporal trends in dioxins (polychlorinated dibenzo-p-dioxin and dibenzofurans) and dioxin-like polychlorinated biphenyls in baltic herring (clupea harengus). Marine pollution bulletin, 73(1):220–230, 2013.

(55)

CHAPTER 6. CONCLUSIONS AND FUTURE WORK

[39] K Wiberg, AT Assefa, KL Sundqvist, IT Cousins, J Johansson, MS McLachlan, A Sobek, G Cornelissen, A Miller, J Hedman, et al. Managing the dioxin problem in the baltic region with focus on sources to air and fish. Swedish Environmental Protection Agency Report, 6566, 2013.

[40] EC Commission, Regulation No, et al. 2006 of 19 december 2006, setting maximum levels for certain contaminants in foodstuffs. Official Journal of the European Union, 2006.

[41] Ingemar Cato and Bernt Kjellin. The national swedish status and trend monitoring programme based on chemical contamination in offshore sediment – an overview of the results from 2003. Swedish Geological Survey, 25, 2005.

[42] Lauri Niemistö. A gravity corer for studies of soft sediments. Merentutkimuslait. Julk./Havsforskningsinst. Skr, 238:33–38, 1974. [43] JC Callaway, RD DeLaune, and WH Patrick. Chernobyl 137cs used to

determine sediment accretion rates at selected northern european coastal wetlands. Limnology and Oceanography, 41(3):444–450, 1996.

[44] PG Appleby and F Oldfield. The calculation of lead-210 dates assuming a constant rate of supply of unsupported 210 pb to the sediment. Catena, 5(1):1–8, 1978.

[45] JD Eakins, , and RT Morrison. A new procedure for the determination of lead-210 in lake and marine sediments. The International Journal of Applied Radiation and Isotopes, 29(9):531–536, 1978.

[46] Kristina L Sundqvist, Mats Tysklind, Ingemar Cato, Anders Bignert, and Karin Wiberg. Levels and homologue profiles of pcdd/fs in sediments along the swedish coast of the baltic sea. Environmental Science and Pollution Research, 16(4):396–409, 2009.

[47] John J Ryan, Henry BS Conacher, Luz G Panopio, Benjamin P-Y Lau, Jacques A Hardy, and Yoshito Masuda. Gas chromatographic separations of all 136 tetra-to octapolychlorinated dibenzo-p-dioxins and polychlorinated dibenzofurans on nine different stationary phases. Journal of Chromatography A, 541:131–183, 1991.

[48] EN Cen. 12341 (1998) air quality-determination of the pm10 fraction of suspended particulate matter–reference method and field test procedure to demonstrate reference equivalence of measurement methods. European Committee for Standardization (CEN), Brussels, Belgium, 1998.

(56)

CHAPTER 6. CONCLUSIONS AND FUTURE WORK

[49] R R Draxler and GD Rolph. Hysplit (hybrid single-particle lagrangian integrated trajectory) model. noaa air resources laboratory, silver spring, 2003.

[50] Scott Bridger and Mike Knowles. A complete method for environmental samples by simultaneous axially viewed icpaes following us epa guidelines. Varian ICP-OES At Work, (29), 2000.

[51] JE Longbottom, TD Martin, KW Edgell, SE Long, MR Plantz, BE Warden, R Baraona, D Bencivengo, D Cardenas, L Faires, et al. Determination of trace-elements in water by inductively-coupled plasma-mass spectrometry – collaborative study. Journal of AOAC International, 77(4):1004–1023, 1994.

[52] Anders Bignert, Sara Danielsson, Alma Strandmark, Elisabeth Nyberg, Lillemor Asplund, Ulla Eriksson, Urs Berger, Anders Wilander, and Peter Haglund. Comments concerning the national swedish contaminant monitoring programme in marine biota, 2008. 2008.

[53] Tjelvar Odsjö. The swedish environmental specimen bank with reference to the national contaminant monitoring programme in sweden. Science of the total environment, 139:147–156, 1993.

[54] Anders Bignert, Elisabeth Nyberg, Kristina L Sundqvist, and Karin Wiberg. Spatial variation in concentrations and patterns of the pcdd/f and dioxin-like-pcb content in herring from the northern baltic sea. Journal of Environmental Monitoring, 9(6):550–556, 2007.

[55] Philip K Hopke. Receptor modeling for air quality management, volume 7. Elsevier, 1991.

[56] Philip K Hopke. Recent developments in receptor modeling. Journal of Chemometrics, 17(5):255–265, 2003.

[57] Pentti Paatero and Unto Tapper. Positive matrix factorization: A non-negative factor model with optimal utilization of error estimates of data values. Environmetrics, 5(2):111–126, 1994.

[58] Gary Norris, Ram Vedantham, K Wade, Steve Brown, Jeff Prouty, and Chuck Foley. Epa positive matrix factorization (pmf) 3.0 fundamentals & user guide. US Environmental Protection Agency, Office of Research and Development, Washington, DC, 2008.

[59] Ian Jolliffe. Principal component analysis. Wiley Online Library, 2002. [60] Svante Wold, Kim Esbensen, and Paul Geladi. Principal component

analysis. Chemometrics and intelligent laboratory systems, 2(1):37–52, 1987.

References

Related documents

The ambiguous space for recognition of doctoral supervision in the fine and performing arts Åsa Lindberg-Sand, Henrik Frisk &amp; Karin Johansson, Lund University.. In 2010, a

In the project, a battery of biotest methods currently in use in toxicity assessments were applied using a contaminated Baltic Sea harbour sediment as a model

Source-to-receptor multivariate modeling by Positive Matrix Factorization (PMF) showed that six types of PCDD/F sources are and have been important for the Baltic Sea

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

a) Inom den regionala utvecklingen betonas allt oftare betydelsen av de kvalitativa faktorerna och kunnandet. En kvalitativ faktor är samarbetet mellan de olika

Industrial Emissions Directive, supplemented by horizontal legislation (e.g., Framework Directives on Waste and Water, Emissions Trading System, etc) and guidance on operating

Tommie Lundqvist, Historieämnets historia: Recension av Sven Liljas Historia i tiden, Studentlitteraur, Lund 1989, Kronos : historia i skola och samhälle, 1989, Nr.2, s..

Detta element beskriver hur de ska gå till väga för att börja använda sig utav tjänsten och enligt resultatet visade det sig vara en stor bidragande faktor till att tjänsten