You are on page 1of 44

SCIENCE DOSSIER

Pentachlorobenzene – Sources, environmental fate


and risk characterization

Robert E. Bailey

July 2007
July 2007

1
July 2007
This publication is the eleventh in a series of Science Dossiers providing the scientific community with reliable
information on selected issues. If you require more copies, please send an email indicating name and mailing
address to eurochlor@cefic.be.
The document is also available as a PDF file on www.eurochlor.org
Science Dossiers published in this series:
1. Trichloroacetic acid in the environment (March 2002)
2. Chloroform in the environment: Occurrence, sources, sinks and effects (May 2002)
3. Dioxins and furans in the environment (January 2003)
4. How chlorine in molecules affects biological activity (November 2003)
5. Hexachlorobutadiene – sources, environmental fate and risk characterisation (October 2004)
6. Natural organohalogens (October 2004)
7. Euro Chlor workshop on soil chlorine chemistry
8. Biodegradability of chlorinated solvents and related chlorinated aliphatic compounds (December 2004)
9. Hexachlorobenzene - Sources, environmental fate and risk characterisation (January 2005)
10. Long-range transport of chemicals in the environment (April 2006)
Copyright & Reproduction
The copyright in this publication belongs to Euro Chlor. No part of this publication may be stored in a retrieval system or transmitted in
any form or by any means, electronic or mechanical, including photocopying or recording, or otherwise, without the permission of Euro
Chlor. Notwithstanding these limitations, an extract may be freely quoted by authors in other publications subject to the source being
acknowledged with inclusion of a suitable credit line or reference to the original publication.

2
Foreword

The Environmental Working Group (EWG) is a science group of Euro Chlor, which
represents the European chlor-alkali industry. Objectives of the group are to identify both
natural and anthropogenic sources of chlorinated substances, study their fate, gather
information on the mechanisms of formation and degradation in the environment, and
achieve a better knowledge of the persistence of such substances and communicate these
findings to a wide audience in order to promote science-based decision making. The EWG
often uses external specialists to assist in developing reports that review the state of
existing knowledge of the different aspects mentioned.

Dr. Robert E. Bailey is an independent consultant on environmental chemistry of industrial


products. Prior to setting up his own business, he worked for The Dow Chemical Company
in environmental chemistry from 1969 to 1995. He was supervisor of the corporate
environmental chemistry group from 1976 to 1990 studying all aspects of the fate of
different industrial chemicals. Dr. Bailey’s recent work has focused primarily on the sources
and fate of chlorinated chemicals and the fate of chemicals used in the polyurethane
industry.

Dr Bailey would like to acknowledge the useful contributions to chapter 5.5, “Critical body
burden”, by Dr Paul Thomas, Manager Environmental Chemistry and Regulatory Affairs,
Akzo Nobel Technology & Engineering, Arnhem, The Netherlands.

In Pentachlorobenzene – Sources, environmental fate and risk characterization, Dr Bailey


presents a comprehensive overview of currently available information on this substance.
This Science Dossier presents a thermodynamically self consistent set of physical
properties important to movement in the environment which is suggested for use in future
work. The physical-chemical properties of pentachlorobenzene which are important for the
environmental fate of the substance are presented, covering degradation, bioaccumulation
and the distribution in environmental compartments. There are no uses of PeCB known at
present and an extensive inventory of current emission sources demonstrates that
combustion is the major contributor. A summary of measured concentrations and trends in
various environmental compartments, including biota, is presented, as well as a summary
of ecotoxicity data. Finally, the major findings are summarized in a concluding chapter.

1
Summary

A thermodynamically self consistent set of physical properties important to movement in the


environment has been reported and is suggested for use in future work.
Pentachlorobenzene (PeCB) is a semivolatile chemical with an overall half-life in the
environment measured in years and capable of long range transport.

There are no large scale uses of PeCB at present. Current emissions of PeCB to the
environment are estimated to be about 85,000 kg/year, based on published information.
The largest sources appear to be combustion of solid wastes, 33,000 kg/y, and biomass
burning, 44,000 kg/y, with industrial processes less important.

PeCB has been measured in many environmental media over the past 30 years. Low but
detectable concentrations of PeCB have been reported in sediments and biota in remote
areas all over the world. Concentrations of PeCB in the environment have declined over
the period of monitoring with a 90% decrease of PeCB concentrations in herring gull eggs
from Lake Superior, Canada.

To protect the general public for lifetime exposure, the toxicity of PeCB has been studied
and an oral reference dose of 0.8 µg/kg body weight per day has been established by the
USEPA. The corresponding Canadian tolerable daily intake has been set at 0.5 µg/kg
bw/day. The toxicity of PeCB to aquatic organisms has a lowest reported sublethal effect
concentration of about 10 µg/L for daphnids and amphipods.

Concentrations of PeCB in the environment are low and are expected to continue to drop
with improving technology and waste handling.

2
Table of contents

1 Physical and Chemical properties ....................................................................4


2 Environmental Fate .............................................................................................6
2.1 Environmental Degradation ....................................................................... 6
2.2 Bioconcentration/bioaccumulation ............................................................ 7
2.3 Distribution in environmental compartments ............................................ 8
3 PeCB Sources in the environment ................................................................ 10
3.1 Industrial uses of PeCB............................................................................ 10
3.2 Byproduct formation of PeCB .................................................................. 10
3.3 Combustion formation of PeCB............................................................... 12
3.3.1 Waste combustion ......................................................................... 12
3.3.2 Coal combustion ............................................................................ 13
3.3.3 Biomass combustion ..................................................................... 14
3.4 PeCB from degradation of other chemicals in the environment ......... 15
3.5 PeCB global emissions inventory............................................................ 15
4 Concentrations and trends of PeCB in the environment........................... 16
4.1 Atmosphere................................................................................................ 16
4.2 Water ......................................................................................................... 16
4.3 Sediments .................................................................................................. 18
4.4 Soils ......................................................................................................... 20
4.5 Biota ......................................................................................................... 20
4.6 Comparison of environmental concentrations with
emission and fate ...................................................................................... 26
4.7 PeCB in the Environment – Summary.................................................... 27
5 Ecotoxicity 28
5.1 Aquatic........................................................................................................ 28
5.2 Plants ......................................................................................................... 30
5.3 Wildlife ......................................................................................................... 30
5.4 Soil toxicity.................................................................................................. 30
5.5 Critical Body Burden ................................................................................. 30
6 Conclusions....................................................................................................... 34
7 References ......................................................................................................... 34

3
1 Physical and Chemical properties

PeCB is is a chlorinated aromatic hydrocarbon with the molecular formula C6HCl5. Its CAS
registry number is 608-93-5. The structural formula of PeCB is shown below. A summary
of the physico-chemical properties is given in Table 1.1

Cl

Cl Cl

Cl Cl
MolWt: 250.34 C6 H1 CL5
000608-93-5 Benzene, pentachloro-

Table 1.1. Physico-chemical properties


Property Value Reference.
Molecular mass 250.34
Melting point 84.6 °C (Shen and Wania 2005)
Boiling point 277 °C (Weast 1983)
3
Density of solid 1.8342 g/cm (Weast 1983)
Vapor pressure, solid, 25 °C 0.29 Pa Calculated*
Vapor pressure, sub-cooled liquid, 25 °C 1.0 Pa (Shen and Wania 2005)
Log octanol/water partition coefficient, 25 °C 5.19 (Shen and Wania 2005)
3
Water solubility, solid, 25 °C 0.0027 mol/m (Shen and Wania 2005)
Water Solubility, solid, 25 °C 0.68 mg/L (Shen and Wania 2005)
3
Henry’s Law constant calculated. 25 °C 72 Pa m /mol (Shen and Wania 2005)

* Calculated by linear regression of the literature values used by Shen and Wania.

The properties of PeCB that are key to predicting its movement in the environment are its
vapor pressure (sub-cooled liquid) and partition coefficients between environmental
phases. Because the different partition coefficients are thermodynamically related, it is
possible to adjust values to reduce the likely errors in the individual values. The values
listed above from Shen and Wania are their “final adjusted values” where the literature has
been evaluated to derive a consistent set of the best partitioning values for each chemical
studied. They also derived internally consistent energies of phase transfer for PeCB which
can be used for calculating partition coefficients at different temperatures.
ΔUA = 60.22 kJ/mol
ΔUW = 16.39 kJ/mol
ΔUAW = 43.83 kJ/mol
ΔUOW = -25.11 kJ/mol
ΔUOA = -68.94 kJ/mol
ΔUO = -8.72 kJ/mol (Shen and Wania 2005)

4
The physical properties values listed above have been used in the following model
calculation. The level I model is a static slice of the world containing air, water, soil and
sediments (Mackay 2001). This simple model shows where a chemical accumulates at
equilibrium in the absence of any degradation processes or flow.
PeCB in Level I model
Air 13%
Soil 84%
Water 0.6%
Sediments 2%

The physical properties of PeCB suggest that it will strongly sorb to sediments and soil.
Thus many studies have monitored the environmental distribution of PeCB by analysis of
sediments. Desorption of PeCB from Rhine River and Ketelmeer sediments was reported
-1
to be very slow, 0.00022 h , a half-time of over 4 months, with only a small fraction
desorbing rapidly (ten Hulscher et al. 2002).

5
2 Environmental Fate

2.1 Environmental Degradation

PeCB, like most chlorinated aromatics, is not expected to hydrolyze under environmental
conditions (Boethling and Mackay 2000). PeCB photodechlorination has been reported in
surfactant micelles using UV radiation at 254 nm as a method for treating contaminated soil
(Chu et al. 2002). Mill and Haag reported photolysis of HCB in a water acetonitrile mixture
in California summer sunshine with a half-life of 70 d, although they state that some of the
HCB may have been lost by volatilization (Mill and Haag 1986). The UV absorption range
of HCB extends above 300 nm so that there is some opportunity for environmental
photodegradation. PeCB is expected to behave similarly. Indirect photolysis of HCB in the
presence of humic substances in lake water has been reported and a similar process may
be significant for PeCB (Grannas et al. 2003).

No convincing evidence of aerobic biodegradation of PeCB in the environment was found.


During a biodegradation study of quintozene (pentachloronitrobenzene) which is used as a
fungicide, the fate of PeCB which was present as an impurity was also monitored (Beck
and Hansen 1974). Beck and Hansen reported a half-life for PeCB in soil of 194 and 345
days, an average of 270 days. These slow rates of reported PeCB degradation could
easily have been losses by volatilization from the soil in their experiments. The results of
Beck and Hansen showed that PeCB was apparently formed as an intermediate
degradation product of quintozene as the concentration of PeCB increased during the first
100 days of laboratory incubation before starting its slow decline. The observations by
Beck and Hansen of PeCB in soils which had been treated over a period of years with
quintozene is consistent with their laboratory data, i.e., conversion of some quintozene to
PeCB and its slow removal from soil.

It is probable that PeCB can be biodegraded by the type of fungi that have been shown to
be able to biodegrade HCB and other highly chlorinated aromatics. There have been many
reviews of fungal capabilities and PeCB has not attracted specific attention (Aust and
Benson 1998). For example, white rot fungi, Phanerochaete chrysosporium, was used in a
laboratory incubation of contaminated soil with 100% removal of 1,2,3,4-tetrachorobenzene
and some of the other contaminants (D'Annibale et al. 2005). The contaminated soil only
contained 1 mg/kg PeCB so the extent of its removal was not estimated. A large number
of fungal isolates were screened for their ability to tolerate and degrade HCB and
pentachlorophenol in a highly contaminated soil (Matheus et al. 2000). Of the 36 isolates
screened, 11 were able to grow in soil with greater than 2% HCB, however separate tests
with carbon-14 labeled HCB showed less than 1% yield of carbon-14 CO2 after 128 days
incubation.

Dechlorination of PeCB by anaerobic bacteria has been reported by many workers (Tiedje
et al. 1987; Fathepure et al. 1988; Pardue et al. 1993; Ramanand et al. 1993; Pavlostathis
and Prytula 2000; Brahushi et al. 2004). The dechlorination rate of PeCB was reported to
be faster than that of HCB in a system enriched for HCB degradation (Pavlostathis and
Prytula 2000). The initial source of inoculum was contaminated sediment from Bayou
d’Inde, Louisiana, USA. They reported that the isomer distribution of tetrachlorobenzenes
(TeCB) from dechlorination of PeCB was 91% 1,2,3,5-TeCB and 9% 1,2,4,5-TeCB. Some
of the other reports indicated that alternative paths were preferred with different microbial
systems and they all agree that PeCB goes to a TeCB. Pavlostathis and Prytula reported a
half-life for PeCB in their enriched system of 1.2 days. Brahushi et al. reported that HCB
was converted to PeCB and then to TeCBs as described above in arable soil (Brahushi et
al. 2004). Simply saturating the soil induced native anaerobic microbial communities. The
dechlorination was delayed and reduced by addition of other organics such as straw.
Beurskens et al. reported the anaerobic dechlorination of PeCB to approximately equal
amounts of 1,3,5-trichlorobenzene and 1,2,4-trichlorobenzene by a mixed inoculum derived
from sediments from the Rhine River in Lake Ketelmer (Beurskens et al. 1994).

6
PeCB has been observed in sediments that have been dated more than 60 years earlier
(Durham and Oliver 1983). Therefore, in spite of the laboratory evidence for anaerobic
dehalogenation by anaerobic microorganisms from many sources, PeCB can persist for
many decades in sediments such as those in Lake Ontario. Lake Ketelmeer in the
Netherlands accumulates much of the suspended solids that flow down the Rhine River.
Beurskens et al. have proposed that HCB and the resulting PeCB have been anaerobically
dechlorinated in these sediments to 1,3,5-trichlorobenzene and 1,2- and 1,3-
dichlorobenzenes (Beurskens et al. 1993) . They found that the concentrations of HCB
and PeCB in the ca. 1970 portions of sediment cores taken in 1988 had much lower
concentrations of HCB and PeCB than surface sediments collected in 1972. In addition,
the sediment cores had increased concentrations of tri- and dichlorobenzenes presumably
from the anaerobic dechlorination. A half-life of about 7 years for HCB in these sediments
was derived from the extent of decrease in its concentration. This same activity in
sediments was proposed by Bailey based on the work of Durham and Oliver (Bailey 1983).
However, this interpretation was not supported by the authors (Oliver and Nicol 1983).

Wang and Jones studied the fate of PeCB in soils with and without added organic matter in
the form of sludge (Wang and Jones 1994). They report that the primary mechanism for
PeCB loss from soil is volatilization. There is the potential for photodegradation on the
exposed surface of soil as has been reported in solution. One can also suppose that some
aerobic microorganisms will slowly metabolize PeCB over a period of years. Additionally,
even in surface soils there can be anaerobic zones during extended rainy periods where
PeCB could be dechlorinated. In cold sediments, the presence of many other organic
materials may delay and retard the rate of anaerobic dechlorination. Thus in the following
discussion of environmental fate of PeCB, its degradation half-life has been arbitrarily taken
as 10 years in soils, sediments and water.

Degradation of PeCB in the atmosphere is expected to be primarily through reaction with


OH radicals (Atkinson 1990). The estimated reaction rate of PeCB with OH radicals is
-12 3
0.0579 x 10 cm /molecule-sec at 25 °C (USEPA 2004). This corresponds to an
atmospheric half-life of 185 days calculated assuming an OH radical concentration of 1.5 x
6 3 6
10 radicals/cm for 12 hours per day, a diurnally averaged concentration of 0.75 x 10
3
radicals/cm (USEPA 2004). However the average tropospheric temperature is less, 277
K, so its atmospheric half life will be greater. No activation energy for the reaction of OH
radicals with PeCB was found but if it is assumed that the activation energy is about the
same as that of HCB, 24.3 kJ/mol (Brubaker Jr. and Hites 1998), the predicted half-life of
PeCB in the atmosphere at 277 K is about 370 days.

2.2 Bioconcentration/bioaccumulation

Bioconcentration describes a chemical’s tendency to be taken up by an animal from water


solution. Bioaccumulation is uptake from both water and food. Van de Plassche et al.
summarized bioconcentration (BCF) and bioaccumulation factors (BAF) for PeCB with few
details given (van de Plassche et al. 2001). Most of the studies are described in more
detail below. The reported BAF for PeCB in bluegill sunfish (Lepomis macrochirus) was
3400. In rainbow trout (Oncorhynchus mykiss) a BAF range from 4000 to 8400 and in
guppy a BAF of 13,000 are reported by van de Plassche et al. The much higher BAF
values, based on concentration only in lipid, 155,000 to 260,000, also included by van de
Plassche et al. should not be compared with the whole body BAF values used for
regulatory purposes.

The BCF of chlorobenzenes in guppies (Poecilia reticulata) gave a calculated BCF for
PeCB of 4,700 from its uptake and elimination kinetics (Van Hoogen and Opperhuizen
1988). The fish were exposed for 5 days without feeding and the uptake of PeCB
determined followed by elimination of the PeCB in freshwater for 21 days. Another study of
the PeCB BCF in guppies as a function of temperature reported lipid basis BCF values
ranging from a log of 5.11 at 286 K to 5.28 at 306 K (Opperhuizen et al. 1988). The

7
guppies were 5±2% lipid so that these BCF values can be compared with the whole fish
values above by multiplying by 0.05 to get a range of 6,400 to 9,500.

Detailed studies of PeCB uptake were conducted to determine lethal body burdens of
chlorobenzenes in the amphipod Hyalella azteca (scuds) and fathead minnows
(Pimephales promelas) (Landrum et al. 2004; Schuler et al. 2007). The average BCF,
derived from the study of scuds at the five lowest concentrations tested, was 1900 with up
to 28 days exposure. Higher exposure concentrations yielded lower BCF values for the
scuds. Exposure of the minnows to PeCB for 28 days gave an average log of the whole-
body BCF of 3.07 which equals 1175.

Bioconcentration of a mixture of chlorobenzenes at low concentrations in 250 g rainbow


trout (Salmo gairdneri, now Oncorhynchus mykiss) was studied (Oliver and Niimi 1983).
They reported BCF values for PeCB of 13,000 and 20,000 after 119 and 105 days
exposure, respectively. These results led to predicted concentrations of PeCB in Lake
Ontario trout which were close to those found in the field. They stated “These results would
strongly suggest that, excluding HCB, the chemical concentrations [of chlorobenzenes] in
the water largely control the concentrations in fish for these chemicals.”

(Carlson and Kosian 1987) studied the bioconcentration and toxicity of PeCB to embryo
and juvenile fathead minnows (Pimephales promelas). After 31 days exposure to PeCB
the average BCF was reported to be 8,400.

An extensive field study using samples collected from 1982 to 1986 in Lake Ontario and
selected tributaries examined the concentrations of PeCB, HCB and other chlorinated
chemicals in a variety of environmental samples and biota (Oliver and Niimi 1988). This
study is hard to interpret quantitatively because fish and water samples were collected at
different times from different locations. However, bioaccumulation factors (BAF) for PeCB
for several aquatic organisms representing different trophic levels, including several trout
species, can be calculated from the results presented in their publication. PeCB
concentrations were 72±15 pg/L in water, 8.4±6.5 ng/g wet weight in mysids, 5.0±3.7 ng/g
ww in amphipods, 2.6 ng/g ww in sculpin, 2.1 ng/g ww in large smelt, and 5.0±3.1 ng/g ww
in large trout. The BAF for PeCB in trout can be calculated from this data as (5.0 ng/g)/(72
pg//L) = 69,000 L/kg. Similar concentrations of PeCB at different levels of the food web
shows a minimal effect of PeCB uptake from food.

A field study of PeCB concentrations in the industrially polluted Bayou d’Inde, Louisiana,
USA reported apparent average log BAF values on a lipid basis (Pereira et al. 1988).
These lipid basis values have been converted to whole-body BAF values by multiplying by
the reported lipid percent in the different species. The calculated whole-body BAF values
in fish were: Atlantic croaker (Micropoganias undulates) 18,700: spotted sea trout
(Cynoscion nebulosis) 2,100; blue catfish (Ictalurus furcatus) 12,300. The calculated BAF
in blue crab (Callinectes sapidus) was 6,600. A later study in the same area using
samples collected in 1990 gave similar BAF values but the lipid contents were not reported
so that whole-body BAF values could not be calculated (Burkhard et al. 1997).

2.3 Distribution in environmental compartments

Degradation of PeCB in different environments is described above. However, extrapolating


from laboratory conditions to define the expected environmental degradation rate involves
many approximations. The most effective way to look at the fate of a chemical in the
environment is to consider not only its degradation rate in each compartment of the
environment but also its partitioning between compartments. This way one can estimate an
“overall” degradation rate or half-life (lifetime).
The level III model adds degradation processes, plausible transfer rates between phases,
and PeCB flow into and out of the model space to the partitioning based on physical
properties (Mackay 2001). Using the half-lives above, 370 days in the atmosphere and 10
years in soil, water and sediments, and assuming all PeCB is emitted into the atmosphere,

8
the percentages of PeCB accumulating in each compartment and degrading in each
compartment are:

Compartment Percentage of PeCB Percentage of PeCB


accumulated degrading
Air 30.6 81
Soil 68.1 18
Water 0.2 <1
Sediments 1.1 <1

The conclusion from this simple modeling exercise is that atmospheric transport and
degradation are predicted to be important even though only about one third of the PeCB is
predicted to be present in the atmosphere at equilibrium. The dramatic difference between
equilibrium partitioning (see section 1 above) and the result above is due to the relatively
slow equilibration between atmospheric PeCB and soil below the surface. Vulykh et al. at
the Meteorological Synthesizing Centre – East came to the same conclusion, 63% is
degraded in the atmosphere, using their much more complex model developed for
evaluation of POPs (Vulykh et al. 2005). This means that on a global scale PeCB is
expected to be found widely distributed. Because the average temperature is lower in the
polar regions, semivolatile chemicals can tend to accumulate in polar environments. This
behavior has been explained and modeled by Wania and Mackay (Mackay and Wania
1995; Wania and Mackay 1995).

9
3 PeCB Sources in the environment

No global inventory of possible emissions of PeCB to the environment was found.


However, in Canada total releases of PeCB have been estimated to be 41.8 kg/y. PeCB
from backyard trash burning is estimated to be 21.93 kg/y, pentachlorophenol treated
wood, 2.34 kg/y, pesticide use 6.2 kg/y, dielectric fluid 5.6 kg/y, municipal solid waste
incineration 2.36 kg/y, hazardous waste incineration 1.84 kg/y, magnesium production 1.53
kg/y and solvent use 0.04 kg/y (Environment Canada 2005). No estimates of PeCB
emissions in the remainder of the world were found . Thus, the following factors for
estimating global PeCB emissions were developed for this report

3.1 Industrial uses of PeCB

PeCB is not known to have any commercial uses at present (Beck 1986; Environment
Canada 1993). However, in the past, PeCB was one component of a chlorobenzenes
mixture used to reduce the viscosity of PCB products employed for heat transfer
(Environment Canada 1993; King et al. 2003). PeCB has also been used in a
chlorobenzenes mixture with PCBs in electrical equipment (Environment Canada 2005).
PCBs are still in use in some old electrical equipment in North America and Europe so that
there is a small potential for release of PeCB from this source (AMAP 2000; Environment
Canada 2003). It can be presumed that some PCBs are also still in use elsewhere in the
world and some fraction of them contain PeCB.

PCBs are being taken out of service in many countries of the world so that any related
PeCB emissions are expected to be decreasing with time. Global releases of PCBs have
been estimated to be approximately 40 metric tons in 2000 (Breivik et al. 2002). Breivik
noted a high level of uncertainty in the emissions estimate, plus and minus a factor of 10.
Breivik’s estimate of Canadian PCB emissions was 0.6 metric tons per year. Assuming the
same ratio of PeCB emission to PCB emission, the global emission of PeCB from use with
PCBs can be roughly calculated:
((5.6 kg/y PeCB)/(0.6 t/y PCB))*(40 t/y PCB) = 373 kg/y

PeCB was used in the past as an intermediate in manufacture of pentachloronitrobenzene


(quintozene) (van de Plassche et al. 2001). However, quintozene is now made by
chlorination of nitrobenzene (Feiler 2001). PeCB may also have been used in the past as a
fungicide and flame retardant (van de Plassche et al. 2001).

3.2 Byproduct formation of PeCB

The U. S. Toxics Release Inventory (TRI) includes PeCB. A total of 20 industrial facilities
are listed as emitting or transferring 2533 pounds (1151 kg) of PeCB in 2004 (USEPA
2006). The industries represented are chemical, waste treatment, and coal burning electric
power. Total reported air emissions were 220 pounds (100 kg) and water emissions of 17
pounds (7.8 kg) (USEPA 2006). In industrial chlorination reactions it is possible to produce
PeCB as a byproduct and it probably accounts for some of the emissions reported in the
TRI. There are other processes which produce a variety of chlorinated aromatics that may
contribute PeCB even if PeCB has not been explicitly detected and reported yet. TRI
includes only industrial facilities handling relatively large amounts of chemicals so that
additional emissions are expected.

One way to estimate global emissions from the global chemical and related industry sectors
covered by the TRI program in the USA is to simply multiply by the ratio of world gross
domestic product (GDP) to the US GDP. Since the USA accounts for about 25% of world
GDP, multiplying the TRI reported emissions to air and water by 4 leads to estimated global

10
emissions of 431 kg/y. This value is uncertain because of both the uncertainty in the US
emissions and waste handling practices in developing industrial countries. North American
emissions of PeCB were greater in the past due to less stringent waste handling
regulations as shown by monitoring studies described in Section 4.

Global emissions from chlorinated solvent use and pesticide use have been estimated by
scaling from the Canadian emissions reported by Environment Canada (Environment
Canada 2005). In round numbers, the Canadian GDP is about 2.5% of the global GDP.
Therefore, multiplying the PeCB emissions from the use of chlorinated solvents, 0.04 kg/y,
by 40 gives an estimate of 1.6 kg/y for global emissions from chlorinated solvents.
Emissions from pesticide and pentachlorophenol use are calculated similarly:
(6.2 + 2.34 kg/y)*(40) = 341.6 kg/y

The emission of PeCB has been reported from the use of hexachloroethane (HCE) to
remove dissolved hydrogen from molten aluminum in foundries (Westberg et al. 1997).
They reported an emission rate of 310 µg PeCB/g HCE used at a typical HCE application
rate. While the use of HCE for degassing aluminum in foundries is not widespread in the
USA, several small foundries were reported to use HCE (Streeter 1998). Because HCE is
easier to use than gaseous chlorine for degassing aluminum, it is probably used in some of
the less developed countries as well as in smaller facilities in developed countries. In 1998
it was reported that about 3,500,000 kg/y HCE was used in this application (Scottish
Chemical 1998). Using the Westberg et al. factor, this could result in the emission of:
-9
(310 µg/g)*(3,500,000 kg/y)*(10 kg/µg)*(1000g/kg) = 1100 kg/y PeCB.

Vogelgesang also reported the release of PeCB and other compounds from degassing of
aluminum (Vogelgesang 1986). PeCB emissions from an aluminum smelter using a
3 3
sodium/potassium chloride flux were reduced from 18.05 µg/m to 10.34 µg/m by
installation of a baghouse filter (Aittola et al. 1996). The other chlorobenzenes and
chlorophenols were reduced by a similar fraction. In contrast, the filter reduced the
polychlorinated dioxins and furans by 98-99%.

There are several processes for production of metals involving the treatment of ores with
carbon and chlorine to yield a soluble or volatile compound which can be purified.
Carbochlorination of magnesium oxide, heating MgO with coke to 700-800 °C in a chlorine
atmosphere, yields MgCl2 and CO along with traces of PeCB, HCB and dioxins/furans and
other chlorinated compounds (Knutzen and Oehme 1989). Knutzen and Oehme estimated
that about 50 kg/year PeCB along with up to 300 kg HCB may have been emitted to the
Frierfjord, Norway during the 1980s (Knutzen and Oehme 1989). Norwegian regulations
called for dramatic reduction in dioxin/furan emissions in the 1990s so that it may be
assumed that PeCB emissions have decreased also (although not necessarily in direct
proportion). Recovery of copper from a slag by roasting it with coal and sodium chloride
produced a variety of highly chlorinated organics (Doering et al. 1992). The residue from
this process was an attractive red gravel which was used in landscaping. Concentrations
of PeCB in this red slag ranged from 0.6 to 1200 µg/kg and HCB ranged from 5.3 to 11000
µg/kg in samples. Vogelgesang reported release of PeCB and other chlorinated
compounds from the production of niobium and tantalum (Vogelgesang 1986). The author
of this report, Bailey, speculates that other processes which use chlorine and carbon at
high temperatures may produce PeCB, such as the chloride process to produce TiO2
pigment. No quantitative estimates are provided because there is no quantitative
information on which to base them.

11
3.3 Combustion formation of PeCB

The observation of PeCB as a trace product of incomplete combustion has been widely
reported. Nearly all fuels contain some chlorine, especially biomass and waste. The yield
of PeCB from combustion of different fuels under different conditions has been reported to
vary widely. The goal in this section is to extract PeCB yield factors from the various
research studies and apply these factors to combustion processes for estimation of global
PeCB emissions from combustion.

3.3.1 Waste combustion


Open burning of household waste and the resulting emissions of PeCB averaged 0.076
-8
(range 0.033-0.162) mg/kg mass actually burned or a factor of 7.6 x 10 kg/kg (Lemieux
1997). The variation between duplicate burns in this small project prevent any conclusions
from the two different waste mixtures used and illustrate the uncertainty in the emission
factor. This study was designed to represent uncontrolled combustion of solid waste in a
perforated barrel. It probably applies to much of the uncontrolled burning of trash which
takes place in piles all over the world as well as landfill fires.

Emissions of PeCB and other chlorobenzenes during incineration have been studied as a
potential way to estimate emissions of chlorodioxins and furans (PCDD/F) with simpler
analytical technology. The use of PeCB for monitoring has been suggested by Kato and
Urano who found a correlation between PeCB concentrations and the international toxic
equivalent factor (I-TEQ) in 5 different operating waste disposal plants in Japan (Kato and
Urano 2001). They found that under normal operating conditions PeCB correlated with I-
TEQ within a factor of 3.

Other workers have also reported correlations between the emission of PeCB and
polychlorodibenzodioxins and furans (PCDD/F) (Oberg and Bergstrom 1987; Kaune et al.
1994). Aittola et al. pointed out that a filter that removed over 90% of the PCDD/F,
removed only about 50% of PeCB and HCB from an aluminum smelter (Aittola et al. 1996).
Thus reductions in PCDD/F emissions may not lead to equal reductions in PeCB
emissions.

The reported yields of PeCB to HCB formed in different combustion processes vary widely,
depending on combustion conditions and the presence (or not) of catalytic materials. For
example, in a fluidized bed solid waste incinerator the yield of total chlorobenzenes and
PCDDs increased 20 fold as metals accumulated in the sand, while the PeCB/HCB ratio
was approximately the same (Akimoto et al. 1997). Because combustion conditions, for
example, oxygen, time and temperature, seem to dominate in most situations, the influence
of fuel composition, concentration of chlorine beyond a minimum concentration, on the
PeCB emissions is only one of several factors.

Many of the studies of trace products of incomplete combustion report in terms of mass per
volume of stack gas which is what the regulations require. To convert to a yield of PeCB in
3
terms of mass of waste burned, a gas flow of 7,000 m /metric ton has been used
(Carpenter et al. 1986). Combustion experiments are difficult to control and the PeCB from
nominally duplicate runs varied by factors of 2 to 5 in those studies where replicates were
reported. The variety of parameters studied also led to wide variation in the PeCB yields.
Table 3.1 has only average PeCB yields and illustrates some of the variation reported. The
yield of PeCB is certainly affected by poor combustion conditions. Another significant
factor in the overall yield of chlorinated aromatics is the presence of solid surfaces,
especially if metals are present. One conclusion from all this work is that PeCB is probably
emitted from all combustion processes of fuels containing any chlorine as organochlorine or
chloride.

12
Table 3.1. Yields of PeCB in various waste incineration experiments. The results of
different experiments in each study have been averaged. PeCB concentrations have been
3
converted to emission factors by multiplying by 7000 m of stack gas per ton of fuel.
PeCB Emission PeCB Experimental Study
Concentration Emission
Factor
3
0.91 µg/std m 6.4 mg/ton Lab fluidized bed (Wikstrom et al. 1999)
3
39 µg/std m 273 mg/ton City incinerator (Tiernan et al. 1983)
3
1 µg/std m 7 mg/ton 24 city incinerators (Kato and Urano 2001)
3
23 µg/std m 161 mg/ton Lab fluidized bed (Fängmark et al. 1993)
3
12 µg/std m 84 mg/ton Lab fluidized bed (Fängmark et al. 1994)
3
0.42 µg/std m 2.9 mg/ton City incinerator (Jay and Stieglitz 1995)
3
25 µg/std m 175 mg/ton City fluidized bed (Akimoto et al. 1997)
3
2 µg/std m 14 mg/ton Hazardous waste (Oberg et al. 1985)
incinerator
Special purpose experiments not used in waste combustion averaging.
3
87 ng/std m 0.6 mg/ton Wood burner (Zimmerman et al. 2001)
969 mg/ton Pure PVC in pilot plant (Ahling et al. 1978)
burner
300 mg/ton Pure PVC in lab burner (Kim et al. 2004)

The range of PeCB yields above extends over a factor of 100 from experiments designed
to clarify the processes taking place during municipal incineration. The geometric mean of
the emission rates from the waste combustion experiments above, except the Tiernan 1983
study, is about 25 mg/ton. For the purpose of an order of magnitude estimate of PeCB
emissions, the emission factor for municipal and hazardous waste incineration will be 25
-8
mg/ton of waste or 2.5 x 10 kg/kg. The total amount of waste incinerated with the low
emission factor above is estimated to be about 200 million tons per year which could yield
PeCB.
-8 6
(2.5 x 10 )*(200 x 10 tons) = 5 tons = 5000 kg/y PeCB emitted.

In most of the developing world trash is burned casually to reduce its volume and attraction
of vermin. The production of solid waste is estimated to be about 0.4 kg per capita-day
(Zurbruegg 2003). For this report it is estimated that about half of the solid waste is
9 9
burned. Thus a population of 5 x 10 people would generate 2 x 10 kg/day of total solid
waste. Using the average PeCB emission factor from Lemieux, see above (Lemieux 1997),
the yield of PeCB on an annual basis would be:
9 -8
(2 x 10 kg/day)*(365 d/y)*(0.5 burned)*(7.6 x 10 ) = 27,740 kg/y

3.3.2 Coal combustion


Coal use in 2005 was estimated to be 4990 million metric tons (World Coal Institute 2006).
Coal typically contains about 0.1% chlorine. No information on PeCB emissions from coal
3
combustion was found. However 0.07 µg/std m HCB has been reported from coal
combustion in a lab experiment (Oberg and Bergstrom 1985). Oberg and Bergstrom did
not report the conditions of combustion but simply reported “satisfactory and fully
comparable”. In contrast, HCB was not detected in a series of coal fired utility boilers
tested in the USA (USEPA 1998). Because a sizable fraction of coal is burned in small

13
domestic units, some PeCB emissions are expected. The references cited above generally
reported several times more PeCB than HCB from combustion processes with low
concentrations of chlorine. For the purposes of this report, half of the coal used is assumed
to be burned domestically and the reported HCB emission concentration from Oberg and
Bergstrom will be multiplied by 5. The PeCB emission factor is calculated:
3 3 -9
(0.07 µg/std m )*(7000 m /ton)*(5) = 2450 µg/ton or 2.45 x 10 kg/kg PeCB per kg coal.
Multiplying this factor by coal burned in small scale combustion yields:
-9 9
(2.45 x 10 kg/kg)*(0.5)*(4990 x 10 kg) = 6113 kg PeCB

3.3.3 Biomass combustion

Total biomass burning in the 1990s was estimated to be 3716 million metric tons of carbon
with a chlorine content ranging from 2.4 mg/kg (tulip trees in Ohio, USA) to 9000 mg/kg
(citrus leaves in Southern California) (Lobert et al. 1999). The only emission factor for
PeCB from burning biomass found was that of Zimmerman where waste wood was burned
in a high temperature incinerator (Zimmerman et al. 2001). No information on PeCB
emissions from smoldering combustion or other combustion conditions was found.
However, Gullet and coworkers have reported a series of studies on PCDD and PCDF
emissions, expressed as international toxic equivalents (I-TEQ), from combustion of
biomass under conditions designed to simulate agricultural burning and forest fires (Gullet
and Touati 2003a; Gullet and Touati 2003b; Gullet et al. 2006). Table 3.2 summarizes the
I-TEQ emissions from the different studies.

Table 3.2. Comparison of emission factors from biomass combustion studies.


Substance I-TEQ ng/kg C PeCB ng/kg C Reference
emission factor emission factor
Wood 0.8* 1200 ng/kg C Zimmerman et al. 2001
Sugarcane 1.7-25** Gullet et al. 2006
Forest fire 4-30 Gullet and Touati 2003a
Straw ~1 Gullet and Touati 2003b
3 3
* Emissions converted from concentration (ng/m ) to emission factor (ng/kg C) by multiplying by 7000 m flue gas
per ton of wood burned, and assumes wood to be 50% carbon.
** The very high emission factor, 253 ng TEQ/kg C from Hawaiian sugarcane, was not used because it seems to
be an outlier.

Several researchers have noted a correlation between the emission of PeCB and PCDD/Fs
from waste combustion (Oberg and Bergstrom 1985; Kato and Urano 2001). Zimmerman
et al. reported the average emission of PeCB was 1475 times the average I-TEQ emission
3 3
in their wood burning experiment (87 ng/m /0.059 ng/m = 1475). This ratio of PeCB to I-
TEQ is within the range, 500 to 2500 ng PeCB per ng of TEQ, reported in the studies
summarized in Table 3.1, above. Based on the I-TEQ emissions from uncontrolled
biomass burning in Table 3.2, it is reasonable to estimate the PeCB emission factor for
most biomass burning is somewhat greater than that reported by Zimmerman et al. For the
purposes of this report, developing an approximate global emission inventory for PeCB, the
average emission PCDD/Fs (I-TEQ) from combustion of biomass will be taken as 8 ng/kg C
burned. Using the PeCB/I-TEQ ratio from Zimmerman the emission factor to be used for
PeCB is 11,800 ng/kg C burned. Multiplying this factor by the amount of biomass burned
yields an emissions estimate.
6 16
(11,800 ng/kg)*(3716 x 10 ton)*(1000 kg/ton) = 4.39 x 10 ng PeCB
= 43,900 kg PeCB

14
3.4 PeCB from degradation of other chemicals in the environment

Dechlorination processes for HCB have been studied in the laboratory. HCB was
photodechlorinated to PeCB in surfactant micelles (Chu et al. 2002). The PeCB thus
formed was subsequently dechlorinated at about the same rate as HCB. Therefore a
negligible net production of PeCB is expected from this environmental process. Indirect
photolysis of HCB in arctic lake water was reported to be facilitated by dissolved organic
matter; however there was no mention of the products (Grannas et al. 2003).

Anaerobic biological dechlorination of HCB to PeCB followed by dechlorination of PeCB


has been reported by many workers (Tiedje et al. 1987; Fathepure et al. 1988; Pardue et
al. 1993; Ramanand et al. 1993; Beurskens et al. 1994; Pavlostathis and Prytula 2000).
However, the dechlorination rate of PeCB is reported to be faster than that of HCB so that
PeCB does not accumulate but is further dechlorinated to tetrachlorobenzenes and lower
chlorobenzenes. Thus anaerobic biological dechlorination of HCB is not expected to lead
to a net accumulation of PeCB in the environment.

Pentachloronitrobenzene (PCNB) has been reported to degrade in soil forming a small


yield of PeCB along with much higher yields of other related compounds (Beck and Hansen
1974) . PCNB has also been reported to be photolyzed to PeCB along with other related
compounds (Crosby and Hamadmad 1971). Global agricultural use of PCNB in the 1990s
was reported to be 880,590 kg (Landell Mills Market Research 1996). Thus there is the
potential for the release of some PeCB. PeCB from this source was not included in the
summary of PeCB sources, Table 3.3, because the available information did not allow
calculation of the potential quantity.

3.5 PeCB global emissions inventory

Using the factors and logic described above, the estimated global emissions of PeCB
around the year 2000 are summarized in Table 3.3, 85,000 kg/y. As described above there
is considerable uncertainty about the size of each of these estimated PeCB emissions,
potentially an order of magnitude. Global emissions are clearly dominated by combustion
sources. Of all sources, combustion of biomass (43,900 kg/y), combustion of solid waste
(32,740 kg/y) and combustion of coal (6,113 kg/y) represent the three largest emissions.
Industrial sources are relatively minor and improvements in industrial practices have
probably led to the reductions in environmental concentrations of PeCB noted in Section 4.

Table 3.3. Summary of estimated annual global emissions of PeCB.


PCB use losses 373 kg
Chlorinated solvents <2 kg
Pesticide use 342 kg
Chemical manufacturing and waste disposal 431 kg
Aluminum casting 1100 kg
Combustion of solid waste 32,740 kg
Combustion of coal 6113 kg
Combustion of biomass 43,900 kg
Total annual emission 85,001 kg

15
4 Concentrations and trends of PeCB in the
environment

4.1 Atmosphere

During a project to monitor PeCB and other chemicals in the Canadian atmosphere to look
for seasonal changes in 1988-1989, PeCB was detected in 133 out of 143 samples (Hoff et
3
al. 1992). Its mean concentration was reported as <8.0 pg/m . Some PeCB was reported
lost (volatilized) during summer sampling so that it was impossible to calculate a true mean.

Passive air samplers were deployed across North America for one year from 82° north
(Alert, Canada) to 10° north (Monteverde, Costa Rica) in 2000-2001 to determine the long
range transport and distribution of a variety of organochlorine compounds (Shen et al.
2005). The reported concentrations of PeCB were relatively constant across the continent
3 3
with an annual average of 45 pg/m with an apparent range of 17 to 136 pg/m . The
3
corresponding annual average for HCB was 89 pg/m with an apparent range of 49 to 131
3
pg/m . This agrees well with the average HCB concentration reported from the Integrated
Atmospheric Deposition Network (IADN) above the North American Great Lakes in 2000,
3
about 72 pg/m (Buehler et al. 2004). This wide distribution of relatively constant PeCB
atmospheric concentrations is consistent with a compound having a long atmospheric
lifetime and release from a wide variety of processes. One parameter that apparently does
influence the measured concentrations of PeCB (and HCB) is the altitude at a given
location. Both in the work by Shen et al. and in a study using analysis of mosses in the
Andean Mountains concentrations were reported to be higher at higher elevations (Grimalt
et al. 2004).

A study of the influence of different emission sources on the mixtures of chloroorganics


present in the atmosphere in Germany showed the chlorobenzenes, including PeCB, at
higher concentrations near industrial or urban locations than at a rural reference site
(Wenzel et al. 2006). Samples were collected July through September 1995 with a winter
urban sample in February 1996. Reported concentrations of PeCB ranged from 5.7 to 28.6
3 3
pg/m in the industrial and urban sites and 0.31 pg/m at the rural reference site. Kaj and
Palm reported atmospheric concentrations of PeCB in Southern Sweden in 2003 to
3
average 0.033 ng/m , similar to the concentrations reported elsewhere (Kaj and Palm
2004).

4.2 Water

Concentrations of chemicals in the Niagara River were monitored for two years, 1981-1983
(Oliver and Nicol 1984). During this period reported concentrations of PeCB ranged from
0.34 to 6.4 ng/L with a mean of 1.3±1.0 ng/L. HCB concentrations ranged from 0.16 to 29
ng/L and the mean concentration was 1.1±2.9 ng/L. A more comprehensive monitoring
program was started later which separated dissolved and particulate bound chemicals.
Concentrations of PeCB at the mouth of the Niagara River where it flows into Lake Ontario
have been monitored carefully during the period 1987-1997 showing annual mean total
(dissolved plus particulate) concentrations dropping from 0.351 ng/L to 0.093 ng/L, a 75%
reduction in the amount of PeCB entering Lake Ontario (Williams et al. 2000). PeCB mean
concentrations at the head of the Niagara River were 0.024 to 0.015 ng/L over the same
period. The estimated total concentration of 93 pg/L at the mouth of the Niagara River in
1996-1997 corresponds to a loading of 52 g per day to Lake Ontario (Williams et al. 2000).
Most of this PeCB was not present at the head of the Niagara River but was added from
various tributaries. Approximately 25% of this PeCB was present on the suspended solids.
The concentrations of PeCB on the suspended solids are shown in Figure 4.1.

16
PeCB has also been monitored in the St. Clair River (Chan 1993). Concentrations of PeCB
in the St. Clair River were greater at the mouth than at the head, and the mean PeCB
concentration at the mouth dropped between 1988 and 1989. Both of these monitoring
programs have continued but more recent data is not expected to be available until 2007
(Neilson 2006).

A survey of Great Lakes water to determine concentrations of chlorobenzenes in water and


sediments was carried out April-November 1980 (Oliver and Nicol 1982). PeCB was not
detected at some of the open water sites and mean

Figure 4.1. PeCB in Niagara Bar sediment and Niagara River suspended solids. Durham
and NYDEC data are from sediment cores and Williams data is from suspended solids.
Note logarithmic concentration scale.

1000

100
PeCB ng/g

Durham
NYDEC
Williams
10

1
1900 1920 1940 1960 1980 2000
Year

concentrations of PeCB were reported to be 0.2 ppt (part per trillion) in Lake Ontario and
0.04 ppt in Lake Huron. Water samples collected in the spring of 1986 in the Canadian
Great Lakes showed detectable PeCB in many samples (Stevens and Neilson 1989).
None was detected in Lake Superior at a detection level of 0.007-0.011 ng/L. In Lake
Huron PeCB was detected in 11% of the samples with an average concentration in those
samples of 0.033 ng/L, and in Georgian Bay PeCB was detected in 29% of the samples at
an average concentration of 0.019 ng/L. In the lower Great Lakes, concentrations were
only a bit higher, 14% detected in Lake Erie with an average of 0.058 ng/L and in Lake
Ontario 70% detected with an average PeCB concentration of 0.062 ng/L. Water sampling
results in the mid to late 1990s were statistically treated to calculate the upper 90%
confidence interval of compounds like PeCB which were only detected part of the time
(Williams et al. 2001). Thus the upper 90% confidence interval reported for Lake Superior
in 1997 was 0.024 ng/L PeCB; Lake Erie in 1998 was 0.02 ng/L; and Lake Ontario in 1998
PeCB was 0.05 ng/L.

Precipitation has been monitored on the Canadian side of the Great Lakes for
contamination by organochlorine compounds, including PeCB (Chan et al. 2003). PeCB
was detected in less than 50% of samples so that total amounts of these compounds could
not be calculated. A study of PeCB in snow and ice from the Russian arctic reported only
traces (Melnikov et al. 2003). In southeast Africa, on the shore of Lake Malawi, PeCB
concentrations in precipitation samples collected in 1997 and 1998 ranged from 1 to 48
pg/L with an average of 10±14 pg/L (Karlsson et al. 2000).

17
A 1993 study of the distribution of chloroorganics in the North Pacific Ocean, the Bering
and Chukchi Seas found PeCB in all samples (Strachan et al. 2001). PeCB concentrations
were an average of 16 pg/L in the dissolved phase. Suspended solids represented just a
small fraction of the total PeCB, adding 0.38 pg/L. Strachan et al. calculated that the flow
of water northward into the Arctic Ocean carries 0.31-0.52 metric ton of PeCB per year.

Water and sediment in the Yangtse River near Nanjing were analyzed for organochlorines
in 1998 (Jiang et al. 2000). The total PeCB concentrations were about 0.4 ng/L in the river
water. An average of 57% of the PeCB was in the dissolved phase and on particulate
matter less than 0.7 µm. The highest sediment concentration of PeCB at one location was
3 ng/g. Water from four rivers draining industrial regions in the U.K. were sampled weekly
for two years to determine the chlorobenzenes contributed to the Humber estuary (Meharg
et al. 2000). The flux of chlorobenzenes into the Humber was dominated by 1,2-
dichlorobenzene, 56 kg/year, and 1,4-dichlorobenzene, 65 kg/year, with only 0.8 kg/year of
PeCB discharged.

4.3 Sediments

Sediment cores from Lake Ontario near the mouth of the Niagara River reveal a history of
PeCB loading to Lake Ontario which parallels that of the other chlorobenzenes and
chlorinated chemicals. Figure 4.1 shows PeCB reached its peak concentration about 1960
followed by a substantial decline to about 10% of the peak by 1980 (Durham and Oliver
1983; NYDEC 1998) in this industrially impacted area. Concentrations of PeCB on
suspended solids in the Niagara River flowing into Lake Ontario have continued to drop
during the 1990s as shown by suspended solids analyses (Williams et al. 2000). It is
interesting to note how well the PeCB concentrations on suspended solids agree with the
concentrations in sediment cores. Note also that this more recent concentration trend is
the same as observed in the earlier sediments.

Kaminsky et al. analyzed sediment samples collected from a number of sites in western
Lake Ontario in October 1980 to trace the movement of contaminants from the Niagara
River (Kaminsky et al. 1983). PeCB concentrations ranged from 2 ppb to 32 ppb. In
another study the distributions of organochlorines in Lake Ontario sediments were studied
to estimate total quantities which had accumulated in surface sediments as of 1981 (Oliver
et al. 1989). At that time they estimated a total inventory of about 3 metric tons of PeCB in
Lake Ontario surface sediments. Surface sediments in Lakes Erie, Huron, and St. Clair
were analyzed for chlorobenzenes and other chlorinated chemicals in 1980 and 1982
(Oliver and Bourbonniere 1985). Mean PeCB concentrations in sediments in Southern
Lake Huron were 1.5 ng/g, Lake St. Clair 5.8 ng/g, Western Lake Erie 2.9 ng/g, Central
Lake Erie 1.0 ng/g, and Eastern Lake Erie 0.9 ng/g. The implication is that one source of
PeCB was along the St. Clair River and perhaps another along the Detroit River.

Muir et al. have determined the concentrations of PeCB in the sediment of a series of
remote lakes in northern Canada (Muir et al. 1995). Sediment surface layer concentrations
(representing a period of time approximately centered on 1979-1988) of PeCB in these
northern lake sediments ranged from less than 0.01 to 0.73 ng/g sediment. Table 4.1
summarizes these results. The PeCB concentrations have been divided by the fraction of
organic carbon (OC) in the sediments to allow comparison with regulatory limits. The
sediment PeCB concentrations found in four Alaskan Arctic lakes sampled in 1991-1993
averaged 0.10 ±0.10 ng/g dry weight, approximately the same as the Canadian lakes
described in Table 4.1 (Allen-Gil et al. 1997).

18
Table 4.1. Concentration of PeCB in remote Lake Sediments in Northern Canada
(Muir et al. 1995).
Lake PeCB ng/g % OC PeCB ng/g OC
Lake 375 0.28 13.4 2.1
Lake 382 <0.01 16.9 <0.06
Far Lake 0.21 9.4 2.2
Hawk Lake 0.32 13.5 2.4
Amituk Lake 0.73 1.2 61
Sophia Lake 0.01 3.0 0.33
Buchanan Lk <0.01 1.3 <0.08
Hazen Lake 0.01 1.7 0.59

Chlorinated contaminants have been determined in the sediments and flood plain soils of
the River Elbe watershed (Witter et al. 1998; Schwarzbauer et al. 2001). Two of three
locations on the flood plain had concentrations of PeCB as high as 64 and 71 µg/kg (Witter
et al. 1998). The third site was reported to have only 1 µg/kg or less PeCB. Concentrations
of PeCB appear to have peaked in the 1960s as shown by concentrations at different
depths. Sediment samples collected in 1993 and 1994 in the Berlin area along the Spree
and Havel Rivers also contained PeCB (Schwarzbauer et al. 2001). Most of the sampling
locations had <10 µg/kg PeCB but the maximum concentration reported in the Havel River
system was 76 µg/kg and in the Spree was17 µg/kg. Sediments from seven locations on
the Elbe River downstream from Hamburg were analyzed for PeCB by three different
methods (Eder et al. 1987). In general the PeCB concentrations reported ranged from 0.47
to 4.4 ng/g wet weight.

Sediment samples collected from the Lippe River, Germany, between 1999 and 2001 were
analyzed for PeCB and other organics (Kronimus et al. 2004). PeCB was detected in many
of the samples at concentrations ranging up to 6 ng/g from a location near its mouth at the
Rhine River. There was considerable temporal variation with even the station showing the
highest concentrations of PeCB reporting 1 ng/g at one sampling time. A survey of
pollutants in sediments of the Ebro River reported PeCB in 25% of the samples with an
LOD of 0.48 µg/kg (Lacorte et al. 2006). The mean PeCB concentration of detected
samples was 1.17 µg/kg.

The concentration of PeCB in sediment from the Mulde Reservoir (Saxony, Germany) was
determined in a study of chemical partitioning and mobilization under anoxic conditions
(Zoumis et al. 2001). The reported concentration of PeCB was 0.4 µg/kg. In Ontario
(Canada) the Sir Adam Beck Reservoir takes water from the Niagara River and after a
relatively brief holding period returns it to the Niagara River through a power plant. In 1998
the sediment was sampled in several locations for determination of contaminant
concentrations (Williams et al. 2003). The highest reported PeCB concentration was 0.4
ng/g at one location with only a trace at other locations.

Chlorobenzenes were determined in surface sediments collected off the coast of Taiwan in
the vicinity of several sewage outfalls in 1996 (Lee et al. 2000). PeCB was detected at all
but one of the 40 sampling stations at concentrations ranging up to 15.7 ng/g. These same
locations were sampled a year later, 1997, to allow more statistical analysis to locate
pollution sources (Lee et al. 2005). PeCB concentrations were similar in this second study
and the authors did not report locations of any point sources for PeCB.

19
Low concentrations of PeCB, 0.01 to 0.06 µg/kg were reported in sediment from the
industrially polluted Kishon River, Israel (Oren et al. 2006). As in many of the studies,
concentrations of PeCB were correlated with the organic carbon content of the sediment.
Surface sediments in Masan Bay, Korea, were analyzed in 1997 for chlorinated compounds
(Hong et al. 2003). PeCB was detected in only a few samples at up to 0.28 ng/g.

4.4 Soils

The concentrations of chlorobenzenes in surface soil at five sites in the Niagara Falls
(USA) area were studied (Ding et al. 1992). The five sites were chosen to show if
significant quantities of chlorobenzenes were migrating through the air from the Love Canal
hazardous waste site or accumulating from contaminated water. The site near Love Canal
was compared with nearby industrial sites and a suburban site 12 miles from the canal.
PeCB was detected in soil samples from all the sites. Concentrations in individual soil
samples ranged from non detect up to 1700 pg/g. The suburban site had a mean PeCB
concentration of 480 pg/g with a range of 180 to 1200 pg/g. The conclusion was that the
Love Canal hazardous waste site had not released excessive amounts of chlorobenzenes
to the atmosphere which deposited in the region.

The European Reference Soil-Set was analyzed for a variety of chemical


microcontaminants including PeCB (Gawlik et al. 2000). PeCB concentrations in the five
soils ranged from 0.10 to 0.79 ng/g. Analysis of digested sewage sludges from five
wastewater treatments near Vancouver (Canada) to determine their suitability for
application to cropland did not detect PeCB at a detection level of <0.01 µg/g (Bright and
Healey 2003).

4.5 Biota

PeCB has been monitored in Great Lakes herring gull eggs annually since the 1970s by the
Canadian Wildlife Service. Concentrations at all herring gull colonies have dropped
dramatically since then (Bishop et al. 1992; Petit et al. 1994; Pekarik et al. 1998; Jermyn-
Gee et al. 2005; Havelka 2006). The exact extent of decrease varies at the different sites
and the trend is partially obscured by the non-detected PeCB in many of the more recent
samples. For example, the PeCB concentrations in herring gull eggs from Muggs
Island/Leslie Spit near Toronto dropped from 50 ng/g in 1979 to non-detected at 1 ng/g by
the mid 1990s, a 98% decrease. Figure 4.2 shows concentrations of PeCB in herring gull
eggs at three representative sites on the Great Lakes. Herring gull eggs on Lake Superior
probably reflect the background atmospheric concentration of PeCB because they are
generally distant from most industrial activity. The trend lines indicate an average decrease
in PeCB concentration of about 17% and 14% per year for the eggs from the shores of
Lake Ontario and Lake Superior, respectively. Table 4.2 summarizes the decreases in
PeCB concentrations from all the sampling sites. In the future it will not be possible to
reliably evaluate trends because most of the PeCB concentrations are less than the limit of
quantitation in the CWS herring gull egg program. A recent publication (Hebert and
Weseloh 2006) adjusts the rates of decline in chemical concentrations for changes in the
trophic position of the herring gulls in different colonies as a result of environmental
changes. Because some herring gulls have moved down the food web as a result of eating
more terrestrial food, the true decline in contaminant concentration may not be as great as
indicated by the straight concentrations. For HCB this has the result of bringing the
observed rate of decline in herring gull egg concentrations into closer agreement with the
reported declines in atmospheric concentrations. There is no comparable data set for
PeCB in the atmosphere.

20
Figure 4.2. Concentrations of PeCB in herring gull eggs from colonies on the shores of the
Great Lakes. Note the logarithmic concentration scale. Concentrations less than the limit
of quantitation (LOQ) have been plotted as 0.5 the LOQ. The trend lines for eggs from
Lakes Superior and Ontario show the average rate of decrease in PeCB concentration.

1000.0
y = 3E+165e
-0.1907x Ontario
Michigan
100.0
Superior
PeCB ng/g

Expon. (Ontario)
10.0 Expon. (Superior)

LOQ 1 ng/g
1.0
-0.1462x
y = 6E+126e
0.1
1970 1980 1990 2000 2010
Year

Analyses of PeCB and other chloroorganics in Lake Ontario biota and environment
collected 1981-1986 enabled determination of bioconcentration and biomagnification
factors (Oliver and Niimi 1988). The concentrations of PeCB are listed in Table 4.3.

Table 4.2. Great Lakes herring gull colonies and rates of PeCB concentration decrease in
herring gull eggs for the entire period of monitoring through 2004. See the CWS reports
cited above for exact dates and monitoring sites.
Location Rate Constant Percent Decrease Half-
-1 -1
(Year ) (Year ) life
(Years)
St Lawrence R., Strachan Island -0.071 6.9 9.8
Lake Ontario, Snake Island - West -0.144
Brothers Island 13.4 4.8
Lake Ontario, Mugg’s Island - Leslie -0.191
Spit 17.4 3.6
Hamilton Harbor -0.112 10.6 6.2
Niagara River -0.167 15.4 4.1
Lake Erie, Colbourne Light -0.149 13.8 4.7
Lake Erie, Middle Island -0.115 10.9 6.0
Detroit River, Fighting Island -0.158 14.6 4.4
Lake Huron, Chantry Island -0.107 10.1 6.5
Lake Huron, Channel Shelter I. -0.173 15.9 4.0
Lake Huron, Double Island -0.136 12.7 5.1
Lake Michigan, Gull Island -0.111 10.5 6.2
Lake Michigan, Big Sister Island -0.105 10.0 6.6
Lake Superior, Agawa Rock -0.146 13.6 4.7
Lake Superior, Granite Island -0.121
11.4 5.7

21
A survey of Great Lakes fish caught during the late 1970s found PeCB only in fish from
Lake Ontario and the Ashtabula River, Ohio, and its tributary (Kuehl et al. 1981). More
recently the US National Study of Chemical Residues in Fish (Kuehl et al. 1994) detected
PeCB at about 22% of the 388 sites nationwide, with the highest concentrations near
chemical manufacturing plants. Pereira et al. and Burkhard et al. reported PeCB along with
other halogenated organics in biota and sediments from the lower Calcasieu River and the
Bayou d’Inde, Louisiana (Pereira et al. 1988; Burkhard et al. 1997). A 1993 study of
contaminants in brown bullhead fish from the Detroit River, Michigan found PeCB along
with much higher concentrations of PCBs and other chloroorganics in this industrially
impacted river (Leadley et al. 1998). Mean concentrations of PeCB were 13.0, 29.4 and
16.1 µg/kg wet weight in the fish from Amherstburg Channel (east side), Trenton Channel
(west side) and Peche Island at the head of the river, respectively.

Table 4.3. Concentrations of PeCB in Lake Ontario biota and environment (Oliver
and Niimi 1988).
Species, units PeCB
Water, pg/L 72±15
Bottom sediments, ng/g dry wt. 33±14
Suspended sediments, ng/g dry wt. 13±3.9
Plankton, ng/g wet wt. 0.6±0.3
Mysids, ng/g wet wt. 8.4±6.5
Amphipods, ng/g wet wt. 5.0±3.7
Oligochaetes, ng/g wet wt. 0.8±0.3
Sculpin, ng/g wet wt. 2.6
Alewive, ng/g wet wt. n.d.
Small smelts, ng/g wet wt. n.d.
Large Smelts, ng/g wet wt. 2.1
Fish (large salmonids), ng/g wet wt. 5.0±3.1

Zebra mussels and eel in the Rhine and Meuse Rivers in 1994 were analyzed for a wide
variety of chemicals and metals (Hendriks et al. 1998). PeCB concentrations in zebra
mussels from the Rhine, Meuse, and Ysselmeer were 0.49, 0.27, 0.50 µg/kg wet weight,
respectively. PeCB in eel from the Rhine, Meuse, and Hollands Diep were 15, 2.9, and 7.7
µg/kg wet weight, respectively. A survey of chemicals in fish from the Ebro River in Spain
detected PeCB in 14 out of 18 samples with an LOD of 0.30 µg/kg. The mean
concentration of PeCB was 1.10 µg/kg with a range of 0.32 to 3.31 µg/kg (Lacorte et al.
2006).

A quantity of PeCB, estimated at 59 kg, mixed in 5400 kg of PCB heat transfer fluid was
released in the Gulf of St. Lawrence after a barge sank in 1970 (King et al. 2003). The
barge was raised in 1996 at which time it was discovered that much of the heat transfer
fluid had leaked into the environment. Snow crab digestive glands were monitored for their
PeCB content and showed a rapid decline over the five year period at the sampling point
right where the barge had rested. Concentrations were 150 ng/g wet weight in 1996, 12
ng/g wet weight in 1997, 3.6 ng/g wet weight in 1998, 3,6 ng/g wet weight in 1999, 3.1 ng/g
wet weight in 2000. The other sampling sites, one or more miles distant, showed no
increased concentrations over what is apparently the PeCB background concentration in
that region, even in 1996.

A study of organochlorines in fin whale blubber collected in 1971 and 1972 from the
Newfoundland area showed mean concentrations of PeCB of 0.96 and 0.01 ng/g lipid for
females and males, respectively (Hobbs et al. 2001). From Nova Scotia the concentrations
were 0.23 ng/g lipid and non detected for females and males, respectively. Because HCB
was still in wide use in agriculture at that time, the corresponding concentrations of HCB
were 244 and 333 ng/g lipid for females and males from Newfoundland and 217 and 221
ng/g lipid for females and males from Nova Scotia. Blubber biopsies from St. Lawrence
Estuary beluga whales in 1994-1998 showed PeCB concentrations ranging widely from

22
1.56 ng/g lipid to 1510 ng/g with a geometric mean of 24.5 in females (Hobbs et al. 2003).
In males the range was similar, 1.5-1500 ng/g lipid with a geometric mean of 144 ng/g.

A study of PeCB in fish from four Alaskan arctic lakes found mean concentrations (ng/g wet
weight) of PeCB of 1.42±1.82 in grayling liver, 0.06±0.08 in grayling muscle, 0.48±0.35 in
lake trout liver and 1.21±3.66 in lake trout muscle (Allen-Gil et al. 1997). A terrestrial top
predator, the arctic fox has been studied for accumulation of chlorinated chemicals
(Hoekstra et al. 2003). Samples were collected at three sites: Arviat on Hudson Bay,
Canada, Holman, Northern Territory, Canada and Barrow, Alaska. About 20 animals at
each site were collected during 1999-2001 at some distance from human habitation to
minimize effects of garbage scavenging. The PeCB concentrations found in arctic foxes
were: Arviat, muscle, 0.61±0.12; Holman, muscle, 0.29±0.06; Holman, liver, 0.57±0.11;
Barrow, muscle, 0.55±0.20; Barrow, liver, 0.73±0.17.

The polar bear has been studied as another top predator whose diet allows for
accumulation of persistent chemicals. A wide ranging study of bears from Alaska, Canada,
East Greenland and Svalbard sampled between 1996 and 2002 looked for geographical
variations in chemical concentrations (Verreault et al. 2005). Unfortunately for this purpose
the data was reported in terms of HCB and the sum of PeCB, HCB and 1,2,3,4-
tetrachorobenzene (TeCB). Thus only the sum of PeCB and TeCB can be calculated
which can be interpreted as an upper bound on the PeCB concentration. HCB constituted
about 75% of the sum of these three chlorobenzenes. Table 4.4 shows the reported
concentrations and the sum of TeCB and PeCB. Concentrations of chlorobenzenes were
relatively uniform between these polar bear populations spread over about half of the arctic.

Body burdens and concentrations of chlorobenzenes in polar bears of different ages have
been studied before and after their seasonal fasts (Polischuk et al. 2002). Polischuk et al.
reported that none of the chlorobenzenes included in their summation (1,2,4,5-TeCB,
PeCB, and HCB) were excreted or metabolized during the fast so that concentrations
increase as fat was metabolized. They also reported that nursing polar bear cubs received
increased amounts of chlorobenzenes so that the concentration of chlorobenzenes in cubs
is greater than that in adult bears.

Table 4.4. Concentrations of chlorobenzenes in polar bear lipid from adipose tissue,
geometric mean in ng/g lipid (range) (Verreault et al. 2005).
Location CBz HCB PeCB+TeCB
calculated
Alaska (males) 113 (70.4-181) 84.5 (35.1-157) 28.5
Alaska (females) 118 (70.0-277) 85.7 (41.7-230) 32.3
Amundsen Gulf 113 (71.1-190) 75.5 (43.9-146) 37.5
W. Hudson Bay 97.5 (55.9-257) 75.3 (39.3-229) 22.2
Foxe Basin/Gulf of Boothia 127 (73.4-329) 87.3 (46.3-305) 39.7
Lancaster Sound/Jones Sound 148 (111-186) 107 (74.9-146) 41
N. Baffin Island 191 (108-656) 152 (76.7-620) 39
S. Baffin Island 111 (42.0-275) 87.1 (33.2-249) 23.9
E. Greenland 79.1 (36.5-323) 60.0 (25.5-311) 19.1
Svalbard 105 (49.0-248) 90.4 (37.9-229) 14.6

Chlorobenzenes and other organochlorine compounds have been determined in a variety


of Greenland biota collected in 1998-2001 (Vorkamp et al. 2004). As above, only the
concentrations of HCB and the sum of 1,2,3,4-TeCB, PeCB and HCB are reported so that
the sum of TeCB and PeCB have been determined by difference. As shown in Table 4.5,

23
the concentration of PeCB+TeCB in lipid is much less than that of HCB in most cases. The
concentration of PeCB+TeCB is very small in most tissues when they are adjusted for the
lipid content of the tissues. In most species the chlorobenzenes concentration in lipid is
about the same in the different tissues.

Table 4.5. Median concentrations and ranges of chlorobenzenes concentrations in


Greenland biota in ng/g lipid weight (Vorkamp et al. 2004). The percentage of lipid in
tissues was used to calculate the wet weight concentration of PeCB+TeCB, the potential
PeCB dose to predators from consumption of prey.
Species Tissue % Lipid CBz, ng/g lw HCB, ng/g lw PeCB+TeCB PeCB+TeCB
, ng/g lw , ng/g ww
Ptarmigan Liver 6.7 6.6 (4.3-29) 2.9 (2.1-5.1) 3.7 0.248
Muscle 3.8 5.4 (3.5-9.6) 3.6 (3.5-6.8) 1.8 0.068
Hare Liver 4.0 170 (120-330) 170 (120-330) 0 0.000
Muscle 3.1 12 (3.1-56) 9.9 (3.0-40) 2.1 0.065
Kidney 36 17 (15-170) 17 (15-170) 0 0.000
Lamb Liver 9.1 4.8 (1.8-16) 4.6 (1.7-16) 0.2 0.018
Muscle 8.8 1.5 (0.65-4.4) 1.2 (0.53-3.6) 0.3 0.026
Kidney 3.9 5.4 (3.1-12) 4.8 (2.9-11 0.6 0.023
Blubber 91 0.24 (0.11-0.48) 0.21 (0.089-0.41) 0.03 0.027
Caribou Liver 7.3 6.3 (3.9-7.6 6.2 (3.9-7.4) 0.1 0.007
Muscle 1.5 9.1 (7.0-10) 8.7 (6.8-9.5) 0.4 0.006
Kidney 3.4 3.8 (2.3-5.6) 3.7 (2.2-5.4) 0.1 0.003
Blubber 64 7.5 (3.8-9.6) 7.3 (3.7-9.3) 0.2 0.128
Muskox Liver 9.8 150 (46-190) 150 (45-190) 0 0.000
Muscle 2.0 620 (23-1100) 620 (23-1100) 0 0.000
Kidney 3.1 180 (4.9-460) 180 (4.3-460) 0 0.000
Blubber 89.8 0.79 (0.29-7.7) 0.46 (0.13-6.2) 0.33 0.296
Arctic char Muscle 1.1 29 (18-75) 28 (18-48) 1 0.011
(From three Muscle 1.5 52 (37-110) 49 (36-88) 3 0.045
Locations) Muscle 2.8 28 (19-53) 25 (17-48) 3 0.084
Shrimp Muscle 0.95 75 (41-110) 15 (12-19) 60 0.570
Snow crab Liver 4.0 49 (27-66) 40 (21-53) 9 0.360
Muscle 0.72 63 (38-86) 45 (25-55) 18 0.130
Iceland Muscle 0.40 1.8 (0.80-3.3) 0.44 (n.d.-1.2) 1.36 0.005
scallop
Atlantic cod Liver 58 30 (28-33) 27 (25-30) 3 1.740
Muscle 0.68 38 (18-71) 33 (16-49) 5 0.034
Redfish Muscle 1.8 38 (20-45) 34 (19-39) 4 0.072
Atlantic Muscle 9.2 16 (14-24) 13 (11-18) 3 0.276
salmon
Greenland Liver 39 48 (39-730) 42 (34-710) 6 2.340
halibut
Muscle 10 51 (24-62) 44 (19-54) 7 0.700
Wolffish Liver 21 38 (27-58) 34 (24-53) 4 0.840
Muscle 1.7 42 (26-60) 37 (23-54) 5 0.085
Capelin Muscle 1.7 51 (37-79) 47 (35-73) 4 0.068
Shorthorn Liver 17 60 (45-110) 52(39-94) 8 1.360
sculpin
Shorthorn Liver 23 53 (39-110) 44 (34-91) 9 2.070
sculpin
Shorthorn Liver 15 16 (10-37) 14 (8.7-32) 2 0.300
sculpin
Shorthorn Liver 10 17 (13-26) 14(12-22) 3 0.300
sculpin

24
Species Tissue % Lipid CBz, ng/g lw HCB, ng/g lw PeCB+TeCB PeCB+TeCB
, ng/g lw , ng/g ww
Common Liver 5.0 81 (45-89) 71 (38-75) 10 0.500
eider
Muscle 3.5 60 (42-110) 50 (36-90) 10 0.350
King eider Liver 5.3 65 (49-130) 56 (42-110) 9 0.477
Muscle 3.5 73 (52-120) 62 (45-100) 11 0.385
Kittiwake Liver 5.7 13 (9.4-30) 2.2 (1.0-6.3) 10.8 0.616
Muscle 14 20 (9.6-71) 3.6 (0.57-44) 16.4 2.296
Thick-billed Liver 5.5 110(77-170) 97 (69-150) 13 0.715
murre
Muscle 3.6 89 (46-360) 76 (41-320) 13 0.468
Ringed seal Liver 5.4 20 (13-34) 12 (8.9-17) 8 0.432
Muscle 12 20 (9.4-71) 13 (6.7-29) 7 0.840
Blubber 97 16 (1.1-60) 9.3 (0.65-27) 6.7 6.499
Ringed seal Liver 5.3 11 (7.0-22) 5.2 (4.1-16) 5.8 0.307
(Second Muscle 3.2 11 (6.9-43) 7.6 (5.2-14) 3.4 0.109
Location) Kidney 3.8 10 (4.7-18) 5.4 (2.7-11) 4.6 0.175
Blubber 92 14 (8.6-39) 7.9 (5.0-13) 6.1 5.612
Harp seal Liver 6.2 78 (21-250) 72 (13-250) 6 0.372
Muscle 1.7 58 (15-140) 48 (12-120) 10 0.170
Kidney 3.0 21 (12-52) 19 (9.5-48) 2 0.060
Blubber 87 71 (14-120) 66 (11-110) 5 4.350
Minke whale Liver 5.8 170 (150-250) 170 (150-250) 0 0.000
Muscle 1.2 120 (71-240) 120 (71-240) 0 0.000
Kidney 3.4 120 (52-250) 120 (52-250) 0 0.000
Blubber 18 n.a. 160 (15-610) n.a.
Beluga Liver 6.3 210 (40-410) 190 (34-380) 20 1.260
Muscle 1.8 370 (210-570) 350 (190-510) 20 0.360
Kidney 3.2 250 (31-360) 230 (26-310) 20 0.640
Skin 3.6 160 (16-320) 150 (15-310 10 0.360
Blubber 88 260 (27-690) 250 (21-550) 10 8.800
Narwhal Liver 4.7 430 (290-810) 420 (280-780) 10 0.470
Muscle 2.2 450 (50-800) 440 (45-770) 10 0.220
Kidney 2.5 400 (180-470) 390 (170-450) 10 0.250
Skin 3.5 310 (250-410) 300 (240-390) 10 0.350
Blubber 87 490 (390-830) 470 (370-790) 20 17.400

An extensive study of organochlorine compounds in seals from the east and west sides of
the Northwater Polnya between Canada and Greenland looked for influences of diet (Fisk
et al. 2002). Tissue samples from the ringed seals were collected by Inuit hunters during
the spring of 1998. Fisk et al. reported 8.4±1.1 ng/g wet weight in females and 7.3±1.9
ng/g PeCB in males from the west side, Grise Fiord. On the east side, Qanaq, females
contained 5.0±0.5 and males 7.0±1.5 ng/g wet weight of PeCB.

A study of organochlorine concentrations in seal blubber, fishes and invertebrates from the
White Sea in Northwestern Russia found PeCB along with other compounds (Muir et al.
2003). The mean concentrations of PeCB in the different species are shown in Table 4.6.
Harp seal pups collected in 1992 and 1998 were analyzed to look for trends in
contamination. The mean concentration (± standard deviation of the 10 samples) of PeCB
in 1992 was 11±2.0 ng/g lipid weight. In 1998 the concentration of PeCB was 5.0±1.8 ng/g
lipid weight. Apparently, the concentrations of PeCB dropped by approximately 50% over
the period of 1992 to 1998 as did concentrations of nearly all the other organochlorines
measured.

25
Table 4.6. Mean concentrations of PeCB in seal blubber, fish and invertebrates from
the White Sea, Russia (Muir et al. 2003).
PeCB
Bearded seal blubber 0.9 ng/g lipid
Harp seal (adult) blubber 12 ng/g lipid
Ringed seal (juvenile) blubber 2.5 ng/g lipid
Ringed seal female blubber 2.9 ng/g lipid
Ringed seal male blubber 2.1 ng/g lipid
Navaga whole fish 5.06 ng/g wet wt.
Bullrout whole fish 0.01 ng/g wet wt.
White Sea cod muscle 0.01 ng/g wet wt.
White Sea herring whole fish 3.77 ng/g wet wt.
Smelt whole fish 3.81 ng/g wet wt.
Isopod 0.04 ng/g wet wt.
Zooplankton 0.07 ng/g wet wt.
Spider crab 0.08 ng/g wet wt.
Whelk <0.01 ng/g wet wt.

Samples of cod liver and halibut liver from near Adak Island in the Aluetian Chain of
islands, Alaska were analyzed for organochlorine compounds (Arend et al. 2001). The
PeCB liver concentrations were 1.4, 0.8 and 1.4 µg/kg lipid for Sweeper Cove cod, Kuluk
Bay cod and Kuluk Bay halibut, respectively. A study of organochlorine compounds in
bowhead whales from Barrow, Alaska found mean PeCB concentrations of 0.3±0.01 ng/g
wet weight in liver and 0.8±0.1 ng/g in blubber (Hoekstra et al. 2002).

PeCB has also been detected in Antarctic biota (Corsolini et al. 2006). Corsolini et al.
report the concentrations of PeCB to be 0.05 ng/g in krill, 0.08 ng/g in emerald rockcod
muscle, 0.09 ng/g in rockcod whole body and 0.68 ng/g in adelie penguin eggs.

4.6 Comparison of environmental concentrations with emission and


fate

One potential concern is the risk of missing a significant source which could result in an
unexpected exposure of humans or the environment to PeCB. One way of checking for
unknown sources is by back calculating the expected emissions from measured
environmental concentrations using a model. This predicted emission rate can be
compared with reported emissions. Obviously the uncertainties in concentration
measurements, estimated emissions and models need to be remembered when
interpreting the results which can only give a rough indication at best.

In the modeling above, section 2.3, illustrating the movement of PeCB in the environment,
PeCB can accumulate in the soil but the majority, 80%, is degraded in the atmosphere.
Thus, on a global basis atmospheric degradation is expected to be important. Thus a very
simple model can be used to look for glaring inconsistencies.

26
Consider the northern hemisphere as a continuous stirred pot reactor. The volume of the
atmosphere can be calculated from its depth, 8000 m (Weast 1983) and the area of the
2
northern hemisphere, 255 million km . If the concentration of PeCB in the atmosphere is
3
45 pg/m , then the total amount of PeCB in the atmosphere can be calculated:
6 2 6 2 2 18 3
NH atmosphere (255 x 10 km )*(8000 m)*(10 m /km ) = 2.04 x 10 m
-12 3 18 3
PeCB (45x10 g/m )*(2.04 x 10 m ) = 91,800,000 g PeCB

The globally, seasonally, diurnally averaged atmospheric half-life of 370 days is equivalent
-1
to a first order reaction rate of 0.69 y , see Section 2.1 above. Thus the amount of PeCB
3
that needs to be added to the atmosphere to maintain a concentration of 45 pg/m is:
-1
0.69 y * 91,800 kg = 63,200 kg/y.
This is about the same as found in the “bottom-up” inventory based on sources reported in
the literature, section 3. This calculated number of 63,000 kg/y ignores any degradation of
PeCB dissolved in water, sorbed to soil or sediments or exported irreversibly to polar
regions and across the equator. Inclusion of the additional fates for PeCB as done by the
EMEP environmental model (Vulykh et al. 2005) would increase the amount needed to
maintain the atmospheric concentration. It is important to be aware of the uncertainties
involved in all aspects of these calculations, PeCB emissions, modeling simplifications and
environmental concentrations. The surprisingly close agreement between the calculated
emissions and the calculated total degraded suggests that there are no major hidden
emission sources which are dramatically affecting environmental concentrations of PeCB.

Because only atmospheric degradation has been considered, it is conservative in that the
inclusion of additional degradation processes will increase the overall degradation rate.
Because only published sources of emissions of PeCB in the environment have been
included there could be additional sources which have not been considered. The assumed
atmospheric PeCB concentration is based on only one study (Shen et al. 2005), and it is
generally consistent with the higher concentrations reported from some of the other studies
of PeCB in the atmosphere.

4.7 PeCB in the Environment – Summary

• PeCB has been observed at low concentrations essentially everywhere in the


environment that has been carefully analyzed.
• Polar bear adipose tissue had the highest reported concentrations of PeCB+TeCB
with an average concentration of 30 ng/g lipid.
• Among the highest reported PeCB concentrations in prey organisms (e.g. for polar
bears) is about 5 ng/g wet weight in the blubber of Arctic seals.
• PeCB concentrations in herring gull eggs on the shore of Lake Superior, Canada
have dropped by over 90% since the 1970s.
• Concentrations of PeCB have dropped by over 90% since the 1960s in sediments
near the industrially impacted Niagara Falls area of the US and Canada.
• PeCB concentrations in sediments from remote lakes in northern Canada averaged
0.20 ng/g compared to about 8 ng/g off the mouth of the Niagara River in Lake
Ontario.

27
5 Ecotoxicity

5.1 Aquatic

Acute toxicity information on PeCB to aquatic organisms was summarized by van de


Plassche et al. and is included in Table 5.1 (van de Plassche et al. 2001). Environment
Canada extrapolated from the toxicity of the lower chlorobenzenes to derive an estimate for
toxicity of PeCB to benthic organisms because there was no information on toxicity of
PeCB to benthic organisms (Environment Canada 2003). The extrapolation was based on
the observation that the chlorobenzenes appear to be non-specific in their toxic action,
narcosis. That is, the toxic internal concentrations of the different chlorobenzenes tested in
the organisms, daphnids, were about the same on a molar concentration basis. The
internal concentrations were calculated on the basis of chlorobenzenes partitioning
between water and the daphnid. The partitioning of PeCB in sediments was also used to
derive a critical toxicity value for freshwater sediments (CTVsed). The resulting CTVsed for
PeCB was 2500 µg/g organic carbon (OC). In marine sediments the extrapolated CTVsed
for PeCB was 3080 µg/g OC for the LOEC.

A recent detailed study on the time-dependent toxicity of PeCB to Hyalella azteca, a scud,
has been published (Landrum et al. 2004). The LC50 water concentration was about 125
µg/L at 28 days. The 28-day growth rate of Hyalella azteca was reduced from exposure to
a water concentration of approximately 0.05 µmol/L (12.5 µg/L).

The toxicity of PeCB to juvenile sand crabs (Portunus pelagicus) was studied (Mortimer
and Connell 1994). They reported an LC50 of about 0.3 µmol/L (75 µg/L) at 96 hours, after
the crabs had molted. There appeared to be a discontinuity in the line of time versus LC50
which they stated may be the result of increased sensitivity during molting. A comparison
of toxicities of a wide variety of chemicals listed the mean LC50 of two PeCB studies with
Cladocera (Daphnia) as 3218 µg/L and the LC50 for PeCB with Anostraca (fairy shrimp),
376 µg/L (Sánchez-Bayo 2006).

Fathead minnow (Pimephales promelas) embryo to early life stages were used to measure
chronic toxicity of PeCB (Carlson and Kosian 1987). They reported that a concentration of
55 µg/L PeCB for 32 days, which was the highest concentration that they could maintain in
their system, led to no statistically significant mortality or reduction in growth, a NOEC. The
concentration of PeCB in the minnows exposed to 55 µg/L PeCB averaged 380 µg/g (1.52
µmol/g). They also exposed 12 juvenile minnows to 130 µg/L for 6 days to look for acute
toxicity. Only one minnow died during the experiment but some of the others showed loss
of equilibrium after 48 hours. Another study of toxicity and the critical body burden (CBB)
of PeCB in fathead minnows reported an LC50 between 200 and 250 µg/L at both 12 and
28 days (Schuler et al. 2007).

28
Table 5.1. Toxicity of PeCB to aquatic organisms.

29
5.2 Plants

Lettuce (Lactuca sativa) was grown in soil containing PeCB. A 14-day study indicated an
EC50 of 280 µg/g and an NOEC of 50 µg/g for growth (Environment Canada 1993).

5.3 Wildlife

No wildlife toxicity reports were found. A detailed review of mammalian toxicity is beyond
the scope of this report. However, both the USEPA and Environment Canada have
interpreted key animal studies to derive a daily intake of PeCB deemed safe for the general
human population including sensitive subpopulations for lifetime exposure. The USEPA
Integrated Risk Information System (IRIS) has derived an RfD (reference dose) of 0.8 µg/kg
bw/day (USEPA 1988). This value was taken from a subchronic study on rats (Linder et al.
1980). The lowest effect dose was divided by a factor of 10 to correct it for chronic toxicity,
a factor of 10 to drop it to the no effect level, a factor of 10 to extrapolate to humans and
other species and a final factor of 10 to protect sensitive subpopulations. Environment
Canada derived a Tolerable Daily Intake (TDI) for PeCB of 0.5 µg/kg bw/day (Environment
Canada 1993). This TDI was based on a subchronic study in mice (McDonald 1991) with
an adjustment factor of 10,000 applied to the lowest observed effect dose, 5.2 mg/kg
bw/day, to protect the general human population.

5.4 Soil toxicity

The toxicity of PeCB to two species of earthworms, Eisenia andrei and Lubriculus rubellus,
was determined in two different soils (van Gestel et al. 1991). For Eisenia andrei the 14
day LC50s were 134 (>100 <180) mg/kg and 238 (>180 <320) mg/kg. For Lumbriculus
rubellus the 14 day LC50s were 115 (109-122) mg/kg and 210 (>150 <270) mg/kg. Belfroid
et al., in a study of toxicity and the CBB of PeCB in earthworms (Eisenia andrei), used
several different exposure systems and reported an LC50 between 250 and 400 mg/kg in
soil (Belfroid et al. 1993). Worms exposed to water containing approximately 0.2 mg/L
PeCB died between 4 and 7 days. Feeding worms manure containing 60 mg/g PeCB killed
over half of them between 19 and 23 days. Worms exposed to filter paper treated with 75
2 2
and 7.5 µg/cm PeCB died between 24 and 45 hours. Exposure to 0.75 µg/cm PeCB on
filter paper was not lethal to any of the worms.

5.5 Critical Body Burden (with contributions by Dr. Paul Thomas)

According to the critical body burden (CBB) concept, originally proposed more than two
decades ago (McCarty 1986), chemicals will have a specific effect (lethal or sub-lethal)
upon organisms once the latter have attained a specific internal tissue concentration. The
internal concentration is rather independent of external influences and mode of exposure
(e.g. aquatic or via the food) and should occur at the same level regardless of the organism
or species used. This concept, which is also referred to as ‘critical tissue residue’ or
‘internal dose’ (e.g. Escher and Hermens, 2004) is now reasonably well-established,
particularly for acute effects for chemicals that act via a narcotic mode of action. According
to Thompson & Stewart (2003), a reasonable estimate for Lethal body burdens (LBB) with
a non-polar narcotic mode of action is 2 to 8 mmol/kg with a more approximate indicative
range of 1 to 10 mmol/kg. Chronic CBB was estimated with less confidence but data
provided by Thompson & Stewart suggest their estimate of 0.2 to 0.8 (0.1-1 approximate
indicative range) mmol/kg is appropriate. Terminology is not always consistent, but
generally the acronym “LBB” is used to describe the tissue concentration corresponding to
an EC50 (generally obtained from acute studies), while “CBB” corresponds to a no effect
level from a longer term study.

30
Apart from reducing the uncertainty of identifying safe levels as compared to standard tests
using ambient concentrations, an additional advantage is that the CBB method can be used
to compare biomonitoring results for chemicals with potential for adverse effects in the
environment (McCarty and Mackay 1993). In this chapter, we consider available body
burden data from the literature and attempt to calculate a value for CBB for PeCB that can
be compared with existing biomonitoring data. Table 5.2 summarizes CBB data which were
collated by (Jarvinen and Ankley 1999). Studies are described individually due to technical
variations in design.

Recently, Schuler et al. (2007) studied the CBB of PeCB in a fish, the juvenile fathead
minnow (Pimephales promelas). They found the LBB for 50% mortality to be 1.26 mmol/kg
whole body wet weight and the lowest-observed-effect CBB to be 0.66 mmol/kg for reduced
growth rate over 28 days. For comparison with monitoring data, these values can be
converted to µg/g by multiplying by the molecular weight of PeCB, 250.34. The CBB results
for this study are 315 µg/g and 165 µg/g for 50% mortality and reduction in growth rate,
respectively.

Further studies on fish support this recent study. An early life stage study on fathead
minnow (Pimephales promelas) embryos carried out over a period of 31 days, to measure
chronic toxicity of a series of chlorobenzenes, failed to observe lethality or any statistically
significant effects on growth at the highest achievable water concentration of 55 µg/l (±16
µg/l) for PeCB (Carlson and Kosian 1987). At the conclusion of the study, the concentration
of PeCB in the whole fish was 1.52 mmol/kg. As no effects were observed the CBB value
can be considered as greater than 1.52 mmol/kg.

Acute values from the same study after 6 days exposure of the fish to 130 µg/l PeCB led to
a whole body PeCB concentration of 0.23 mmol/kg. During this test one fish died and five
out of twelve exhibited loss of equilibrium. No further mortalities or effects were observed.
The authors recalculated the body burdens of fish in both chronic and acute tests
normalized to 1% lipid. These were found to be twice as high in the acute test (1.16
mmol/kg) as in the chronic study (0.63 mmol/kg) but the reason for the ten-fold lower lipid
concentration measured in acutely exposed fish than in all other tissues analysed was not
explained.

In a further study on PeCB using fish (Van Hoogen and Opperhuizen 1988) LBB
concentrations on dead guppies (Poecilia reticulata) were reported to be 2.54±0.59
mmol/kg after 4 days exposure and 2.11±0.39 mmol/kg after 8 days exposure. These
concentrations resulted from exposure of the fish to 135 µg/l and 100 µg/l PeCB in aqueous
solution, respectively. These concentrations were achieved by using a generator column to
maximize solubility.

Monitoring data of water concentrations in Lake Ontario were used to predict tissue levels
for 250 g sub-adult rainbow trout (Salmo gairdneri now Oncorhynchus mykiss) from
laboratory based BCFs of PeCB (Oliver and Niimi 1983). An average body burden of
0.00087 mmol PeCB /kg was found in fish from this BCF test further to exposure to 9.3±7.6
ng/l PeCB in a mixture of chlorobenzenes for 105 days. No adverse effects were observed
at this concentration.

Several invertebrate studies have provided results comparable to those in fish. A recent
study reported the CBB of PeCB associated with 50% mortality and growth of the
amphipod, Hyalella azteca, over an exposure period of 1 to 28 days (Landrum et al. 2004).
The organisms were exposed to radiolabeled PeCB in acetone (100 µl/l), diluted to between
4.8 and 920 µg/l, in a semi-static test with partial (chronic exposure) or total (short term
exposure) replacement of test solutions every 1-2 days. The highest concentrations
employed are probably well above the solubility level of the substance in water. Similar
results were found regardless of whether the body burden values were based on living or
dead organisms. Short term exposure (1 to 2 days) resulted in LBBs from 1 to 3 mmol/kg
while exposure from 10 to 28 days led to LBBs ranging from 0.4 to 0.5 mmol/kg. A
statistically significant reduction in growth rate after 28 days exposure was observed

31
starting at whole body concentrations of approximately 0.15 mmol/kg so the CBB can be
based on the no observed effect body concentration for growth of 0.1 mmol.kg.

In the marine environment the lethal body burden of PeCB was studied for juvenile sand
crabs (Portunus pelagicus) (Mortimer and Connell 1994). The crabs were exposed to
PeCB in seawater with a range of concentrations from 0.03 to 3 ppm in solvent
(concentration unspecified) for up to approximately 1000 hours. Many of the concentrations
used were above the aqueous solubility of PeCB. The crabs molted during the test and
there was an apparent discontinuity in the LC50 versus time plot after a few days, with
toxicity increasing at molt. Mortimer and Connell stated that other authors have reported
increased sensitivity to chemicals while molting although this is possibly related to
increased potential for uptake due to loss of the impermeable exoskeleton. They reported
their results as an equation for the line of LC50 versus time with results expressed in terms
of µmol/kg lipid. Extrapolating back to time = 0, the theoretical CBB was reported to be 3.24
mmol/kg wet weight. Calculating from their equation, the CBB associated with the LC50 at
7 days is 0.2 mmol/kg lipid and based on the authors’ estimate of 0.4% lipid, 0.09 mmol/kg
wet weight. This study is the only aquatic study that does not provide results that fit exactly
within the predictive range for a mode of action I chemical (non-polar narcotics). Due to the
non-standard test method the inaccurate reporting of exposure concentration, clear effect
of molting on the effect concentration, not observed elsewhere, and information and
methodological deficiencies, the results of this study should not be included to provide a
value for CBB.

One terrestrial study was found in the literature. The CBB concept was applied to toxicity
tests for PeCB with earthworms (Eisenia andrei) in different exposure systems, water, soil,
food and filter paper (Belfroid et al. 1993). Exposure concentrations causing mortality and
time to death varied dramatically, presumably according to the changing uptake rates from
the medium in which the worms were placed. For PeCB mixed in soil (at 250 to 567 mg/kg)
in an LBB ranging from 2.34 ± 0.73 mmol/kg within 7 to 10 days was observed. Exposure
of worms to a 0.2 mg/l water solution of PeCB resulted in a lethal CBB of 1.31±0.6 mmol/kg
after 1 to 7 days. Oral exposure to 60 mg/kg PeCB mixed in the manure (food) supply led
to death after 19 to 23 days. The lethal CBB was 1.46±0.71 mmol/kg. The surviving worms
averaged 0.44 mmol/kg PeCB. Earthworms were exposed on filter paper treated with 75
2
and 7.5 µg/cm of PeCB. Death was observed between 24 and 45 hours with a lethal CBB
of 1.29±0.61 mmol PeCB /kg. All results fall within the expected range for LBB for non-polar
narcotics.

Conclusion on critical body burden


While several methods, exposure routes and species with very different feeding strategies
were used to determine the lethal and critical body burden of PeCB, remarkably similar
results were found with the exception of the non-standardised and incompletely
documented marine study on crabs. Based on this evidence an LBB of 1 to 2.5 mmol/kg
(250 to 626 mg PeCB/kg) would be expected. Based on the general knowledge on
substances with a narcotic mode of action and the available evidence on PeCB, such as
the Hyalella growth/mortality study and other information discussed, an estimation of 0.1
mmol PeCB/Kg (25 mg/kg) can be tentatively proposed as a CBB for chronic effects.

32
Table 5.2. Summary of critical tissue residue data from Jarvinen and Ankley (1999).

Tissue Tissue
Residue Residue
Taxonomic Lofe Test Exposure mg/kg mmol/kg
Species group stage conc. (days) (wet) (wet) Effect Reference
Survival -
Eisenia andrei Annelid Reduced - Belfroid et al,
(Fw) earthworm Adult 0.2 mg/l 10 325.4 1.3 Death 1993
Survival -
Portunus Crustacean 0.1-100 Reduced - Mortimer and
pelagicus (Sw) sand crab Juvenile µmol/l 7 861 3.44 Death Connell, 1994
Fish
Oncorhynchus Rainbow Subadult, 0.009 Survival - Oliver and
mykiss (Fw) trout 250g µg/l 105 0.22 0.0009 No effect Niimi, 1983
Fish Survival,
Pimephales Fathead Embryo- 55.0 Growth - Carlson and
promelas (Fw) minnow Juvenile µg/l 31 380 1.5182 No effect Kosian, 1987
van Hoogen
Survival - and
Poecilia 0.40 Reduced - Opperhuizen,
reticulata (Fw) Fish Guppy Adult µmol/l 8 528 2.11 Death 1988
van Hoogen
Survival - and
Poecilia 0.54 Reduced - Opperhuizen,
reticulata (Fw) Fish Guppy Adult µmol/l 4 635.8 2.54 Death 1988
Sw= Salt water, Fw=Fresh water

33
6 Conclusions

• The physical properties and low reactivity of PeCB in the environment suggest it is
susceptible to long range transport.
• PeCB has been detected in many remote environments at low concentrations
suggesting long range transport.
• The primary sources of PeCB in the environment are combustion processes.
• PeCB concentrations have decreased in the environment.
• PeCB is bioaccumulated in aquatic organisms, but there is little biomagnification in
aquatic food webs.

7 References

Ahling, B., Bjorseth, A., Lunde, G. 1978. Formation of chlorinated hydrocarbons during
combustion of poly(vinyl chloride). Chemosphere 10: 799-806.
Aittola, J.-P., Paasivirta, J., Vattulainen, A., Sinkkonen, S., Koistinen, J., Tarhanen, J. 1996.
Formation of chloroaromatics at a metal reclamation plant and efficiency of stack filter in
their removal from emission. Chemosphere 32: 99-108.
Akimoto, Y., Nito, S., Inouye, Y. 1997. Comparative study on formations of polychlorinated
dibenzo-p-dioxins, polychlorinated dibenzofurans and related compounds in a fluidized bed
solid waste incinerator using long term used sand and fresh sand. Chemosphere 34: 791-
799.
Allen-Gil, S. M., Gubala, C. P., Wilson, R., Landers, D. H., Wade, T. L., Sericano, J. L.,
Curtis, L. R. 1997. Organochlorine pesticides and polychlorinated biphenyls ((PCBs) in
sediments and biota from four US arctic lakes. Arch. Environ. Contam. Toxicol 33: 378-387.
AMAP 2000. PCB in the Russian Federation: Inventory and proposals for priority remedial
actions. Prepared for the Arctic Council by Arctic Monitoring and Assessment Programme,
State Committee of Russian Federation for Environmental Protection and Center for
International Projects. AMAP Report 2000:3.
Arend, M. W., Jarman, W. M., Ballschmiter, K. 2001. Determination of organochlorine
POPs in Pacific halibut- and cod liver with SPE, NP-HPLC and HRGC-ECD/MSD. Preprints
of Extended Abstracts presented at the ACS National Meeting, American Chemical Society,
Division of Environmental Chemistry 41(1): 90-94.
Atkinson, R. 1990. Gas-phase tropospheric chemistry of organic compounds: a review.
Atmos. Environ. 24A: 1-41.
Aust, S. D., Benson, J. T. 1998. The fungus among us: Use of white rot fungi to biodegrade
environmental pollutants. Environmental Health Perspectives 101(3).
Bailey, R. E. 1983. Comment on “Chlorobenzenes in sediments, water and selected fish
from Lakes Superior, Huron, Erie, and Ontario”. Environ. Sci. Technol. 17: 504.
Beck, J., Hansen, K. E. 1974. The degradation of quintozene, pentachlorobenzene,
hexachlorobenzene and pentachloroaniline in soil. Pestic. Sci. 5: 41-48.
Beck, U. (1986). Chlorinated hydrocarbons. Ullmann's Encyclopedia of Industrial
Chemistry, Fifth Edition. VCH. A6: 338.
Belfroid, A., Seinen, W., van Gestel, K., Hermans, J. 1993. The acute toxicity of
chlorobenzenes for earthworms (Eisenia andrei) in different exposure systems.
Chemosphere 26(12): 2265-2277.
Beurskens, J. E. M., Dekker, C. G. C., Jonkhoff, J., Pomstra, L. 1993. Microbial
dechlorination of hexachlorobenzene in a sedimentation area of thje Rhine River.
Biogeochemistry 19: 61-81.
Beurskens, J. E. M., Dekker, C. G. C., van den Heuvel, H., Swart, M., de Wolf, J. 1994.
Dechlorination of chlorinated benzenes by an anaerobic microbial consortium that
selectively mediates the thermodynamic most favorable reactions. Environ. Sci. Technol
28(4): 701-706.

34
Bishop, C. A., Weseloh, D. V., Burgess, N. M., Struger, J., Norstrom, R. J., Turle, R.,
Logan, K. A. (1992). An atlas of contaminants in eggs of fish-eating colonial birds of the
Great Lakes (1970-1988). Technical Report Series Canadian Wildlife Service; Ontario
Region.
Boethling, R. S., Mackay, D. (2000). Handbook of Property Estimation Methods for
Chemicals Environmental and Health Sciences. Boca Raton, FL, Lewis Publishers.
Brahushi, F., Doerfler, U., Schroll, R., Munch, J. C. 2004. Stimulation of reductive
dechlorination of hexachlorobenzene in soil by inducing the native microbial activity.
Chemosphere 55(11): 1477-1484.
Breivik, K., Sweetman, A., Pacyna, J. M., Jones, K. C. 2002. Towards a global historical
emission inventory for selected PCB congeners -- a mass balance approach. 2.
Emissions. Sci Total Environ 290: 199-224.
Bright, D. A., Healey, N. 2003. Contaminant risks from biosolids land application:
contemporary organic contaminant levels in digested sewage sludge from five treatment
plants in Greater Vancouver, British Columbia. Environmental Pollution 126(1): 39-49.
Brubaker Jr., W. W., Hites, R. A. 1998. OH reaction kinetics of gas-phase alpha- and
gamma-hexachlorocyclohexane and hexachlorobenzene. Environ. Sci. Technol. 32: 766-
769.
Buehler, S. S., Basu, I., Hites, R. A. 2004. Causes of variability in pesticide and PCB
concentrations in air near the Great Lakes. Environ. Sci. Technol 38(2): 414-422.
Burkhard, L. P., Sheedy, B. R., McCauley, D. J., DeGraeve, G. M. 1997. Bioaccumulation
factors for chlorinated benzenes, chlorinated butadienes, and hexachloroethane. Environ.
Toxicol. Chem 16: 1677-1686.
Carlson, A. R., Kosian, P. A. 1987. Toxicity of chlorinated benzenes to fathead minnows
(Pimephales promelas). Arch. Environ. Contam. Toxicology 16: 129-135.
Carpenter, C., Schweer, G., Stinnet, G., Gabel, N. 1986. Exposure Assessment for
hexachlorobenzene. Prepared for USEPA, Report No. EPA-560/5-86-019.
Chan, C.-H., Williams, D. J., Neilson, M. A., Harrison, B., Archer, M. L. 2003. Spatial and
temporal trends in the concentrations of selected organochlorine pesticides (OCs) and
polynuclear aromatic hydrocarbons (PAHs) in Great Lakes basin precipitation, 1986 to
1999. J. Great Lakes Res. 29(3): 448-459.
Chan, C. H. 1993. St. Clair River head and mouth water quality monitoring, 1987-1989.
Water Poll. Res. J. Canada 28: 451-471.
Chu, W., Hunt, J. R., Jafvert, C. T. 2002. Modeling the sequential photodechlorination of
hexachlorobenzene in surfactant micelles Water Research 36(4): 843-850.
Corsolini, S., Covaci, A., Ademollo, N., Focardi, S., Schepens, P. 2006. Occurrence of
organochlorine pesticides (OCPs) and their enantiomeric signatures, and concentrations of
polybrominated diphenyl ethers (PBDEs) in the Adélie penguin food web, Antarctica
Environmental Pollution 140(2): 371-382.
Crosby, D. G., Hamadmad, N. 1971. The photoreduction of pentachlorobenzenes. J. Agr.
Food Chem. 19(6): 1171-1174.
D'Annibale, A., Ricci, M., Leonardi, V., Quarantino, D., Mincione, E., Petruccioli, M. 2005.
Degradation of aromatic hydrocarbons by white-rot fungi in a historically contaminated soil.
Biotechnology Bioengineering 90(6): 723-731.
Ding, W.-H., Aldous, K. M., Briggs, R. G., Valente, H., Hilker, D. R., Connor, S., Eadon, G.
A. 1992. Application of multivariate statistical analysis to evaluate local sources of
chlorobenzene congeners in soil samples. Chemosphere 25: 675-690.
Doering, J., Damberg, M., Gamradt, A., Oehme, M. 1992. Screening method based on the
determination of perchlorinated aromatics for surface soil contaminated by copper slag
containing high levels of polychlorinated dibenzofurans and dibenzo-p-dioxins.
Chemosphere 25(6): 755-762.
Durham, R. W., Oliver, B. G. 1983. History of Lake Ontario contamination from the Niagara
River by sediment radiodating and chlorinated hydrocarbon analysis. J. Great Lakes Res
9(2): 160-168.
Eder, G., Sturm, R., Ernst, W. 1987. Chlorinated hydrocarbons in sediments of the Elbe
River and Elbe Estuary. Chemosphere 16(10-12): 2487-2496.
Environment Canada 1993. Priority Substances List Assessment Report
Pentachlorobenzene.

35
Environment Canada 2003. Follow-up Report on Five PSL1 Substances for Which There
Was Insufficient Information to Conclude Whether the Substances Constitute a Danger to
the Environment. Accessed at www.ec.gc.ca/substances/
ese/eng/PSAP/assessment/PSL1_chlorobenzenes_followup.pdf.
Environment Canada 2005. Pentachlorobenzene (QCB) and tetrachlorobenzenes (TeCBs)
proposed risk management strategy.
Escher, B.I. and Hermens, J.L.M., 2004. Internal exposure: Linking bioavailability to effects.
Environ Sci Technol 38:455A-462A.
Fängmark, I., Strömberg, B., Berge, N., Rappe, C. 1994. Influence of postcombustion
temperature profiles on the formation of PCDDs, PCDFs, PCBzs, and PCBs in a pilot
incinerator. Environ. Sci. Technol 28: 624-629.
Fängmark, I., van Bavel, B., Marklund, S., Strömberg, B., Berge, N., Rappe, C. 1993.
Influence of combustion parameters on the formation of polychlorinated dibenzo-p-dioxins,
dibenzofurans, benzenes, and biphenyls and polyaromatic hydrocarbons in a pilot
incinerator. Environ. Sci. Technol 27: 1602-1610.
Fathepure, B. Z., Tiedje, J. M., Boyd, S. A. 1988. Reductive dechlorination of
hexachlorobenzene to tri- and dichlorobenzenes in anaerobic sewage sludge. Appl.
Environ. Microbiology 54: 327-330.
Feiler, W. (2001). Amvac Chemical Co. Personal communication.
Fisk, A. T., Holst, M., Hobson, K. A., Duffe, J., Moisey, J., Norstrom, R. J. 2002. Persistent
organochlorine contaminants and enantiomeric signatures of chiral pollutants in ringed
seals (Phoca hispida) collected on the east and west side of the Northwater Polynya,
Canadian Arctic. Arch. Environ. Contam. Toxicology 42(1): 118-126.
Gawlik, B. M., Martens, D., Schramm, K. W., Kettrup, A., Lamberty, A., Muntau, H. 2000.
On the presence of PCDD/Fs and other chlorinated hydrocarbons in the second generation
of the European Reference Soil Set--the EUROSOILS. Fresenius’ Journal of Analytical
Chemistry 368(4): 407-411.
Grannas, A. M., Miller, P. L., Chin, Y.-P. 2003. Indirect photolysis of persistent organic
pollutants by arctic dissolved organic matter. Preprints of Extended Abstracts presented at
the ACS National Meeting 43(2): 1130-1133.
Grimalt, J. O., Borghini, F., Sanchez-Hernandez, J. C., Barra, R., Torres Garcia, C. J.,
Focardi, S. 2004. Temperature dependence of the distribution of organochlorine
compounds in the mosses of the Andean Mountains. Environ. Sci. Technol. 38(20): 5386-
5392.
Gullet, B. K., Touati, A. 2003a. PCDD/F emissions from forest fire simulations. Atmos.
Environ. 37(6): 803-813.
Gullet, B. K., Touati, A. 2003b. PCDD/F emissions from burning wheat and rice field
residue. Atmos. Environ. 37(35): 4893-4899.
Gullet, B. K., Touati, A., Huwe, J., Hakk, H. 2006. PCDD and PCDF emissions from
simulated sugarcane field burning. Environ. Sci. Technol. 40(20): 6228-6234.
Havelka, T. (2006). Canadian Wildlife Service, Ontario Region. Personal communication.
Hebert, C. E., Weseloh, D. V. 2006. Adjusting for temporal change in trophic position
results in reduced rates of contaminant decline. Environ. Sci. Technol 40(18): 5624-5628.
Hendriks, A. J., Pieters, H., de Boer, J. 1998. Accumulation of metals, polycyclic
(halogenated) aromatic hydrocarbons, and biocides in zebra mussel and eel from the Rhine
and Meuse Rivers. Environ. Toxicol. Chem 17(10): 1885-1898.
Hobbs, K. E., Muir, D. C. G., Michaud, R., Béland, P., Letcher, R. J., Norstrom, R. J. 2003.
PCBs and organochlorine pesticides in blubber biopsies from free-ranging St. Lawrence
River Estuary beluga whales (Delphinapterus leucas), 1994–1998 Environmental Pollution
122(2): 291-302.
Hobbs, K. E., Muir, D. C. G., Mitchell, E. 2001. Temporal and biogeographic comparisons
of PCBs and persistent organochlorine pollutants in the blubber of fin whales from eastern
Canada in 1971–1991 Environmental Pollution 114(2): 243-254.
Hoekstra, P. F., Braune, B. M., O'Hara, T. M., Elkin, B., Solomon, K. R., Muir, D. C. G.
2003. Organochlorine contaminant and stable isotope profiles in Arctic fox (Alopex lagopus)
from the Alaskan and Canadian Arctic. Environmental Pollution 122(3): 423-433.
Hoekstra, P. F., O'Hara, T. M., Pallant, S. J., Solomon, K. R., Muir, D. C. G. 2002.
Bioaccumulation of organochlorine contaminants in bowhead whales (Balaena mysticetus)
from Barrow, Alaska. Arch. Environ. Contam. Toxicology 42(4): 497-507.

36
Hoff, R. M., Muir, D. C. G., Grift, N. P. 1992. Annual cycle of polychlorinated biphenyls and
organohalogen pesticides in air in Southern Ontario. 1. Air concentration data. Environ. Sci.
Technol 26(2): 266-275.
Hong, S. H., Yim, U. H., Shim, W. J., Oh, J. R., Lee, I. S. 2003. Horizontal and vertical
distribution of PCBs and chlorinated pesticides in sediments from Masan Bay, Korea.
Marine Pollution Bulletin 46(2): 244-253.
Jarvinen, A. W., Ankley, G. T. 1999. Linkage of effects to tissue residues: Development of a
comprehensive database for aquatic organisms exposed to inorganic and organic
chemicals, SETAC Technical Publication 99-1, Society of Environmental Toxicology and
Chemistry, Pensacola, FL, USA.
Jay, K., Stieglitz, L. 1995. Identification and quantification of volatile organic components in
emissions of waste incineration plants. Chemosphere 30: 1249-1260.
Jermyn-Gee, K., Pekarik, C., Havelka, T., Barrett, G. C., Weseloh, D. V. (2005). An atlas of
contaminants in eggs of fish-eating colonial birds of the Great Lakes (1998-2001). Techical
Report Series Number 417, Canadian Wildlife Service, Environment Canada.
Jiang, X., Martens, D., Schramm, K.-W., Kettrup, A., Xu, S. F., Wang, L. S. 2000.
Polychlorinated organic compounds (PCOCs) in waters, suspended solids and sediments
of the Yangtse River. Chemosphere 41(6): 901-905.
Kaj, L., Palm, A. 2004. Screening av hexaklorbutadien (HCBD) i miljoen. IVI Rapport
B1543. IVI Svenska Miljoeinstitutet AB.
Kaminsky, R., Kaiser, K. L. E., Hites, R. A. 1983. Fates of organic compounds from Niagara
Falls dumpsites in Lake Ontario. J. Great Lakes Res. 9(2): 183-189.
Karlsson, H., Muir, D. C. G., Teixiera, C. F., Burniston, D. A., Strachan, W. M. J., Hecky, R.
E., Mwita, J., Bootsma, H. A., Grift, N. P., Kidd, K. A., Rosenberg, B. 2000. Persistent
Chlorinated Pesticides in Air, Water, and Precipitation from the Lake Malawi Area,
Southern Africa. Environ. Sci. Technol 34(21): 4490-4495.
Kato, M., Urano, K. 2001. Convenient substitute indices to toxic equivalent quantity for
controlling and monitoring dioxins in stack gas from waste incineration facilities. Waste
Management 21(1): 55-62.
Kaune, A., Lenoir, D., Nikolai, U., Kettrup, A. 1994. Estimating concentrations of
polychlorinated dibenzo-p-dioxins and dibenzofurans in the stack gas of a hazardous waste
incinerator from concentrations of chlorinated benzenes and biphenyls. Chemosphere
29(9-11): 2083-2096.
Kim, K.-S., Hong, K.-H., Ko, Y.-H., Kim, M.-G. 2004. Emission characteristics of PCDD/Fs,
PCBs, chlorobenzenes, chlorophenols, and PAHs from polyvinylchloride combustion at
various temperatures. J. Air Waste Manage. Assoc. 54: 555-562.
King, T. L., Lee, K., Yeats, P., Alexander, R. 2003. Chlorobenzenes in snow crab
(Chionoecetes opilio): time-series monitoring following an accidental release. Bull. Environ.
Contam. Toxicology 71(3): 543-550.
Knutzen, J., Oehme, M. 1989. Polychlorinated dibenzofuran (PCDF) and dibenzo-p-dioxin
(PCDD) levels in organisms and sediments from the Frierfjord, Southern Norway.
Chemosphere 19(12): 1897-1909.
Kronimus, A., Schwarzbauer, J., Dsikowitzky, L., Heim, S., Littke, R. 2004. Anthropogenic
organic contaminants in sediments of the Lippe river, Germany. Water Research 38(16):
3473-3484.
Kuehl, D. W., Butterworth, B., Marquis, P. J. 1994. A national study of chemical residues in
fish. III: Study results. Chemosphere. 29 523-535.
Kuehl, D. W., Johnson, K. L., Butterworth, B. C., Leonard, E. N., Veith, G. D. 1981.
Quantification of octachlorostyrene and related compounds in Great Lakes fish by gas
chromatography - mass spectrometry. J. Great Lakes Res 7(3): 330-335.
Lacorte, S., Raldua, D., Martinez, E., Navarro, A., Diez, S., Bayona, J. M., Barcelo, D.
2006. Pilot survey of a broad range of priority pollutants in sediment and fish from the Ebro
river basin (NE Spain) Environmental Pollution 140(3): 471-482.
Landell Mills Market Research 1996. Landell Mills agrochemical database report for Bailey
Associates, Chlorothalonil, chlorthal and quintozene-(PCNB). Landell Mills Market
Research, Ltd., Bath, U.K.
Landrum, P. F., Steevens, J. A., Gossiaux, D. C., McElroy, M., Robinson, S., Begnoche, L.,
Chernyak, S., Hickey, J. 2004. Time-dependent lethal body residues for the toxicity of
pentachlorobenzene to Hyalella azteca. Environ. Toxicol. Chem 23(5): 1335-1343.

37
Leadley, T. A., Balch, G., Metcalfe, C. D., Lazar, R., Mazak, E., Habowsky, J., Haffner, G.
D. 1998. Chemical accumulation and toxicological stress in three brown bullhead
(Amereiurus nebulosus) populations of the Detroit River, Michigan, USA. Environ. Toxicol.
Chem 17(9): 1756-1766.
Lee, C. L., Song, H. J., Fang, M. D. 2000. Concentrations of chlorobenzenes,
hexachlorobutadiene and heavy metals in surficial sediments of Kaohsiung coast, Taiwan.
Chemosphere 41(6): 889-899.
Lee, C. L., Song, H. J., Fang, M. D. 2005. Pollution topography of chlorobenzenes and
hexachlorobutadiene in sediments along the Kaohsiung coast, Taiwan--a comparison of
two consecutive years' survey with statistical interpretation. Chemosphere 58(11): 1503-
1516.
Lemieux, P. M. 1997. Evaluation of emissions from the open burning of household waste in
barrels, Volume 1. Technical Report, USEPA Report EPA-600/R-97-134a.
Linder, R., Scotti, T., Goldstein, J., McElroy, K., Walsh, D. 1980. Acute and subchronic
toxicity of pentachlorobenzene. 4:183-196. J. Environ. Pathol. Toxicol. 4: 183-196.
Lobert, J. M., Keene, W. C., Logan, J. A., Yevich, R. 1999. Global chlorine emissions from
biomass burning: Reactive chlorine emissions inventory. J. Geophys. Res. 104(D7): 8373-
8389.
Mackay, D. 2001. Multimedia Environmental Models: The Fugacity Approach. Second
Edition, Lewis Publishers, Boca Raton. The models are explained in this book but the
actual computer programs were obtained from the Canadian Environmental Modeling
Centre at Trent University, Peterborough, Ontario, Canada,
www.trentu.ca/academic/aminss/envmodel/.
Mackay, D., Wania, F. 1995. Transport of contaminants to the Arctic: partitioning,
processes and models. Sci. Total Environ. 160/161: 25-38.
Matheus, D. R., Bononi, V. L. R., Machado, K. M. G. 2000. Biodegradation of
hexachlorobenzene by basidiomycetes in soil contaminated with industrial residues. World
J. Microbiology Biotechnology 16: 415-421.
McCarty, L. 1986. The relationship between aquatic toxicity QSARs and bioconcentration
for some organic chemicals. Environmental Toxicology and Chemistry 5: 1071-1080.
McCarty, L. S., Mackay, D. 1993. Enhancing ecotoxicological modeling and assessment.
Environ. Sci. Technol 27(9): 1719-1727.
McDonald, M. M. 1991. Toxicity studies of pentachlorobenzene (CAS No. 608-93-5) in
F344/N rats and B6C3F1 mice (feed studies). National Toxicology Program, U.S.
Department of Health and Human Services, Public Health Service, National Institutes of
Health. NIH Publication No. 91-3125.
Meharg, A. A., Wright, J., Osborn, D. 2000. Chlorobenzenes in rivers draining industrial
catchments Science Total Environment 251-252: 243-253.
Melnikov, S., Carroll , J., Gorshkov, A., Vlasov, S., Dahle, S. 2003. Snow and ice
concentrations of selected persistent pollutants in the Ob–Yenisey River watershed Sci.
Total Environ. 306(1-3): 27-37.
Mill, T., Haag, W. 1986. The environmental fate of hexachlorobenzene. IARC Scientific
Publications No. 77: 61-66.
Mortimer, M. R., Connell, D. W. 1994. Critical internal and aqueous lethal concentrations of
chlorobenzenes with the crab Portunus pelagicus. Ecotoxicology Environ. Safety 28: 298-
312.
Muir, D., Savinova, T., Savinov, V., Alexeeva, L., Potelov, V., Svetochev, V. 2003.
Bioaccumulation of PCBs and chlorinated pesticides in seals, fishes and invertebrates from
the White Sea, Russia. Science of the Total Environment 306(1-3): 111-131.
Muir, D. C. G., Grift, N. P., Lockhart, W. L., Wilkinson, P., Billeck, B. N., Brunskill., G. J.
1995. Spatial trends and historical profiles of organochlorine pesticides in Arctic lake
sediments. Sci. Total Environ. 160/161: 447–457.
Neilson, M. A. T. (2006). Environment Canada, Ontario Water Quality Monitoring. Personal
communication.
NYDEC (1998). Enhanced toxics sampling for trace organic chemicals in Lake Ontario.,
April, 1998, New York State Department of Environmental Conservation, Division of Water.
Oberg, T., Aittola, J.-P., Bergström, J. G. T. 1985. Chlorinated aromatics from the
combustion of hazardous waste. Chemosphere 14: 215-221.

38
Oberg, T., Bergstrom, J. G. T. 1985. Hexachlorobenzene as an indicator of dioxin
production from combustion. Chemosphere 14(8): 1081-1086.
Oberg, T., Bergstrom, J. G. T. 1987. Emission and chlorination pattern of PCDD/PCDF
predicted from indicator parameters. Chemosphere 16: 1221-1230.
Oliver, B. G., Bourbonniere, R. A. 1985. Chlorinated contaminants in surficial sediments of
Lakes Huron, St. Clair, and Erie: implications regarding sources along the St. Clair and
Detroit rivers. J. Great Lakes Res. 11(3): 366-372.
Oliver, B. G., Charlton, M. N., Durham, R. W. 1989. Distribution, redistribution, and
geochronology of polychlorinated biphenyl congeners and other chlorinated hydrocarbons
in Lake Ontario sediments. Environ. Sci. Technol 23(2): 200-208.
Oliver, B. G., Nicol, K. D. 1982. Chlorobenzenes in sediments, water, and selected fish
from Lakes Superior, Huron, Erie, and Ontario. Environ. Sci. Technol. 16: 532-536.
Oliver, B. G., Nicol, K. D. 1983. Authors reply on comments of R. E. Bailey Environ. Sci.
Technol 17: 505.
Oliver, B. G., Nicol, K. D. 1984. Chlorinated contaminants in the Niagara River, 1981-1983.
Sci Total Environ 39: 57-70.
Oliver, B. G., Niimi, A. J. 1983. Bioconcentration of chlorobenzenes from water by rainbow
trout: Correlations with partition coefficients and environmental residues. Environ. Sci.
Technol 17(5): 287-291.
Oliver, B. G., Niimi, A. J. 1988. Trophodynamic analysis of polychlorinated biphenyl
congeners and other chlorinated hydrocarbons in the Lake Ontario ecosystem. Environ.
Sci. Technol 22: 388-397.
Opperhuizen, A., Serne, P., Van der Steen, J. M. D. 1988. Thermodynamics of fish/water
and octan-1-ol/water partitioning of some chlorinated benzenes. Environ. Sci. Technol
22(3): 286-292.
Oren, A., Aizenshtat, Z., Chefetz, B. 2006. Persistent organic pollutants and sedimentary
organic matter properties: A case study in the Kishon River, Israel Environmental Pollution
141(2): 265-274.
Pardue, J. H., DeLaune, R. D., Adrian, D. D., Patrick Jr., W. H. 1993. Reductive
dechlorination of hexachlorobenzene in wetland soils, pages 145-152 in Sorption and
degradation of pesticides and organic chemicals in soil. Soil Science Society of America
special publication no. 32.
Pavlostathis, S. G., Prytula, M. T. 2000. Kinetics of the sequential microbial reductive
dechlorination of hexachlorobenzene. Environ. Sci. Technol. 34: 4001-4009.
Pekarik, C., Weseloh, D. V., Barrett, G. C., Simon, M., Bishop, C. A., Petit, K. E. (1998). An
Atlas of contaminants in the eggs of fish-eating colonial birds of the Great Lakes Technical
Report Series No. 322, Canadian Wildlife Service, Ontario Region.
Pereira, W. E., Rostad, C. E., Chiou, C. T., Brinton, T. I., Barber II, L. B., Demcheck, D. K.,
Demas, C. R. 1988. Contamination of estuarine water, biota, and sediment by halogenated
organic compounds: A field study. Environ. Sci. Technol. 22: 772-778.
Petit, K. E., Bishop, C. A., Weseloh, D. V., Norstrom, R. J. (1994). An atlas of contaminants
in eggs of fish-eating colonial birds of the Great Lakes (1989-1992). Technical Report
Series No. 194, Canadian Wildlife Service.
Polischuk, S. C., Norstrom, R. J., Ramsay, M. A. 2002. Body burdens and tissue
concentrations of organochlorines in polar bears (Ursus maritimus) vary during seasonal
fasts Environmental Pollution 118(1): 29-39.
Ramanand, K., Balba, M. T., Duffy, J. 1993. Reductive dehalogenation of chlorinated
benzenes and toluenes under methanogenic conditions. Appl. Environ. Microbiology 59:
3266-3272.
Sánchez-Bayo, F. 2006. Comparative acute toxicity of organic pollutants and reference
values for crustaceans. I. Branchiopoda, Copepoda and Ostracoda Environmental Pollution
139(3): 385-420.
Schuler, L. J., Landrum, P. F., Lydy, M. J. 2007. Response spectrum of fluoranthene and
pentachlorobenzene for the fathead minnow (Pimephales promelas). Environ. Tox. Chem.
26(1): 139-148.
Schwarzbauer, J., Ricking, M., Franke, S., Francke, W. 2001. Organic compounds as
contaminants of the Elbe River and its tributaries. Part 5. Halogenated organic
contaminants in sediments of the Havel and Spree Rivers (Germany). Environmental
Science and Technology 35(20): 4015-4025.

39
Scottish Chemical 1998. Internet advertisment www.scottishchemical.com accessed 1998.
Shen, L., Wania, F. 2005. Compilation, evaluation, and selection of physical-chemical
property data for organochlorine pesticides. J. Chem Eng. Data 50(3): 742-768.
Shen, L., Wania, F., Lei, Y. D., Teixeira, C., Muir, D. C. G., Bidleman, T. F. 2005.
Atmospheric distribution and long-range transport behavior of organochlorine pesticides in
North America. Environ. Sci. Technol 39(2): 409-420.
Stevens, R. J. J., Neilson, M. A. T. 1989. Inter- and intralake distributions of trace organic
contaminants in surface waters of the Great Lakes. J. Great Lakes Res 15(3): 377-393.
Strachan, W. M. J., Burniston, D. A., Williamson, M., Bohdanowicz, H. 2001. Spatial
Differences in Persistent Organochlorine Pollutant Concentrations between the Bering and
Chukchi Seas (1993) Marine Pollution Bulletin 43(1-6): 132-142.
Streeter, R. (1998). Aluminum Association, Washington DC. Personal communication.
ten Hulscher, T. E. M., Vrind, B. A., van Noort, P. C. M., Govers, H. A. J. 2002. Resistant
sorption of in situ chlorobenzenes and a polychlorinated biphenyl in river Rhine suspended
matter Chemosphere 49(10): 1231-1238.
Thompson, R. S. and K.S. Stewart, 2003. Critical body burdens: A review of the literature
and identification of experimental data requirements Report BL7549, June 2003, Brixham
Environmental Laboratory AstraZeneca CEFIC Long range Research Initiative Project.
Tiedje, J. M., Boyd, S. A., Fathepure, B. Z. 1987. Anaerobic degradation of chlorinated
aromatic hydrocarbons. in Developments in Industrial Microbiology, Volume 27, pages 117-
127.
Tiernan, T. O., Taylor, M. L., Garrett, J. H., VanNess, G. F., Solch, J. G., Deis, D. A.,
Wagel, D. J. 1983. Chlorodibenzodioxins, chlorodibenzofurans and related compounds in
the effluents from combustion processes. Chemosphere 12: 595-606.
USEPA 1988. Integrated Risk Information System, available at
www.epa.gov/iris/subst/0085.htm.
USEPA 1998. AP 42, Fifth Edition, Volume I, Chapter 1: External Combustion Sources.
Avialable at http://www.epa.gov/ttn/chief/ap42/ch01/index.html.
USEPA 2004. EPI Suite v3.12, available for download from
http://www.epa.gov/oppt/exposure/pubs/episuitedl.htm (Accessed 23 September 2006).
USEPA 2006. Toxic Releases Inventory for 2004. Available at www.epa.gov, accessed
September 2006.
van de Plassche, E., Schwegler, A., Rasenberg, M., Schoulten, G. 2001.
Pentachlorobenzene. Nomination document for inclusion in the Long Range Transport of
Air Pollutants Convention of UNECE.
van Gestel, C. A. M., Ma, W.-C., Smit, C. E. 1991. Development of QSARs in terrestrial
ecotoxicology: earthworm toxicity and soil sorption of chlorophenols, chlorobenzenes and
dichloroaniline. Sci Total Environ 109/110: 589-604.
Van Hoogen, G., Opperhuizen, A. 1988. Toxicokinetics of chlorobenzenes in fish. Environ.
Toxicol. Chem 7: 213-219.
Verreault, J., Muir, D., Norstrom, R., Stirling, I., Fisk, A., Garielsen, G., Derocher, A.,
Evans, T., Dietz, R., Sonne, C., Sandala, G., Gebbink, W., Riget, F., Born, E., Taylor, M.,
Nagy, J., Letcher, R. 2005. Chlorinated hydrocarbon contaminants and metabolites in polar
bears (Ursus Maritimus) from Alaska, Canada, East Greenland, and Svalbard: 1996-2002. .
Science Total Environment 351-352: 369-390.
Vogelgesang, J. 1986. Hexachlorobenzene, octachlorostyrene and other organochlorine
compounds in wastewater from industrial high-temperature processes involving chlorine. Z.
Wasser- Abwasser-Forsch. 19: 140-144.
Vorkamp, K., Riget, F., Glasius, M., Pecseli, M., Lebeuf, M., Muir, D. 2004.
Chlorobenzenes, chlorinated pesticides, coplanar chlorobiphenyls and other organochlorine
compounds in Greenland biota. Sci Total Environ. 331: 157-175.
Vulykh, N., Dutchak, S., Mantseva, E., Shatalov, V. (2005). EMEP contribution to the
preparatory work for the review of the CLRTAP protocol on persistent organic pollutants.
New substances: Model assessment for potential long-range transboundary atmospheric
transport and persistence of pentachlorobenzene. Moscow, Russia, Meteorological
Synthesizing Centre - East.
Wang, M.-J., Jones, K. C. 1994. Behavior and fate of chlorobenzenes in spiked and
sewage sludge-amended soil. Environ. Sci. Technol. 28: 1843-1852.

40
Wania, F., Mackay, D. 1995. A global distribution model for persistent organic chemicals.
Sci. Total Environ. 160/161: 211-232.
Weast, R. C., Ed. (1983). CRC Handbook of Chemistry and Physics. Boca Raton, Florida,
CRC Press, Inc.
Wenzel, K.-D., Hubert, A., Weissflog, L., Kühne, R., Popp, P., Kindler, A., Schüürmann, G.
2006. Influence of different emission sources on atmospheric organochlorine patterns in
Germany Atmospheric Environment 40(5): 943-957.
Westberg, H. B., Selden, A. I., Bellander, T. 1997. Emissions of some organochlorine
compounds in experimental aluminum degassing with hexachloroethane. Appl. Occup.
Environ. Hyg. 12: 178-183.
Wikstrom, E., Tysklind, M., Marklund, S. 1999. Influence of variation in combustion
conditions on the primary formation of chlorinated organic micropollutants during municipal
solid waste combustion. Environ. Sci. Technol 33(23): 4263-4269.
Williams, D. J., Kuntz, K. W., L’Italien, S., Richardson, V. (2001). Great Lakes Surveillance
Program, Organic Contaminants in the Great Lakes 1992 to1998, Intra- and Inter-lake
Spatial Distributions and Temporal Trends. Report No. EHD/ECB-OR/01-01/I,
Environmental Conservation Branch/Ontario Region, Ecosystem Health Division,
Environment Canada.
Williams, D. J., McCrea, R. C., Sverko, E. 2003. The bottom sediments of the Sir Adam
Beck Reservoir, Niagara River, Ontario. Journal of Great Lakes Research 29(4): 630-640.
Williams, D. J., Neilson, M. A. T., Merriman, J., L’Italien, S., Painter, S., Kuntz, K., El-
Shaarawi, A. H. (2000). The Niagara River upstream/downstream program 1986/87-
1996/97. Concentrations, Loads, Trends. Report No. EHD/ECB-OR/00-01/I, Environmental
Conservation Branch/Ontario Region, Ecosystem Health Division, Environment Canada.
Witter, B., Francke, W., Francke, S., Knauth, H.-D., Miehlich, G. 1998. Distribution and
mobility of organic micropollutants in River Elbe floodplains. Chemosphere 37: 63-78.
World Coal Institute 2006. Coal facts.
http://www.worldcoal.org/pages/content/index.asp?PageID=188 accessed October
2006.
Zimmerman, R., Blumenstock, M., Heger, H. J., Schramm, K.-W., Kettrup, A. 2001.
Emission of nonchlorinated and chlorinated aromatics in the flue gas of incineration plants
during and after transient disturbances of combustion conditions: Delayed emission effects.
Environ. Sci. Technol. 35: 1019-1030.
Zoumis, T., Schmidt, A., Grigorova, L., Calmano, W. 2001. Contaminants in sediments:
remobilization and demobilization. Science of the Total Environment 266(1-3): 195-202.
Zurbruegg, C. 2003. Urban solid waste management in low-income countries of Asia. How
to cope with the garbage crisis. Available at
http://www.sandec.ch/SolidWaste/Documents/04-SW-Management/USWM-Asia.pdf.
Accessed September 2006.

41
Euro Chlor
The voice of the European chlorine industry, Euro Chlor plays a key
communication and representation role on behalf of its members,
listening and responding to society’s concerns about the
sustainability of chlorine chemistry.

Euro Chlor helps members improve safety standards whilst


conducting science, advocacy and communications programmes.
The Brussels-based federation was founded in its current form in
1989 and has 120 members comprising 39 chlorine producers, 44
associates and 37 technical correspondents. Euro Chlor speaks on
behalf of 97% of chlorine production in the EU-27.

Euro Chlor • Avenue E. van Nieuwenhuyse 4, box 2 • B-1160 Brussels, Belgium


Tel: +32 2 676 72 11 | Fax: +32 2 676 72 41
Email: eurochlor@cefic.be | www.eurochlor.org

You might also like