Sie sind auf Seite 1von 151

Distribution and Sources of Polycyclic Aromatic Hydrocarbons

in Sediments, Suspended Particulate Matter and Waters from


the Siak River System, Estuary and Coastal Area of Sumatra,
Indonesia

A dissertation submitted for the degree of


- Doktor der Naturwissenschaften (Dr. rer. nat.)
at the Faculty of Biology/Chemistry
the University of Bremen, Germany

Presented by:
Muhammad Lukman

Bremen, 2010

Distribution and Sources of Polycyclic Aromatic Hydrocarbons in


Sediments, Suspended Particulate Matter and Waters from the Siak River
System, Estuary and Coastal Area of Sumatra, Indonesia

A dissertation submitted for the degree of Doktor der Naturwissenschaften (Dr. rer. nat.)
at the Faculty of Biology/Chemistry, the University of Bremen, Germany

Presented by:
Muhammad Lukman

Referees

1. Professor Dr. Wolfgang Balzer


2. Professor Dr. Wolfram Thiemann

Examiners

1. Professor Dr. Gerhard Kattner


2. Professor Dr. Venugopalan Ittekkot

Date of Colloquium: 29 March 2010

ACKNOWLEDGMENTS

First of all, I would like to express my sincerely and great gratitude to Prof. Dr. Wolfgang
Balzer (FB 2, Marine Chemistry, University of Bremen) for his remarkable role in pouring me
with lots of insight, motivation, and supervision throughout my PhD work. I do very much
appreciate for his willingness to offer and provide me the possibility to do PhD in his working
group as well as to join the SPICE Cluster 3.1. (Science for the Protection of Indonesian Coastal
Ecosystem) Project in Riau, Sumatra, Indonesia. Secondly, I am grateful to DAAD (Deutscher
Akademischer Austausch Dienst or German Academic Exchange Service) in providing me a
great support to do my PhD in Germany during the period of 2004 2007.
I would like to thank Prof. Dr. Wolfram Thiemann as the second referee, Prof. Dr.
Gerhard Kattner and Prof. Dr. Venugopalan Ittekkot as the examiners, and all the colleagues
and those previous colleagues in the marine chemistry working group, University of Bremen:
Dr. Uwe Schler, Dr. Wolfgang Barkmann, Immo Becker, Olaf Wilhelm, Timo Daberkow,
Xiaoliang Tang, Jun Fu, Bjrn Bach, Dominique Schobes, Sonia Tambou for their valuable
assistances and discussion in laboratory and analytical aspects, as well as Mrs. Ute Wolpmann
who faithfully helps me with administrative matters during my stay in Bremen.
I would like to extent my appreciation to all my SPICE Cluster 3.1. colleagues: Dr. Tim
Rixen, Dr. Antje Baum (ZMT Bremen), Dr. Herbert Siegel (IOW Warnemuende), Dr. Thomas
Pohlmann (University of Hamburg), Dr. Ralf Woestmann (Terramare, Wilhemshaven), and to
ZMT staffs Dorothee Dasbach and Matthias Birkicht for all their countless assistances, advices
and critics, as well as to Nathan Giles for improving the English.
Also, I would like to thank to Prof. Dr. Gerd Liebezeit (Terramare, University of
Oldenburg) to all insights, advices and discussion. Also, to all Indonesian SPICE colleagues in
University of Riau, Riau during sampling campaigns, particularly Dr. Joko Samiaji, Dr.
Christine Jose, Dewi Kristina, friends at Hasanuddin University, Makassar, and many others. I
thank you so much for your supports.
Last but not least, very special thanks I dedicate to my wife, Rahmawati Yusuf, to my
family my Mother, Father, Sister and Brothers -, and to my all relatives - who always inspire
me during hard time. Finally, I would like to dedicate this work to my country Indonesia and to
those who are fond of better environment.

Kurzfassung
Die vorliegende Arbeit untersucht Ursprung und Verteilung von Polycyclischen Aromatischen
Kohlenwasserstoffen (PAKs) als Indikator fr anthropogene Verschmutzung in den Ksten- und
Flussregionen der Insel Sumatra in Indonesien. Im Vordergrund steht dabei die Analyse der 16 PAKPriorittsverbindungen, von denen Referenzmaterial gem der USA-EPA priority pollutants Liste
vorliegt. Untersucht wurden Proben in der Lsung und von Oberflchensedimenten und
Schwebstoffen (SPM) des Siak-Flusses, seiner Flussmndung und des Riau Kstenbereichs in
Sumatra.
Die Quantifizierung der PAKs wurde unter Einsatz eines Hochleistungs-FlssigkeitsChromatographen (RP-C18-HPLC) mit UV- und Fluoreszenz-Detektion durchgefhrt. Die
Analysenmethode beinhaltete eine Reihe von Probenahmetechniken fr die individuellen Phasen
(Sediment, SPM, Lsung), Probenaufbereitung, Extraktion, Aufarbeitung und HPLC-Quantifizierung.
Die Untersuchung der PAKs im Sediment konzentrierte sich auf zwei Grenklassen: Grobfraktion
(Sand) 2 mm 63 m und Feinfraktion (Schlick) < 63 m. Die Schwebstoffe wurden ber 0,7 m
Glasfaserfilter (GF/F) herausgefiltert. Die PAKs der Lsung wurden dann mittels eines Octadecyl
Festphasen Extraktionssystems (SPE) extrahiert. Die Qualittskontrolle beinhaltete den Gebrauch von
Blindwerten und Ersatzstandards, um Genauigkeit und Effizienz der Analyse und die
Reproduzierbarkeit der Ergebnisse zu gewhrleisten. Die Aufteilung in die verschiedenen
Stoffquellen der PAK-Verbindungen wurde unter Einbeziehung bekannter Indexe der
Molekulargewichte und spezifischen Isomer-Verhltnissen ausgefhrt.
Die Untersuchungen ergaben, dass sowohl der Flusslauf als auch Flussmndung und Kstenbereich
des Siak-Flusses erheblich mit PAKs belastet sind. Die Untersuchungsergebnisse weisen auf krftige
pyrogene Stoffquellen hin, insbesondere auf die Verbrennung von Biomasse und Erdl. Sie knnen
daher als Nachweis fr grorumiges, lnger anhaltendes und intensives Verbrennen
landwirtschaftlicher Nutzflchen im Zusammenhang mit krftigen Wald- und Torf-Feuern, die ber
die letzten Jahrzehnte stattfanden, angesehen werden. In diesen von Buschfeuern heimgesuchten
Gegenden bilden die PAK-Verteilungen zwischen Grob- und Feinfraktion an der Kste und in der
Flussmndung ein deutliches Muster, das von den Verteilungen blicher Kstenbereiche deutlich
abweicht. Ein Vergleich der PAK-Verteilungen in den beiden Grenfraktionen in den Sedimenten
der Siak-Kste mit den Kstenbereichen von Wenchang und Wanquan in China deutet darauf hin,
dass die PAKs der Kstengewsser um Sumatra hauptschlich mit den hohen kohleartigen
Materialien wie Ru und verbranntem Torf assoziiert sind. Wie die Untersuchungen zeigen, knnen
auch andere relevante Stoffquellen wie andauernde Erdlverschmutzung in den Gewssern um die
Stdte, in den industriellen Vororten von Perawang, der lstadt Dumai und der Erdlraffinerien im
Gebiet der Flussmndung fr die Belastung mit PAKs verantwortlich sein.
Als Zusammenfassung der Ergebnisse der Einzeluntersuchungen wurden die folgenden drei
Manuskripte erstellt, die an begutachtete wissenschaftliche Zeitschriften zu versenden sind.

PAKs im Sediment (Manuskript I, Kapitel IV)


Die PAK-Gehalte (Summe der 16 Standard-Verbindungen) in der Sedimentfraktion aller beprobten
Gebiete bewegen sich zwischen 0,13 und 5,47 g/g Trockengewicht (TG). In der Grobfraktion (Sand)
wurden mit Werten zwischen 0,16 bis 5,47 g/g TG (median m = 0,84) weitaus hhere
Konzentrationen gefunden (etwa um einen Faktor 2) als in der Feinfraktion (Schlick) mit Werten
zwischen 0,13 und 1,31 g/g TG (m = 0,52). In der Grobfraktion ist die Anreicherung unerwartet, da
diese in der Regel wegen der greren Oberflche pro Masseneinheit in der Feinfraktion zu erwarten
ist. Ein hnliches Muster wurde fr das organische Material beobachtet. So variiert der Anteil des
organischen Kohlenstoffs an der Gesamtmasse der Grobfraktion zwischen 0,01% und 24%, whrend

sich der Anteil in der Feinfraktion zwischen 0,34% und 3,7% bewegt. Ebenfalls war ein nahezu
linearer Zusammenhang zwischen PAK und organischem Kohlenstoff nur in der Grobfraktion zu
erkennen. Aus diesen Untersuchungsergebnissen kann geschlossen werden, dass eine bestimmte Sorte
organischen Materials fr die Affinitt zwischen Kohlenstoff und PAKs verantwortlich ist, nmlich
vaskulre Pflanzenreste, Torf und Ru, wie auch in hnlichen Untersuchungen festgestellt wurde.
Entlang des Flusslaufes in Richtung Flussmndung konnte kein klares Muster in den PAK-Gehalten
festgestellt werden, auffallend sind nur die Anreicherungen in den urbanen und industriellen
Gebieten. Die weitgehend hohen Molekulargewichte und die Molekularverhltnisse lassen auf
pyrogene Spurenstoffquellen schlieen, insbesondere auf die Verbrennung von Biomasse und Erdl.
Die PAK-Gehalte gelangen daher ber den Land- und Luftweg in die Gewsser.

PAKs in der Lsung und in den Schwebstoffen (Manuskript II, Kapitel V)


Die gemessenen PAK-Gehalte in der Lsung bewegen sich zwischen 0,13 und 5,14 g/L im
Flusslauf, zwischen 0,32 und 0,62 g/L in dem stuar und zwischen 0,12 und 0,13 g/L im
Kstenbereich. In Richtung Kste nimmt die mittlere Konzentration um einen Faktor 3 ab. Die
hchsten PAK-Gehalte wurden an der Einmndung des Mandau Flusses in den Siak Fluss gemessen.
Die PAKs wurden durch 2-, 3- und 4-Ring-Aromate dominiert. Die PAK-Gehalte in den
Schwebstoffen variieren zwischen 1,48 bis 59.1 g/g im Flusslauf, zwischen 0,16 und 7,67 g/g in
der Flussmndung und zwischen 0,33 und 10,2 g/g in der Kstenregion. Auf das Volumen bezogen
bewegen sich die PAK-Gehalte im SPM jeweils zwischen 0,06 und 0,69 g/L, 0,03 und 0,29 g/L
und zwischen 0,01 und 0,15 g/L. Die PAK-Gehalte nehmen im Allgemeinen in Richtung Kste ab,
was auf eine Ablagerung im Sediment und/oder Verdnnung mit Seewasser zurckzufhren ist. Eine
Anreicherung mit PAKs findet sowohl in der Trockenzeit als auch in der Regenzeit statt, die durch
verschiedene Ring-Gren gekennzeichnet sind. Ebenfalls lassen sich dadurch die verschiedenen
Transportwege der PAKs erschlieen. Die Anreicherungen der PAKs in der Regenzeit knnen durch
eine Zunahme des Oberflchenabflusses entstehen, whrend in der Trockenzeit die Anreicherung
wahrscheinlich durch die Atmosphre stattfindet.
Eine Partitionierung der PAKs zwischen Schwebstoffen und Lsung wurde durchgefhrt, um ein
besseres Verstndnis des Verlaufs der Anreicherung dieser Schadstoffe in den Gewssern zu
bekommen. Die gemessenen Partitionskoeffizienten (KD) zeigen eine betrchtliche Streuung zwischen
den beprobten Stationen. Der mittlere KD-Wert im Flusslauf und in der Flussmndung variiert
zwischen 4 und 5 auf der logarithmischen Skala, an der Kste hat er dann einen Wert von 6 erreicht.
Normalisiert auf den organischen Kohlenstoff bewegt sich der Partitionskoeffizient (KOC) zwischen 2
und 4 im Fluss und an der Mndung und zwischen 3 und 5 an der Kste. Diese Variation deutet auf
eine unterschiedliche Qualitt des partikulren organischen Materials hin, in der Ruablagerungen
eine bedeutende Rolle spielen knnen. Anscheinend erhht die Zunahme des Salzgehalts und
eventuell des pH-Wertes den Koeffizienten KD. Auf der anderen Seite hat die Zunahme des gelsten
organischen Kohlenstoffs (DOC) einen umgekehrten Effekt. Dies deutet auf eine wichtige Rolle des
DOC fr die Erhaltung schwerer PAKs in der Lsung hin, was auch eine Erleichterung ihres
Transports beinhaltet.

Ein Vergleich zwischen den Messungen am Siak-stuar (torfhaltiger Boden) und denen
an den Wenchang und Wanquan Flussmndungen (sandiger Boden) (Manuskript III,
Kapitel VI)
Die Verteilung der PAK-Konzentrationen in den Sedimentfraktionen der Siak-Flussmndung und des
Kstengebietes wurde mit den Messungen an den Wenchang (WW)- und Wanquan (WQ)Flussmndungen in Hainan, China verglichen. Die Gebiete zeichnen sich durch ihre unterschiedliche
Bodenformation aus: torfhaltig am Siak-Fluss und eher mineralisch oder huminstoffarmer Boden auf
Hainan. Wie oben erwhnt, sind die PAK-Konzentrationen in den Siak-Sedimenten durch einen
hohen Anteil von PAKs und organischem Material in der Grobfraktion charakterisiert. Dieses Muster
wurde nicht in den WW/WQ-Sedimenten gefunden, in denen die PAKs hauptschlich in der
Feinfraktion zu finden sind. Die gemessenen PAK-Konzentrationen der 15 US EPA priority pollutants
(ausgenommen Acenaphthylen) bewegen sich in der Siak-Flussmndung zwischen 0,13 g/g TG und

1,83 g/g TG (median m = 0,69 g/g TG) in der Grobfraktion und zwischen 0,09 g/g TG und 0,43
g/g TG (m = 0,20 g/g TG) in der Feinfraktion. Dagegen wurden PAK-Werte in der Grobfraktion
der WW/WQ-Sedimente zwischen 0,11 g/g TG und 0,39 g/g TG (m = 0,19 g/g TG) und in der
Feinfraktion zwischen 0,09 g/g TG und 0,68 g/g TG (m = 0,47 g/g TG) gefunden. Die relativ
hohe PAK-Konzentration in der Grobfraktion der torfreichen Bodenformation von Sumatra konnte
den hohen Konzentrationen der kohlenstoffhaltigen Materialien wie Ru, Torf und
Pflanzenablagerungen zugeordnet werden, welche zur Anreicherung der PAKs an den Materialien
beitragen. In den WW/WQ-Sedimenten wurden diese hohen Kohlenstoffkonzentrationen nicht
gefunden. In beiden Gebieten gab es keine Unterschiede in den Molekularverteilungen zwischen
Grob- und Feinfraktion, was als Hinweis gewertet werden kann, dass die PAK-Verunreinigungen von
hnlichen Quellen stammen. Dagegen deuten die Isomerenverhltnisse, die fr die
Quellenbestimmung hinzugezogen wurden, darauf hin, dass die PAKs der Siak-Flussmndung durch
Buschfeuer und hnliche Verbrennungen entstanden sind, wohingegen die PAKs der WW/WQSedimente aus der Kohle- und lverbrennung stammen. Die Konzentration der gelsten PAKs in den
kstennahen stuaren von WW/WQ war relativ gering. Die PAKs waren zwischen 7,36 und 16,2 ng
L-1. Im Gegensatz dazu war die Konzentration im Siak-stuar sowie an der Kste erheblich hher. Sie
betrug zwischen 121 und 619 ng L-1. Huminstoffe in den berliegenden Wasserschichten spielen
ebenfalls eine Rolle fr die Verteilung der PAKs in den Sedimentfraktionen. So enthlt die SiakFlussmndung und die Kstenregion signifikante Mengen an DOC, die in den beprobten WW/WQGebieten nicht gefunden wurden (Balzer, unverffentlichte Daten). Die Verteilungskoeffizienten
Sediment-Wasser der einzelnen PAKs im Siak-stuar ergaben Werte (auf der logarithmischen Skala)
zwischen 2,22 und 5,58 (Median m = 3,37) fr die Grobfraktion und zwischen 1,53 und 5,03 (m =
3,01) in der Feinfraktion. So liegen die KD-Werte im Siak-stuar niedriger als in den WW/WQstuaren, zwischen 1,12 und 5,89 (m = 3,93) fr die Grobfraktion und zwischen 2,58 und 5,85 (m =
4,40) fr die Feinfraktion. Dies lsst vermuten, dass die hohe DOC-Konzentration in der Wassersule
des Siak eine Adsorption an die organischen Sedimentbestandteile behindert.

SUMMARY
This study examined the overall distribution and sources of polycyclic aromatic hydrocarbons
(PAHs), particularly the 16 parent PAHs of the US Environmental Protection Agency priority
pollutant, as an indicator for anthropogenic pollution, in the surface sediments, suspended
particulate matter (SPM) and water solution of the Siak river system, its estuary and the Riau coast,
Sumatra, Indonesia.
The PAHs were determined by high performance liquid chromatography with reverse phase
octadecyl column (RP-C18-HPLC) using ultraviolet and programmable fluorescence detectors.
Method analysis included various sampling techniques for the individual phases, sample preparation,
extraction, work-up procedures and HPLC quantification. Analysis of PAHs in sediment was
focused on the content distribution among two size-fractions: sand/coarse (2 mm - 63 m) and
mud/fine (< 63 m). The particulate PAHs were those embedded in suspended materials retained
by 0.7 m glass fiber filter (GF/F). Dissolved PAHs were obtained from the filtered water-solution,
using an octadecyl solid phase extraction (SPE) system. Quality control measures included the use
of procedural blanks and surrogate standards in order to optimize and validate procedural accuracy,
efficiency and the reproducibility of results. Source apportionment of PAHs was carried out by
applying existing indices of molecular weights and specific isomer ratios.
In general, the results show that PAHs significantly impact the Siak river, the estuary and the
coastal waters. Source apportionment indicated intense signatures of pyrogenic sources, particularly
biomass burnings and petroleum combustion. The results might be evidence of the effects of
widespread, long-term and intense agricultural burnings coupled with multitudinous forest/peat
swamp fires which have occurred frequently over the last decades. In such burning-affected
estuaries and coastal waters, distribution of PAH between the size-fraction in sediments showed
distinctive patterns to those of other coastal areas. Comparison of distribution of PAH in the coarse
and fine fractions between the Siak Sumatra and the Wenchang and Wanquan coastal estuaries of
Hainan China indicates that PAH transferred to the coastal waters of Sumatra were mostly
associated with high carbonaceous materials the burning product particles such as black carbon
and peat. However, the apportionment also showed that another relevant source of PAHs was
chronic petroleum pollution centred in the waters around cities, the industrial estates of Perawang,
oil city of Dumai, and the oil refinery located in the estuary area.
Summarizing the different results the following three manuscripts were produced which are
sent to scientific journals with a referee system.

Sedimentary PAHs (Manuscript I, Chapter IV)


The PAHs for the sedimentary fractions in all sampled areas ranged between values of 0.13
to 5.47 g g-1 dry weight (d.w.) sediment. The PAHs in the sand fraction ranged from 0.16 to 5.47
g g-1 d.w. (median m = 0.84). In general, the sand fraction contained PAH levels higher by a factor
of 2 as compared to those found in the mud fraction that ranged from 0.13 to 1.31 g g-1 d.w. (m =
0.52). The enrichment of PAHs in the sand fraction was quite astonishing, especially since we
assumed that the fine fraction would generally evidence much higher levels of contaminants, due to
its large surface area per unit mass for adsorption. The same pattern of enrichment was shown by
the organic matters. The organic carbon (OC) contents varied greatly from 0.01% to 24% in the
sand, but only slightly in the mud from 0.34% to 3.70%. A linear relationship between the PAH and

the OC was shown only by the sand fraction. These all evidences lead us to the assumption that a
specific kind of organic matter should be responsible for high affinity for PAHs. Thus, it is most
likely materials such as black carbon, vascular plant debris, and peat as figured out by many other
studies. The spatial distribution showed no clear pattern in distance to the river mouth. But,
increased content of the PAHs was centred at urban and industrial areas. The high molecular weight
compounds were widespread predominant and the molecular ratios for source apportionment
provide further evidences for pyrogenic sources: biomass and petroleum combustions. PAHs were
delivered to those aquatic systems by both land-water and air-water transports.

Dissolved and Particulate PAHs (Manuscript II, Chapter V)


Dissolved PAHs ranged from 0.13 to 5.14 g L-1, 0.32 to 0.62 g L-1, and 0.12 to 0.13 g
L-1 in the Siak River, its estuary and the coastal areas, respectively. The mean concentration
decreased by a factor of 3 towards the coast. The highest concentration was observed in the water at
the confluence of the black water Mandau River, a Siak tributary. The PAHs were dominated by 2-,
3-, and 4-ring compounds. The PAHs in the SPM varied greatly from 1.48 to 59.1 g g-1, 0.16 to
7.67 g g-1, and 0.33 to 10.2 g g-1 for the Siak river, its estuary and the coast, respectively. In
volume basis, the concentration of particulate PAHs ranged from 0.06 to 0.69 g L-1, 0.03 to 0.29
g L-1, and 0.01 to 0.15 g L-1 in the River, the estuary and the coast, respectively. The PAHs
generally decreased towards the coast suggesting an entrapment and/or dilution effect of sea water.
PAH enrichment occurred in both wet and dry seasons characterized by different ring-size
dominance. It suggests different mode of transport by which PAH were integrated into the aquatic
environments. The accumulation of PAHs in the river during rainy season could be attributed to an
increasing land-water surface runoff. Meanwhile, in the dry season, the enrichment was most likely
caused by atmospheric deposition.
Partitioning of the PAHs between SPM and water solution was evaluated to understand the
fate of these contaminants in the given aquatic systems. Measured partition coefficient (KD) showed
a considerable variation between the sampling locations. The mean KD values in the River and the
estuary ranged from 4 to 5 on the logarithmic scale, while in the coast was 6. The mean values of
organic-carbon normalized partition coefficient (KOC) ranged from 2 to 4 on the log scale in the
River and the estuary, whereas from 3 to 5 in the coast. This variation suggests a different quality of
particulate organic matter, in which black carbon might play a significant role. The increase in
salinity and possibly also in the pH apparently turns to increase the KD, but enriched DOC affects
negatively. It indicates an important role of DOC in sustaining heavier PAHs in the dissolved phase,
including a facilitation of their transport.

A comparison between peatland aquatic system of Siak Estuary and a non-peatland


aquatic system of Wenchang and Wanquan Coastal Estuaries (Manuscript III, Chapter
VI)
Distribution of PAHs in the two grain-size fractions of the Siak Estuary and the Coast was
compared with those from non-peatland systems of Wenchang and Wanquan (WW/WQ) Coastal
Estuaries, Hainan, China. As described earlier, the PAHs in the Siak sediments were generally
characterized by high content of PAHs and organic matter in the coarse fraction. This kind of
distribution was not confirmed by the WW/WQ sediments which in most cases PAHs enriched in
the fine sediments. The level of the total 15 US EPA priority pollutants excluding acenaphthylene in

the Siak estuary and the coast ranged from 0.13 g/g d.w. to 1.83 g/g d.w. (median m = 0.69 g/g
d.w.) in the coarse fraction, and from 0.09 g/g d.w. to 0.43 g/g d.w. (m = 0.20 g/g d.w.) in the
fine fraction, while the the WW/WQ sediments PAHs ranged from 0.11 g/g to 0.39 g/g d.w.
(median m = 0.19 g/g d.w.), and from 0.09 g/g to 0.68 g/g d.w. (m = 0.47 g/g d.w.) in the
coarse and the fine fractions, respectively. The high content of PAHs in the coarse fraction of the
degraded peatland system of Sumatra was due to the existence of high carbon content carbonaceous
materials i.e. black carbon, peat, and plant debris acting as strong sorbents for PAHs, which were
not found in the WW/WQ sediments of Hainan. In both compared areas, there were no differences
in molecular distribution between the fractions suggesting that PAH contamination stemmed from
similar sources. Furthermore, the isomeric ratios used for source apportionment indicated that the
PAHs found in the Siak basin had mostly been generated through biomass burnings, whereas PAHs
analyzed in the WW/WQ sediments from Hainan Island stemmed from a mixture of coal and
petroleum combustion. The concentration of dissolved PAHs in WW/WQ coastal estuary was
relatively low. The PAHs ranged from 7.36 to 16.2 ng L-1. In contrast, the level of the PAHs
dissolved in the Siak estuary and the coast ranged from 121 to 619 ng L-1. Humic substance in the
overlaying water plays a role in distribution of PAH in the sediment fractions. The Siak estuary and
the coast contained significant amount of DOC compared to the Wenchang and Wanquan coastal
estuaries (Balzer, unpublished data). DOC in the Siak water may sustain the PAHs in the water
column impeding their association onto the sediment organic matter as shown by sediment-water
distribution coefficient (KD). The sediment-water distribution coefficients (KD) values of the
individual PAH in the Siak estuary ranged in logarithmic value from 2.22 to 5.58 (median m = 3.37)
for the coarse fraction, and from 1.53 to 5.03 (m = 3.01) for the fine fraction. The KD values in the
Siak estuary are generally lower than those of WW/WQ estuaries, which greatly ranged from 1.12
to 5.89 (m = 3.93) in the coarse fraction, and from 2.58 to 5.85 (m = 4.40) in the fine fraction. It
suggests that high DOC in the Siak water may sustain the PAHs in the water column impeding their
association with the sedimentary organic matter.

TABLE OF CONTENTS
I.

INTRODUCTION .............................................................................................................1

1.1. Polycyclic aromatic hydrocarbons: definition and physio-chemical characteristics ...........2


1.1.1. Definition ........................................................................................................................2
1.1.2. Selected physio-chemical characteristics........................................................................3
1.2. PAH contamination in the aquatic environments: background to environmental problems5
1.2.1. Toxicity of PAHs ............................................................................................................5
1.2.2. Bioconcentration and magnification in aquatic food webs .............................................7
1.2.3. Persistence, low degradation rates and pollution indicators ...........................................9
1.3. The relevance of rivers, estuaries and coastal areas in Indonesia ......................................10
1.3.1. A Perspective on the global pollution dispersal............................................................10
1.3.2. Environmental settings of the study areas: An Overview.............................................13
1.4. Study Objectives ................................................................................................................19
II.

SOURCES, DISTRIBUTION AND FATE OF PAHs IN AQUATIC


COMPARTMENTS: A SHORT REVIEW ...................................................................20

2.1. Introduction........................................................................................................................20
2.2. Source and Signatures ........................................................................................................20
2.2.1. Natural PAHs ................................................................................................................20
2.2.2. Pyrogenic PAHs ...........................................................................................................22
2.2.3. Petrogenic PAHs...........................................................................................................29
2.2.4. Source Apportionment ..................................................................................................32
2.3. Distribution in Aquatic Compartments ..............................................................................39
2.3.1. Surface Sediment and Grain Size Fractions .................................................................39
2.3.2. Suspended Particulate Matter and Water ......................................................................42
2.3.3. Water Solution as dissolved PAHs ...............................................................................43
2.4. The fate of PAHs in the water: a partitioning concept and the role of natural organic
matter .................................................................................................................................44
III. METHODS OF ANALYSIS ...........................................................................................47
3.1. Introduction........................................................................................................................47
3.2. Sample Collection and Treatments for PAH .....................................................................47
3.2.1. Surface Sediment and Size Fractionation .....................................................................47
3.2.2. Suspended Particulate Matter (SPM)............................................................................48
3.2.3. Solid phase extraction (SPE) for pre-concentration of dissolved PAHs.......................48
3.3. Determination of polycyclic aromatic hydrocarbons using high performance liquid
chromatography coupled with ultraviolet and fluorescence detectors (HPLC UV/FLD)..50
3.3.1. Soxhlet Extraction of sediment and SPM .....................................................................50
3.3.2. Extract Working-Up .....................................................................................................51
3.3.3. Elution of the SPE Cartridges for dissolved PAHs.......................................................52
3.3.4. PAH determination: High Performance Liquid Chromatography with ultraviolet and
fluorescence detectors (HPLC UV/FLDs) ....................................................................53
3.4. Quality Controls.................................................................................................................57
References (Chapter I III) ......................................................................................................59
IV. DISTRIBUTION AND SOURCE OF POLYCYCLIC AROMATIC
HYDROCARBONS (PAHs) IN SURFACE SEDIMENTS FROM THE SIAK
RIVER, ITS ESTUARY AND THE ADJACENT COASTAL AREA OF RIAU
PROVINCE, INDONESIA .............................................................................................69
Abstract ........................................................................................................................................69
4.1. Introduction........................................................................................................................70
4.2. Study Area and Methods....................................................................................................71

4.2.1. Study Area & sampling locations .................................................................................71


4.2.2. Sample Collection and Fractionation............................................................................72
4.2.3. Analytical Methods.......................................................................................................72
4.3. Results & Discussion .........................................................................................................74
4.3.1. Geochemistry of sediment fractions .............................................................................74
4.3.2. Content and Distribution of PAH .................................................................................75
4.3.3. PAH & OC Relationship ..............................................................................................77
4.3.4. Relative Composition of PAHs ....................................................................................79
4.3.5. Source Apportionment ..................................................................................................80
4.4. Conclusion .........................................................................................................................82
V.

POLYCYCLIC AROMATIC HYDROCARBONS IN SURFACE WATERS OF


THE SIAK RIVER, ITS ESTUARY AND THE COASTAL AREAS OF RIAU
PROVINCE, INDONESIA: DISTRIBUTION AND SOURCES .............................86

Abstract ........................................................................................................................................86
5.1. Introduction........................................................................................................................87
5.2. Materials and Methods.......................................................................................................88
5.2.1. Study Areas and Sampling Locations ...........................................................................88
5.2.2. Sample Collection and Treatments ...............................................................................89
5.2.3. Extraction & Work-Up Procedures for particulate PAH ..............................................89
5.2.4. Solid Phase Extraction (SPE) system for dissolved PAH.............................................89
5.2.5. Determination of PAH by HPLC UV/FLD ..................................................................90
5.3. Results and Discussion ......................................................................................................91
5.3.1. Dissolved PAHs ............................................................................................................91
5.3.2. PAHs in the SPM ..........................................................................................................92
5.3.3. Distribution Coefficient of PAHs between SPM and Water Solution ..........................96
5.3.4. Source apportionment ...................................................................................................98
5.4. Conclusion .......................................................................................................................100
VI. A COMPARISON OF POLYCYCLIC AROMATIC HYDROCARBONS (PAHs)
IN PEATLAND AND NON-PEATLAND AQUATIC SYSTEM SURFACE
SEDIMENTS: A STUDY OF THE SIAK ESTUARY, SUMATRA, INDONESIA
AND OF THE WENCHANG AND WANQUAN ESTUARIES, HAINAN ISLAND,
CHINA ............................................................................................................................104
Abstract ......................................................................................................................................104
6.1. Introduction......................................................................................................................105
6.2. Materials and Methods.....................................................................................................106
6.2.1. Study areas, Sample Collection and Fractionation .....................................................106
6.2.2. Determination of PAHs ..............................................................................................108
6.2.3. Determination of Sedimentary Organic Matter ..........................................................110
6.3. Results..............................................................................................................................110
6.3.1. PAHs and Organic Matter in Sedimentary Size Fractions..........................................110
6.3.2. Relative Composition of PAHs ..................................................................................112
6.3.3. Source Apportionment ................................................................................................113
6.3.4. Sediment-Water Distribution Coefficient ...................................................................114
6.4. Discussion ........................................................................................................................115
6.5. Conclusion .......................................................................................................................119
APPENDIXES ..........................................................................................................................124

I.

INTRODUCTION
An examination the distribution and sources of polycyclic aromatic hydrocarbons (PAHs)

in aquatic systems (rivers, estuaries and coastal waters) tackles key factors, which concern both
the overall health of the environment and also these pollutants' behavior before they are
integrated into the open ocean. During the last three decades, there has been a mounting body of
scientific literature about PAHs in aquatic environments from many parts of the world. These
studies indicate substantial and widespread concerns about these compounds as many persistent
organic pollutants (POP) such as pesticides and polychlorinated biphenyls (PCBs). It also
confirms a significant breakthrough in analytical methods such as sample collection, workingup procedure, and determination of PAH with analytical instruments such as capillary GC, GCMS, and high performance liquid chromatography with ultraviolet and fluorescence detectors.
Unfortunately, research in developing countries, in particular Indonesia and other tropical
regions of the globe, has up to the present been sorely lacking. On the other hand, the relevance
of tropical environments for the global redistribution of organic pollutants has been increasingly
recognized by scientists. For instance, Iwata et al. (1994) showed that the discharge of persistent
and semi-volatile compounds taking place in eastern and southern Asia such as India, Thailand,
Vietnam, Malaysia, Indonesia and Oceania had a significant effect on pollutant redistribution on
a global scale, particularly important through air and water phases.
This study attempts to refine the scientific understanding of the distribution and
magnitude of PAH contamination in such aquatic environments, especially those surrounded by
large areas of peat. It attempts to comprehensively examine the concentrations of PAHs in the
sediments, suspended particulate matter (SPM), and water solution of Siak River, its estuary and
the coastal areas of the Riau province, Sumatra, Indonesia. To start off, the definition and
physico-chemical of selected PAHs are concisely introduced to give a general understanding on
the subject of the compounds investigated (Chapter I). Also, in this chapter, the relevance of the
PAH contamination in the aquatic environment is presented for some points of motivation i.e.
toxic/carcinogenic effects, bioconcentration/biomagnification as well as persistence and
degradation aspects of these compounds. In response to those motivations, the distribution and
the sources of PAHs in the aquatic environment are shortly reviewed to examine the key
environmental problems underlying this PAH investigation (Chapter II). Then, the PAH
determination is presented for three aquatic compartments: sediment, SPM and water as for
dissolved PAHs. Methods of analysis are given in Chapter III. The results and discussion are
elaborated in three manuscripts for publication. The first manuscript deals with the distribution
and source of PAHs in the sediment of the Siak river system (Chapter IV). This manuscript
examines the levels and sources of anthropogenic PAH contamination in two sediment fractions

(sand/coarse: 2 mm 63 m and mud/fine: <63 m on the Wentworth scale). The second


manuscript evaluates the concentration, spatial distribution and sources of anthropogenic PAH
contamination on SPM and dissolved PAHs in surface waters of the given study areas (Chapter
V). The last manuscript deals with the character differentiation of PAHs in surficial sediments
from peat- and non-peatland aquatic systems, which is a comparative study of the Siak estuary
of Sumatra and Wenchang/Wanquan estuaries of Hainan, China (Chapter VI).
The work undertaken in this study was an integrated part of Cluster 3.1. of the SPICE
Project (German-Indonesian Science for the Protection Indonesia Coastal Ecosystem) - a
research collaboration on marine biogeoscience carried out between 2004 and 2007. The general
theme of Cluster 3.1 was "coastal ecosystem health: the transfer of natural and anthropogenic
materials from land to the coastal sea", focusing in particular on the Siak River in Riau
Province, Sumatra, Indonesia. It is fully recognized that there has been no analogous study
undertaken for the particular areas mentioned above up to the present. This, of course, brings
with it the added difficulty of comparing the magnitude of contaminants in Indonesia with those
in other areas of the world. The nature of this study is important, because it presents initial
scientific data as part of the first comprehensive study on PAH distribution for this region. In
this respect, the results can serve as a starting point for discussion and reference for the future
analogous studies attempting to further refine the knowledge in this area.

1.1.
1.1.1.

Polycyclic aromatic hydrocarbons: definition and physio-chemical


characteristics
Definition
Polycyclic aromatic hydrocarbons (PAHs), also called polyaromatic hydrocarbons or

polynuclear aromatic hydrocarbons, refer to a group of organic arene compounds composed of


two or more fused aromatic benzene rings. These compounds contain various configurations of
molecular ring structures, however the base-unit rings do not have heteroatoms within the rings.
PAHs are therefore explicitly associated with their parent compounds. A fused ring results from
the sharing of two or three specific carbon atoms by two/three connected aromatic rings. The
carbon-sharing rings lead to the formation of a virtually single-planed structure composed
entirely of carbon and hydrogen atoms (Neff, 1979). This planar structure allows for large and
highly-diverse molecules, which can be constructed with different numbers and positions of the
aromatic rings.
There are possibly hundreds of PAH compounds occurring in an extremely complex
mixture in the environment. For the purposes of this study, however, we restricted ourselves to
the sixteen parent PAH compounds listed by the US Environmental Protection Agency on its
priority pollutant list (the 16 US EPA). These compounds are among those which have been
frequently used for the purposes of environmental quality assessments. The base structures of

the sixteen parent compounds are composed of 2-6 aromatic rings with molecular masses
ranging from 128 Dalton to 278 Dalton (Fig. 1.1). They include naphthalene (NAPH),
acenaphthylene (ACYN), acenaphthene (ACEN), fluorene (FLU), phenanthrene (PHEN),
anthracene (ANTH), fluoranthene (FLA), pyrene (PYR), benzo(a)anthracene (BaA), chrysene
(CHRY), benzo(b)fluoranthene (BbFLA), benzo(k)fluoranthene (BkFLA), benzo(a)pyrene
(BaP), dibenzo(a,h)anthracene (DANTH), benzo(g,h,i)perylene (BPERY) and indeno(1,2,3c,d)pyrene (IPYR).
1.1.2.

Selected physio-chemical characteristics


These unsubstituted aromatic compounds are non-polar, lipophilic and highly

hydrophobic in nature. Selected physio-chemical characteristics of given PAHs are summarized


in Appendix 1 (see for details and references). Briefly, the compounds are highly soluble in
certain organic solvents with logarithmic values of the octanol-water partition coefficient (Log
Kow) for each of the compounds ranging from 3.45 to 6.75 (Williamson et al., 2002). Their water
solubility is very low and ranges between values of only 0.3 g/L and 30.2 mg/L (Williamson et
al., 2002). Low molecular weight compounds i.e. naphthalene, acenaphthene and
acenaphthylene, have the highest water solubility with values of 30.2, 16.1, and 3.93 mg/L,
respectively. The solubility decreases with increasing molecular mass. PAHs generally tend to
be more easily adsorbed onto organic matter. In the environment, PAHs are readily associated
with other natural, organic substances such as biopolymers (e.g. polysaccharides, lipids, protein
and polynucleic acids), humic substances (e.g. humic acids and fulvic acids), kerogens and
black carbon. One important review of the interactions between contaminants and these geosorbents was given by Weber et al. (2001).
The vapor pressure of PAHs is quite low, ranging from 8.89 10-2 2.10 10-11 mmHg,
which often leads to their classification as semi-volatile compounds. Their boiling points ranges
from 218 oC to 542oC, and the melting point span from 80 oC to 279 oC. At ambient conditions,
they usually occur as almost colorless solids, normally being white or pale yellow in color e.g.
IPYR. In the aquatic environment, however, they occur either as free molecules or associated
with dissolved organic matter, particulate phases and sediments.
The benzene ring structures of these compounds are rigid. PAH molecules are
appreciably stable and prefer substitution reactions to additions. With respects to molecular
structure, there is a different degree of thermodynamic stability between peri-condensed PAH
compounds to cata-condensed ones (Grope, 2001). Peri-condensed PAHs such as fluoranthene,
pyrene, benzofluroanthene, benzo(a)pyrene, benzo(g,h,i)perylene, indeno(1,2,3-c,d)pyrene, are
less stable. On the other hand, within the cata-condensed structure, those with linear structure
are less stable than their angular isomers such as anthracene to its isomer phenanthrene (Grope,

2001). PAH compounds are characterized by broad ultraviolet absorbance spectra. In addition,
most PAHs are fluorescent and emit light when excited, with the exception of acenaphthylene.

NAPH
Naphthalene
C10 H 8
MW=128,17

ACYN
Acenaphthylene
C12 H 8
MW=152,20

FLA
Fluoranthene
C16 H 10
MW=202,26

BbFLA
Benzo(b)flouranthene
C20 H 12
MW=252,32

ACEN
Acenaphthene
C12 H 10
MW=154,21

PYR
Pyrene
C16 H 10
MW=202,26

BkFLA
Benzo(k)fluoranthene
C20 H 12
MW=252,32

BPERY
Benzo(g,h,i)perylene
C22 H 12
MW=276,34

FLU
Fluorene
C13 H 10
MW=166,22

PHEN
Phenanthrene
C14 H 10
MW=178,22

ANTH
Anthracene
C14 H 10
MW=178,23

BaA
Benzo(a)anthracene
C18 H 12
MW=228,29

CHRY
Chrysene
C18 H 12
MW=228,29

BaP
Benzo(a)pyrene
C20 H 12
MW=252,30

DANTH
Dibenzo(a,h)anthracene
C22 H 14
MW=278,35

IPYR
Indeno(1,2,3-c,d)pyrene
C22 H 12
MW=276,34

Fig. 1.1. The molecular structures and masses of the 16 parent PAHs as listed in the US EPA priority
pollutant list. The structures are listed by the number of increasing ring groups (from 2 to 6 rings).

1.2.

PAH contamination in the aquatic environments: background to


environmental problems
PAH contamination is one of the primary environmental problems facing humanity at

present. This is due to the fact that PAHs are toxic, mutagenic and/or carcinogenic to both
humans and other organisms. They also are subject to bioaccumulation and concentration in the
aquatic food web. Additionally, they are relatively persistent in the environment. Mounting
literature has provided sufficient evidence of their global distribution and high concentrations in
atmospheric, terrestrial and aquatic systems. For example, Zhang & Tao (2009) published the
global atmospheric emission inventory of 16 PAH priority pollutants from 37 countries. It was
estimated that PAH global emission in 2004 accounted for 520 giga grams per year, with 55%
of the total emissions coming from Asia. China, India and the United State were the top three
countries with the highest PAH emission. However, as environmental investigation increases
worldwide it can be assumed that the number of and places in which significant amounts of
PAHs are found may also increase. This implies an increasing potential of human exposure to
PAHs and a growing number of possible sources for these compounds. Particularly important in
this respect are aquatic systems, which include rivers, estuaries and coastal waters, upon which
tens or even hundreds of millions of people rely upon for food and other resources. Both water
for drinking and hygiene and also the vast number of both freshwater and marine fish species
and other organisms used as food sources for humans are directly tied to such ecosystems. This
is especially true for a country like Indonesia where people rely on river, estuary and coastal
resources.
1.2.1.

Toxicity of PAHs
PAHs pose toxic, mutagenic and/or carcinogenic threats to humans and other organisms.

A number of toxicity and cancer cases in both human and aquatic organisms have partly been
associated with increasing chronic and acute exposure to high concentrations of PAHs from the
ambient environment or specific contaminated sites (e.g. Brasseur et al., 2007; Cachot et al.,
2006; Chiang et al., 2009; Hu et al., 2007; Smith et al., 2000). Many molecular epidemiological
studies have shown evidence of increased levels of several biomarkers indicating PAH
genotoxicity, for example PAH-DNA adducts and oxidative DNA damage in populations
exposed to increasing levels of PAH (e.g. Shou et al., 1996; Hussain et al., 1998; Liu et al.,
2007; Singh et al., 2007). A relationship between PAHs and cancer causes of lung, skin, bladder
(Bofetta et al., 1997) and prostate (Rybicki et al., 2006) was shown to be conclusive.
The mutagenic/carcinogenic effects of PAHs are mainly exerted through electrophilic
metabolic activation of the compounds due to their planar, highly conjugated aromatic
structures. PAH metabolites are then capable of modifying DNA, which is the key to
carcinogenesis. Several mechanisms of metabolite activation have been widely proposed by

scientists, including the formation of dihydrodiol epoxide. This mechanism turns out to be the
most frequent pathway, which is often called the "bay region dihydrodiol epoxides pathway"
(Xue & Warshawsky, 2005). In this fashion, PAHs are oxidized by P450 enzymes in the initial
step in the activation process. This then produces reactive electrophilic metabolites which are
capable of interacting with cellular macromolecules, in particular with nucleic acids and
proteins. A review of the mechanisms of metabolic activation of PAHs exerting carcinogenicity
was provided by Xue & Warshawsky (2005).
Although it is still an area of intensive epidemiological study, toxicological profiles of
PAHs have already been introduced (e.g. ATSDR, 1995 see www.atsdr.crc.gov). The
International Agency for Research on Cancer (IARC), an institute within the World Health
Organization, has classified the carcinogenicity of individual PAH compounds in several of the
parent groups (Table 1.1.). This classification was based on the strength of evidence derived
from studies in humans and experimental animals (http://monographs.iarc.fr). However, even
those individual substances which could not be positively identified as carcinogens may act
synergistically (Wenzl et al., 2006). High molecular mass (HMW) PAHs tend to be more
carcinogenic, but less acutely toxic than their cousins with lower molecular masses (LMW).
Benzo(a)pyrene (BaP) in particular has revealed itself to be the most prominent carcinogen
(Group 1 of IARC) and is often used as a hazard index for PAH exposure (Bofetta et al., 1997;
Ravindra et al., 2008, Rappaport et al., 2004). However, even the LMW compounds like
naphthalene and acenaphthene have been shown to exert carcinogenic effects on animals and
may also carry potential risk for humans (Long et al., 1995; Rappaport et al., 2004).
As a consequence of increasing concentrations of PAHs in the environment, the
likelihood of human exposure could raise. Human exposure can occur through inhalation,
absorption/adsorption (skin), and ingestion (Bofetta et al., 1997; ATSDR, 1995). Inhalation and
skin contact have been proven to be important pathways for atmospheric PAH to enter the
human organism. Both of these routes of entry into the body are strongly related to specific
occupations, such as aluminum and coke production, coal gasification, iron and steel foundry
work, tar distillation, petroleum cracking, shale oil extraction, wood impregnation, roofing, road
paving and carbon production (Bofetta et al., 1997). However, ingestion through food and
drinking remains the most significant route for PAH contamination in humans. As much as 90%
of the total daily intake of persistent pollution compounds into the human organism could result
solely from diet (Binelli & Provini, 2004; Wenzl et al., 2006). In this context, animal food
sources play a significant role, particularly fish and seafood.

Table 1.1. List of priority PAHs and their carcinogenicity classification, water and sediment
quality guidelines (content in g/L for water and ng/g dry sediment weight).
Compounds
Acenaphthene b,c,d
Acenaphthylene b,c,d
Anthracene a,b,c,d
Benz[a]anthracene b,c,d
Benzo[a]pyrene a,b,c,d
Benzo[b]fluoranthene a,b,c,d
Benzo[e]pyrene d
Benzo[g,h,i]perylene a,b,c,d
Benzo[j]fluoranthene d
Benzo[k]fluoranthene a,b,c,d
Chrysene b,c,d
Dibenz[a,h]anthrancene b,c,d
Fluoranthene a,b,c,d
Fluorene b,c,d
Indeno[1,2,3-c,d]pyrene a,b,c,d
Naphthalene a,b,c
Phenanthrene b,c,d
Pyrene b,c,d

Carcinogenicity
Classification*
(group)
3
NI
3
2B
1
2B
3
3
2B
2B
2B
2A
3
3
2B
2B
3
3

Water Quality
Standard**

Sediment Quality Guidelines***

0.006**
0.001e
0.001e

ERL
16
44
85.3
261
430

ERM
500
640
1100
1600
1600

384
63.4
600
19

2800
260
5100
540

160
240
665

2100
1500
2600

0.001e
0.001e

0.01**
0.001e
0.01**

EU WFD, 2000/60/EC in Annex X (http://europa.eu/scadplus/leg/en/lvb/l28108.htm);


the US EPA (http://www.epa.gov/epaoswer/hazwaste/minimize/chemlist.htm);
NPI Australia;
d
Canadian NPRI Substance 2007;
e
Priority & hazardous substances based on the decision # 2455/2001/CE of the European Parliament
* IARC: http://monographs.iarc.fr/ENG/Classification/index.php. Group 1: Carcinogenic to humans; Group 2A: probably
carcinogenic to humans; Group 2B: possibly carcinogenic to humans; Group 3: not classifiable as to carcinogenicity to humans;
Group 4: probably not carcinogenic to humans;
** Maggi et al 2008;
*** adopted from Long et al., 1995: ERL= the effects range-low, ERM= the effects range-median
NI = information not available
b
c

1.2.2.

Bioconcentration and magnification in aquatic food webs


PAHs are bioconcentrated and biomagnified by marine organisms. Bioconcentration and

bioaccumulation of PAHs in organisms occurs via all three chemical exposure routes, including
dietary absorption, transport across respiratory surfaces and dermal absorption (Mackay &
Fraser, 2000). Of these possibilities, the first two routes are the most important. The degree of
bioconcentration is determined by tissue lipid content (Kayal & Connell, 1995). An appreciable
increase in PAH concentration has been observed in various marine organisms as compared to
their environment (water or sediment). For instance, Barbour et al. (2008) observed appreciable
bioaccumulation factors in oysters harvested from a war-induced oil spill zone in the Eastern
Mediterranean Sea, which extend from 242 to 3700.
Along with bioconcentration, biomagnification has been recorded from lower to higher
trophic levels of organisms, e.g. plankton (Carls et al., 2006), benthic amphipods (Vigan et al.,
2007); mussels (Okay et al 2000; Richardson et al., 2003; Prez-Cadaha et al 2004; Hellou et
al., 2005), other bivalve organisms (Oros & Ross 2005), oysters (Mondon et al., 2001), eels
(Ribeiro et al 2005), feral finfish (Hellou et al 2006), and other fish species (Liang et al., 2007).

It has been suggested that the higher the trophic level, the more and various food types are
consumed. As a result, higher trophic levels are more susceptible to pollutant magnification.
However, in aquatic systems benthic organisms such as mussels and clams accumulate more
PAHs than the carnivorous fish which use these organisms as prey. Mart-Cid et al. (2007)
found that shellfish (mussels & clams) and shrimp can contain higher concentrations of PAHs
than fish such as tuna, mackerel and salmon, due to a low capacity to biotransform
contaminants. Physiologically, mollusks do not metabolize PAHs as quickly as fishes. Instead,
they tend to accumulate these toxins.
Bioconcentration/biomagnification is determined by bioavailability of the compounds in
the water phase. Low molecular mass PAHs (those with 2-3 rings) indicated high levels of
bioavailability in benthic organisms such as mussels and clams when compared to high
molecular mass compounds (e.g. Baumard et al., 1999a; Baumard et al., 1999b; Kayal &
Connell, 1995). This is due to the fact that those compounds are highly water-soluble. On the
other hand, HMW PAHs (4-6 rings) have been proven to be relatively non-bioavailable when
compared to the lighter compounds. Baumard et al. (1998) found low concentrations of heavier
PAHs in mussel, despite bottom sediments containing high levels of various pyrogenic-PAHs.
Thorsen et al. (2004) observed that petrogenic PAHs are more bioavailable than pyrogenic
PAHs.
However, the bioavailability of the compounds is also determined by water and sediment
properties i.e. organic matter content, SPM and sediment grain-size. Organic matter found in
both water and sediment affects the bioavailability of PAHs to a large extent. One notable
review of the effects of dissolved organic matter (DOM) on the bioconcentration of several
organic contaminants (PAHs, chlorinated hydrocarbons, and TBT) in aquatic organisms,
including water fleas, mussels, amphipods and fish, was given by Haitzer et al. (1998). The
authors reviewed the lack of bioavailability of DOM-bound chemicals, which would in turn lead
to a decrease in their bioaccumulation. However, at low level of DOM (~10 mg/L),
bioconcentration of pollutant could be enhanced up to 300% (Haitzer et al., 1998). Moreover,
the extent to which DOM affects PAHs is further determined by factors including the quality
and quantity of DOM, the contact time between humic substances and the chemicals before
exposure, and the overall exposure time. Data might therefore differ from place to place
according to environment-specific characteristics.
Further factors such as the presence of suspended particulate matter and the fineness of
sediment grain-sizes are also important for the bioaccumulation of organic pollutants, especially
in filter/suspension-feeding and detritivorous organisms such as bivalves. Baumard et al. (1998)
found a significant bioconcentration of PAHs on organisms living in close contact with
sediments compared to carnivorous organisms. The latter might be exposed to a much lesser
extent to sedimentary particles. Menon & Menon (1999) found in a 10-day experiment that the

bioaccumulation of PAHs in clams exposed to sediment in suspension increased almost twofold when compared to clams tested in undisturbed sediments. This indicates that higher
sediment levels in aqueous suspensions can result in an increased PAH content in the
surrounding water. Such conditions are important in dynamic systems such as those found in
rivers, estuaries and coastal waters. High turbidity should be expected to increase the
concentration of carcinogenic, high molecular weight PAHs in benthic organisms (e.g. Baumard
et al., 1999b). Such benthic invertebrates are important prey (food) for many carnivorous fishes.
Therefore, increased bioaccumulation can also be extrapolated for higher trophic levels of
organisms found in the food webs of these aquatic systems.
1.2.3.

Persistence, low degradation rates and pollution indicators


PAHs are persistent over longer periods of time when found in bottom sediments. Once

bound to aquatic sediment particles, PAHs can effectively survive for years. This fact stems
from their relatively stable chemical structures, particularly under anaerobic conditions (Neff,
1979; Mihelcic & Luthy, 1988). Due to this, PAH signatures in sediments are often used as
indicators in identification of pollutant sources. PAHs have also been tested by geochronicle
studies looking at dated sediment cores (e.g. Gevao et al., 1998; Yunker et al., 1999; Yamashita
et al., 2000; Fabbri et al., 2003; Ricking et al., 2005). Thirty years ago, Hites et al (1977)
published an experiment which showed the relative stability of PAH composition over hundreds
of years. Their analysis documented unsubstituted PAHs and their alkyl homologues in three
sections of core sediments taken from Buzzards Bay, Massachusetts, USA. The researchers
found that the distribution of PAHs remained qualitatively constant, even though the intensity of
the pollution source had increased considerably.
Degradation of PAHs might, however, potentially affect the relative levels and makeup of
persistent PAHs in aquatic environments. Such degradation might occur by photo-/chemical
oxidation (e.g. Behymer & Hites, 1985; Lehto et al., 2000; Shemer & Linden, 2007) or
biodegradation (Poeton et al., 1999; Kanaly et al., 2000; Kot-Wasik et al., 2004). Miller and
Olejnik (2001) studied the photolysis of water-borne PAHs (BaP, Chry, Flu) via UV radiation.
The study revealed that the photo-degradation of PAHs in water involves rather complicated
mechanisms involving oxygen, the function of pH, and scavengers (organic materials).
Degradation of BaP and Chry (high molecular weight PAHs) was found to be retarded in water
with a lack of oxygen, high pH values (alkalinity) and an increase in the level of organic
scavengers. Conversely, FLU (low molecular weight) proved itself to be independent of those
parameters, except for the fact that FLU was eliminated more quickly when oxygen
concentrations were lower. Microbial degradation is often enhanced when chemicals are found
in the dissolved phase (high bioavailability). Low molecular weight PAHs (2 3 rings) are
more susceptible to microbial degradation due to bioavailability than high molecular mass

compounds (>4 rings). The latter are more recalcitrant when it comes to undergoing chemical
changes (Juhaz & Naidu, 2000). However, a great deal of research has confirmed the microbial
degradation in sediment-bound PAHs. Xia et al. (2006) concluded that an increase in the
population of PAH-degrading bacteria and desorption of PAHs from the solid phase increases
the rate of contact between PAH and bacteria. This can subsequently enhance biodegradation
rates. Some of PAH-degrading bacteria are mentioned in the work of Samanta et al. (2002).
However, it is important to note that biodegradation is generally a relatively slow process
(Mnnist et al., 1996).
In addition to being used as indicators for anthropogenic pollutants, the presence and
amount of PAHs often serve as measures for estimating overall environmental quality. This
includes factors such as the degree of toxicity (e.g. Bihari et al., 2006; Cachot et al., 2006),
environmental and human health risk assessment (e.g. Galloway 2006) and petroleum pollution
(e.g. Requejo et al., 1996).
In conclusion, PAHs are regulated not only for food and drink, but also for environmental
aspects, due to their toxicity and carcinogenic properties, confirmed bioaccumulation and
biomagnification and their persistence over longer periods of time. Examples of legal directives
regulating PAHs are: the European Union Water Framework Directory (the EU WFD,
2000/60/EC in Annex X), the US EPA list of priority pollutants, the National Pollutant
Inventory (NPI) for Australia, and the European Scientific Committee on Food (2002) (see
Stolyhwo & Sikorski, 2005). Despite these efforts, PAHs have still not been specifically
mentioned in the list of twelve POPs (Persistent Organic Pollutants) under the Stockholm
Convention, which has so far been signed and ratified by more than 150 countries including
Indonesia (http://chm.pops.int/).

1.3.
1.3.1.

The relevance of rivers, estuaries and coastal areas in Indonesia


A Perspective on the global pollution dispersal
Rivers and their tributaries, estuaries and coastal areas represent a transitional boundary

between the terrestrial and marine aquatic system. These bodies of water act as the "front line"
with respect to receiving the majority of land-based loading materials, including organic
pollutants. Therefore, their role in the distribution and transport of chemicals into the ocean
reservoir is crucial. Chester (2003) pointed out that river runoff is one of the primary, globalscale sources for material to enter the oceans when taken together with atmospheric deposition
and hydrothermal activity. With regard to this, rivers play a large and important role as the main
carriers of numerous chemical signatures into the ocean.
A variety of surface runoff types exists. The flows from diverse landscapes, e.g.
municipalities, various sorts of heavy, medium and light industry, and agriculture, groundwater
seeps into riverine systems, and into the ocean. Atmospheric deposition also enriches the

10

magnitude of chemicals found in aquatic systems. These sources become especially important
for aquatic systems in regions where economic development has been significantly and rapidly
taking place, areas of the world such as Southeast Asia. Accordingly, the presence of pollution
in the rivers, estuaries and coastal waters of these regions represents one of the key
environmental challenges facing humanity today. Schwarzenbach et al. (2006) estimated that
anthropogenic fluxes of organic pollutants stemming from fertilizers, pesticides, synthetic
organic chemical productions and accidental oil spills annually contribute ca. 450 million tons
worldwide to aquatic systems.
Scientists have established the global significance of the rivers draining southern Asia,
which flush large amounts of terrestrial materials into the world ocean. On a global scale, recent
estimation of the annual total sediment load flowing from rivers into the global ocean is ca. 20
Gigatons, suspended sediment load contributes 90%, and the rest is mainly from bed sediment
load (Syvitski et al., 2003).
Asia and Oceania are the largest producers of fluvial sediment, with around 70% of the
overall annual sediment loads coming from these regions (Milliman et al., 1999). In comparison
with the globe's largest landmasses, including Africa, Asia, Europe, North America, and South
America, the region containing Oceania and Indonesia produces the world's highest sediment
yield, as defined in terms of sediment load divided by total drainage area (Syvitski et al., 2005).
In fact, these two regions accounted for ca. 800 and 543 tons of sediment per square kilometer
per annum, respectively. Milliman et al. (1999) pointed out the particular importance of the
rivers in Sumatra, Java, Borneo, Sulawesi, Timor and the New Guinean islands (Fig. 1.2). They
estimated that these six large islands alone significantly discharge about 4.2 Gt of sediment per
year, despite the fact that they constitute only 2% of the world's total land area which drains into
the global ocean. The rivers located on these islands are responsible for about 20-25 % of global
sediment export. Sumatra alone contributes 498 Megatons through fluvial transport per annum,
an amount which makes it the second largest source in this group of six islands (Milliman et al.
1999). Even though some of the larger river basins in Sumatra such as the Siak, Kampar,
Rokan, Indragiri and Batanghari might not have been included in the above-mentioned
calculations, further estimates suggest that the sediment volumes are substantial.
Most of rivers and estuaries in the eastern coast of Sumatra drain large area of peatland
which are characterized by high humic substance of black water masses. Therefore, the
environmental state of the tropical peatlands has attracted a big concern for pollutant transport
and climate change. Page et al. (2002) reveal that the stability of tropical peatlands is
particularly relevance for the climate change because these peatlands are one of the largest nearsurface reserves of terrestrial organic carbon estimated for ca. 26 50 Gt. They also remain that
persistence environmental change in the tropical peatlands owing to drainage and forest clearing
have threaten the stability of ca. 20-m thick of peat deposit, and make the peatlands being

11

susceptible to fires. In Indonesia, peatland fires have long been an environmental problem (see
section 1.3.2. for further explanation). The worst peatland fire episode occurred during severe El
Nio event (1997/1998) damaging of ca. 6.8 Mha (or 34%) of Indonesia peatland, and emitted
up to 2.57 Gt Carbon in 1997 (Page et al., 2002). It is relatively similar to the global net
emission of CO2 from land-use change was recently estimated for 2.4 Gt y-1 (IPCC, 2000). The
environmental and health effects of the haze and smoke had been widely recognized ever since
(e.g. Langmann et al., 2007; Fang et al., 1999; Parameswaran et al. 2004). Of particularly
important is that the peatland fires are mostly anthropogenic as part of land clearance activities
before establishing crops (Page et al., 2002).
Unfortunately, the magnitude of organic pollutants, particularly that of PAHs the
burning by products - found in Indonesia's rivers, estuaries and coastal areas has only been
superficially evaluated at best. To date there have been no long-term or in-depth studies carried
out, especially ones covering extremely large surface areas or multiple islands in this region.
Nevertheless, the significance of organic pollution fluxes stemming from Indonesian rivers can
be assumed to be directly proportional to the significance of material fluxes previously
mentioned by the above literature sources. It is therefore reasonable to assume that increasing
levels of "flushing out" SPM from these river systems will eventually contribute to an increase
in the magnitude of pollutants in the coastal zone, given that suspended particulate matter
(SPM) is a significant carrier for most organic pollutant such as PAHs (e.g. Kayal & Connell,
1989; Ollivon et al, 1995; Deng et al., 2006; Law et al., 1997; Fernandes et al., 1999; Heemken
et al., 2000; Witt & Siegel 2000; Kowalewska et al., 2003; Ross & Oros, 2004 and Cao et al.,
2005), polychlorinated biphenyls (PCBs) (e.g. Mai et al., 2002; Telli-Karako et al., 2002),
polychlorinated dibenzo-p-dioxins (PCDDs) and polychlorinated dibenzofurans (PCDFs) (e.g.
Ishaq et al., 2003). The coastal zone of Sumatra can thus be viewed as being very susceptible to
organic contaminants. Conversely, dissolved organic matter such as dissolved humic substances
has been shown to enhance the solubility of dissolved PAHs in aqueous solution (e.g. Liu &
Amy, 1993). The implication of this additional factor is that the residence times of PAHs in the
aqueous phase will be increased, which in turn promotes more far-reaching transport towards
and into the open ocean. The relevance of Sumatran rivers, which often drain peat bogs and
other humic soils, for the transport of dissolved organic matter into the open ocean has already
been recognized (e.g. Baum et al., 2007).

12

Das Bild k ann nicht angezeigt werden. Dieser Computer v erfgt mglicherweise ber zu wenig A rbeitsspeicher, um das Bild zu ffnen, oder das Bild ist beschdigt. Starten Sie den Computer neu, und ffnen Sie dann erneut die Datei. Wenn weiterhin das rote x angezeigt wird, mssen Sie das Bild mglicherweise lschen und dann erneut einfgen.

Fig. 1.2. Sediment discharge (106 T y-1) from the six East Indies islands. Arrow widths are proportional to
the annual load. The letters S, J, B, C, T and NG refer to Sumatra, Java, Borneo, Sulawesi (Celebes),
Timor and New Guinea, respectively. Shaded areas represent water depths of less than 1000 m, although
most of these areas are in less than 100 m of water depth (adopted from Milliman et al., 1999).

1.3.2.

Environmental settings of the study areas: An Overview


Our chosen study areas include the Siak River, its estuary, and the coastal areas

Indonesia's Riau Province, extending from the Northwest to the Southwest coast in the Selat
Panjang (the coastal channel running between offshore islands and the Sumatra mainland) (Fig.
1.3.). Details of the exact sampling locations are provided in each results chapter. The areas
selected for this study provide an interesting opportunity to comprehensively study the spatial
and phase distributions of PAHs in different river systems, all of which ultimately empty their
waters into the Malacca Strait. The sampled areas are located in the equatorial wet climate of
Sumatra Island, which is affected by a monsoon season mainly bringing rainfall between the
months of September and March. However, rainfall can also occur at almost any time
throughout the year. The average annual rainfall in 2004, for example, was 2088 meters with the
heaviest rainfall occurring between October and December (BPDAS, 2004, unpublished
seminar notes). However, even such small changes in regional weather patterns can affect the
flow of terrestrial pollutants and organic materials into the marine environment through various
pathways such as varying land erosion, storm water discharge, and surface runoff volumes.
The Siak River system and its catchment area are composed of ca. 300 km of waterways,
including several tributary rivers (the Tapung Kanan, Tapung Kiri and Mandau) and various
sub-basin streams. Including the Siak's estuary system, the entire catchment area totals roughly
1 million hectares, amounting to ca. 10% of the total surface area of Riau Province (BPDAS,
2004). The average river flow rate is normally 200 m3 s-1 (a total of 6.307 x 109 m3 y-1), reaching
a peak flow rate of 1700 m3 s-1 during the monsoon season and a minimum of 45 m3 s-1 in the
course of droughts (unpublished data, PEMPROV RIAU, 2005). The Siak and its estuary

13

mainly drain large areas of low-laying land characterized by widespread peat swamps, which
are scattered among various landscapes, including huge palm-oil plantations, forests and
secondary swamp-forests. These areas account for ~53%, ~23%, and ~11% of catchment landuse coverage, respectively (BPDAS, 2004). The main river correspondingly receives large
amounts of water input stemming from smaller channels and tributaries along its course.
Moreover, the main river basin is sparsely inhabited. It nevertheless drains some highlyurbanized centers, such as the Riau Province capital of Pekanbaru, the smaller city of Siak Sri
Indrapura, a pulp-paper industry estate located in Perawang, and an oil refinery in the Siak river
mouth, thus making it inevitable that urban and industrial discharges will enter the river's water.
The Siak River and its estuary dump their waters in the coastal channel located between
Bengkalis Island and the Sumatra mainland.
Adjacent to the Malacca Strait, the coastal areas are constricted and range from Dumai
City in the northwest of the River down to the Selat Panjang channel in the southwest. The
coastal waters are characterized by observable plumes of suspended particulate materials. These
plumes stem mainly from the Siak River and some other neighboring rivers (such as the Rokan
and Kampar River in the northwest and southwest, respectively) and many smaller streams
emerging along the coast (Fig. 1.5.). The Siak River plume is caused by dark-colored peat
materials and stands out distinctively from the plumes of other rivers. The Siak's current pushes
river water through the channel and into the Malacca strait during the low tide. Its plume pulls
back during high tide. The tide is semi-diurnal, meaning that two high tides and one ebb tide
occur during one day, however, the high tides arrive with differing magnitudes. Unfortunately,
there is no information available on the residence times of the coastal water masses.

14

MALAYSIA

Riau Coast
Bengkalis Island

Siak Estuary
Siak River

Riau Province,
Sumatra, INDONESIA
Fig. 1.3. An overview map of the Siak river mouth (satellite image provided by H. Siegel (2004), Baltic
Sea Research Institute (IOW) Warnemuende, Rostock, Germany). The Siak River stretches over ca. 300
km up to the estuary (the mouth). The coastal areas extend to far northwest of the mouth and to southwest
of the coastal channel between the offshore islands and the Sumatra mainland.

Fig. 1.4. (a) Surface runoff of organic-rich black water from a small waterway entering the Siak River
around a palm-oil plantation, (b) typical small water channel found in the estuary. These are typical water
inputs feeding the Siak River (source: courtesy of SPICE).

15

Rokan River

Siak Estuary

Siak River

Sumatra,
INDONESIA

Kampar River

Fig. 1.5. Plumes of SPM and dissolved humic material from several rivers: the Rokan, Siak and Kampar
River (as seen from northwest to southwest) into the Malacca strait. At low tide the plume is flushed
outwards by the Siak River and at high tide it is pushed backwards. Image from H. Siegel (2004), Baltic
Sea Research Institute (IOW) Warnemuende, Rostock, Germany using MODIS Terra, Data: NASA-RRS.

With regards to PAH pollution, there are two potential sources which have their roots in
the expanding economic development of the Province. Riau is well-known for its oil production
industry. Oil-related facilities such as various plants, refineries and ridges are scattered all along
the river, the estuary and coastal areas (Fig. 1.5). Oil spills might appear along the Siak River,
the estuary and the coast, and most likely stem from boats, oil-transporting vessels, harbor
facilities, and many other petroleum-related activities.

Fig. 1.6. Some oil-related activities observed in the studied areas. Refineries are mainly located in the
coast of Dumai city (left), ridges (middle) in the channel (Panjang Strait), and typical port activities in the
Siak River (source: courtesy of SPICE).

16

Moreover, oil industries can also be found in surrounding areas of two neighboring
countries, Malaysia and Singapore (the biggest oil refinery center in Asia). Malacca Strait is
likely destined to become one of the busiest straits in the world, one which is highly susceptible
to oil pollution. Zakaria et al. (2000) found that oil pollution levels detected in the Strait were
due mainly to spillage (through accidents or routine tanker operations such as ballast-water
discharges) from the tens of thousands of vessels transporting crude oil and other petroleum
products.
The region's second-largest industry, plantations for palm-oil production, covers more
than 50% of the total catchment area. The studied areas were affected by high levels of organic
material combustion, in particular anthropogenic, slash-and-burn agriculture and intense
instances of naturally-occurring forest and swamp fires. Especially important for this region is
the documented fact that uncontrolled agricultural burning for purposes such as land clearing
has worsened overall air quality for over a decade.
Fig. 1.7 shows the trend of hotspots from 1996-2006 and Fig. 1.8 shows an example of a
daily hotspot distribution. Data was extracted from the hotspot daily observation provided by
the Indonesian Ministry of Forestry (www.dephut.go.id). Elevated numbers of hotspots were
observed during the severe El Nio event of 1997-1998. It is not to extrapolate that the numbers
of hotspots correlated directly with El Nio, because high numbers have also occurred during
2005 and 2006, which did not fall under the pronounced effects of El Nio. Severe burning
happening during 1997/1998 resulted in a strong impact on the distribution of aerosols over
Sumatra and the neighboring counties (e.g. Fang et al., 1999) and the effects even reached into
the Indian Ocean. Parameswaran et al. (2004) observed a large aerosol plume which formed
over the eastern equatorial (5oN to 10oS) and eventually reached 60o E latitude during
September-November of 1997. This plume was directly tied with the large-scale fires occurring
in Southeast Asia, particularly Indonesia, at the time of a severe drought caused by an El Nio
event. Unfortunately, the observed number of annual hotspots has remained at a level of >3000
ever since then and has recently become even worse. The occurrence of agricultural burning is
most likely during the dry season, the period between March and September. Nonetheless, (as
can be seen in the Fig. 1.7b), large numbers of hotspot have been observed even in the assumed
rainy season. This is to say that forest and swamp fires, added to uncontrolled agricultural
burning, are potentially the largest source of PAH contamination in the studied areas.

17

20000

18959

18645
17090

18000

Numbers of hotspot observed

16000
14000
13711

12000
9360

10000
7587

8000

7323

7818

6000
6077

4000
2000

4865

El Nio

SPICE
Sampling

3780

1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006

Dry Season

number of burning hotspot

6000

Wet Season
>9000

5000
4000

1997
2002

3000

2004
2005

2000

2006
1000
0
May Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr

Fig. 1.7. (a) The number of hotspots (forest/swamp/agricultural fires) observed over the period from
1996-2006 in Riau Province. (b) Monthly distributions of hotspots for selected years. Unexpected raised
number of hotspots during wet season was observed in 2002 and 2005. (Data were extracted from
PUSDALKARHUTLA Riau Province for 1996-2001; and DEPHUT online publications for 2003-2006,
which were calculated from daily available information: www.dephut.go.id).

18

Das Bild k ann nicht angezeigt werden. Dieser Computer v erfgt mglicherweise ber zu wenig A rbeitsspeicher, um das Bild zu ffnen, oder das Bild ist beschdigt. Starten Sie den Computer neu, und ffnen Sie dann erneut die Datei. Wenn weiterhin das rote x angezeigt wird, mssen Sie das Bild mglicherweise lschen und dann erneut einfgen.

Fig. 1.8. An example of daily observation of hotspot distribution, provided by the ministry of Forestry
Indonesia (image source: www.dephut.go.id ). This picture shows intensive burning (red dots) around the
catchment areas of Siak River, estuary and some other rivers on June 19th, 2004.

1.4.

Study Objectives

The objectives of this study were to:


1)

Determine the concentration and distribution of PAHs in three aquatic compartments:


surface sediment fractions, suspended particulate matter (SPM), and water, taken from
three distinctly different aquatic systems: riverine, estuarine and coastal areas,

2)

Compare the level of PAH contamination, both within the three different aquatic systems
and also to similar systems in other geographical areas,

3)

Apportion possible contaminant sources through molecular weight and isomer ratios,

4)

Study the fate of PAHs in water phases by examining the partition equilibrium to SPM
from the bulk water solution (here considered as dissolved phase),

5)

Analyze the role of organic carbon content as a determinant factor affecting PAH fate in
the various aquatic environments and water chemistry (particularly salinity and pH).

19

II.

SOURCES, DISTRIBUTION AND FATE OF PAHs IN AQUATIC


COMPARTMENTS: A SHORT REVIEW

2.1.

Introduction
This chapter briefly reviews and elaborates upon various aspects of PAH sources,

distributions and fates in an aquatic environment. The sources are grouped around the PAH
formation process, magnitude, and composition with respect to the unsubstituted PAH
compounds and their alkyl homologues. This review focusses on PAH genesis which is related
to anthropogenic processes (mainly combustion), rather than discussing natural, early diagenetic
processes. This is because the former is the primary subject of this study. A comparison of
various anthropogenic sources, including their composition, is also provided for the 16 PAHs
defined by the US EPA priority pollutant list (hereafter called PAH16). Distribution of PAHs in
the aquatic environment is discussed based on their existence in sediment, suspended particulate
matter and water solution, whereas the fate of PAHs is elaborated upon based on their solidwater phase partitioning.

2.2.

Source and Signatures


PAHs in aquatic environments stem from a wide range of natural and anthropogenic

sources. The latter is of serious concern to humans, due to the fact that PAHs are by-products of
literally hundreds of different combustion processes from human civilization. These range from
residential heating systems to industrial processes like power generation, incineration,
manufacturing, agricultural burning and petroleum combustion for transportation, among many
others (e.g. Kakareka & Kukharchyk, 2003; Conde et al., 2005; Westerholm et al., 1992;
Zielinska et al., 2004; Yang et al., 2007). Recent global estimates of the atmospheric emission
of PAH16 substances in 2004 (520 Gg per year) showed that biofuels (56.7%), wildfires (17.0%)
and consumer product usage (6.9%) were the three most important sources (Zhang & Tao,
2009). In addition, such PAHs can be delivered into the environment from direct petroleum
discharges and spillages (e.g. Requejo et al., 1996).
PAHs stem from organic matter and are primarily constructed through three major
transformation processes: 1) rapid, early diagenesis ("natural" PAHs), 2) incomplete, hightemperature combustion ("pyrogenic" PAHs), and (3) catagenetic geological alterations which
result in petroleum ("petrogenic" PAHs).
2.2.1.

Natural PAHs
Use of the term natural PAHs differs somewhat in scientific literature, since this term

could potentially include both diagenic and catagenic processes. However, naturally-occurring

20

PAHs mainly refer to those substances resulting from the early diagenetic processing of
biogenic precursors (also called "diagenic" PAHs). The term "early diagenetic processes" refers
to any chemical, physical or biological post-depositional transformation which changes a
biogenic precursor over a relatively short-geological period (Berner, 1980; Libes, 1992). These
PAHs are limited to few groups of compounds which are derived mainly from terpenes, steroids
and pigments via aromatization processes. Wakeham et al. (1980a) observed five common
groupings of natural PAHs that occur in sediments: 1) a series of tetra- and pentacyclic PAHs
derived from pentacyclic triterpenes of the amyrin type, 2) tetra- and pentacyclic PAHs
stemming from pentacyclic triterpenes with five-membered E-rings (see Fig. 2.1), 3) retene and
pimanthrene with diterpenes as their parent materials, 4) an extended series of phenanthrene
homologues, and 5) perylene. These PAHs are commonly used for identification of natural
PAHs in the sediment (e.g. Lipiatou & Saliot, 1991; Silliman et al., 1998). Fig. 2.1 shows an
example of the transformation of triterpenoids precursors such as lupeol (five-membered Erings) and -amyrin (six-membered E-rings) into pentacyclic hydrocarbons by A-ring cleavage
and aromatization proposed by LaFlamme & Hites (1979). The natural PAHs occur mostly in
the form of alkyl homologues, and can also be intermediates in diagenetic transformations. For
example,

simonellite

(1,1,-dimethyl-7-isopropyl-1,2,3,4-tetrahydrophenanthrene)

is

an

intermediate compound which results from the diagenetic transformation of higher plant
diterpene abiatic acid into retene (Wakeham et al., 1980b).
These PAHs are primarily enriched in deeper, anoxic sediments, indicating a distinct
vertical profile when compared to anthropogenic PAHs. Siliman et al. (1998) distinguished
between natural PAHs, i.e. perylene, and other anthropogenics found in Lake Ontario in three
ways. First, the concentration of perylene in the surface sediment was significantly lower than
those for other anthropogenic PAHs. Second, the perylene concentration peak is found at an age
much older than the earliest industrial times. Third, concentrations of anthropogenic PAHs were
close to zero in sediment layers laid down before 1900. This kind of profile has also been
recognized in many studies performed in widely differing locations (e.g. Wakeham et al.,
1980a; Jiang et al., 2000).

21

Fig. 2.1. A hypothetical aromatization scheme for natural precursors: (a) -amyrin, (b) lupeol (adopted
from LaFlamme & Hites., 1979).

2.2.2.

Pyrogenic PAHs
Unlike natural PAHs, pyrogenic PAHs are produced in an extremely short period of time

through high temperature reactions, particularly the incomplete combustion of organic matter
(in the broadest terms: biomass and fossil fuels). The products are unsubstituted compounds
ranging from low molecular weight (100 200 Dalton, mostly 2-3 ring groups) to high
molecular weight (>200 Dalton, or mostly 4-ring groups) compounds, e.g. those composing
the 16-PAH list.
The formation mechanisms of PAHs have been the subject of intensive research over the
years (e.g. Neff, 1979; Glarborg, 2007; Appel et al., 2000; Rockne et al., 2000; Frenklach, 2002;
Ladesma et al., 2000; Dobbins et al., 1998). Two mechanisms have generally been
acknowledged as the best explanation of how PAHs are thermally generated: pyrolysis and
pyrosynthesis. Pyrolysis involves the cracking of complex and high molecular mass organic
molecules into lower molecular weight free radicals. This is immediately followed by
pyrosynthesis in which the newly-created free radicals are reassembled. Benzene and further
non-alkylated PAHs are produced by joining simple, individual benzene rings into double, triple
and larger, multi-ringed, high molecular mass ring structures (see Fig. 2.2). Most recently,
Richter and Howard (2000) reviewed and discussed the potential reaction pathways of PAH
formation which follow pyrolysis. Several kinetically-governed processes have been added to
the list, namely 1) oxidation, resulting in the formation of the first aromatic ring (benzene) and

22

heavy molecular weight PAHs (500 1000 Da),

2) nucleation or inception of nascent soot

particles (ca. 2000 Da, effective  1.5 nm), 3) particle mass growth due to the addition of gas
phase molecules including PAH radicals and 4) coagulation involving reactive particle-particle
collisions.

Fig. 2.2. Proposed mechanism of pyrosynthesis starting with ethane (adopted from Ravindra et al., 2008).

The production of PAHs is closely related to soot formation. Bockholn (1994) previously
proposed a schematic concept for the production of PAHs in which they were actually
precursors for soot and black carbon formation (Fig. 2.3). In principle, Bockholn's scheme
includes the aforementioned pyrolysis and pyrosynthesis processes of an organic fuel. This
results in small hydrocarbon radicals being created, from which acetylene (C2H2) is formed
under fuel-rich conditions. These radicals grow and form aromatic rings. Subsequently, the
formation of larger aromatic rings occurs when a surplus of acetylene molecules is present.
Formation of black carbon results from the coagulation of larger aromatic structures which form
primary soot particles. The growth of black carbon in size and its increase in concentration are
determined both by coagulation (which switches the molecular-length scale into the
macroscopic, particle dimension) and surface growth (which snatches molecules out of the gas
phase), respectively. In this respect, coagulation is responsible for the highly irregular,
disordered structure of soot particles.
At the onset of black carbon formation, the concentrations of PAHs in the flame would
therefore be reduced (Lima et al., 2005). But not all PAHs are transformed into soot particles, as
is evidenced by considerable amounts of PAH residues remaining in both the gas and particulate
phases, and/or adsorbed directly onto the black carbon itself (e.g. Butler & Crossley, 1981;
Koelmans et al., 2006).

23

Fig. 2.3. A conceptual picture proposed for soot formation in homogeneous mixture (adopted from
Bockhorn, 1994).

The magnitude and relative compositions of PAH16 compounds released by combustion


processes vary, resulting from a combination of factors including fuel sources (biomass vs.
fossil fuel, or substrate structures) and combustion/burning conditions (temperature, oxidants).
Literature data taken from various combustion experiments and field measurements is reviewed
here to calculate out the amount (expressed in terms of the mass of PAH emitted per unit mass
of fuel burned) and composition of PAHs. The products are then grouped together as petroleum,
coal or biomass (wood, grasses, agricultural mixtures, and paper) combustions (Fig. 2.4. and
supporting information in Appendix 2.1) Among these groups, coal combustion produces the
highest PAH16 emissions, ranging from ~100 1000 mg/Kg. Biomass burnings emit a wide
range of PAH16 concentrations, varying from ~10 500 mg/Kg, while petroleum combustions
(mainly diesel fuel) release PAHs in the range of 1 10 mg/Kg. Combustion of gasoline fuel
emits PAH16 up to two orders of magnitude less than that of diesel fuel (e.g. Marr et al., 1999;
Miguel et al., 1998).
The relative predominance of aromatic or aliphatic fractions in the fuel structure controls
the amount of PAHs emitted. The higher the aromatic fraction of the fuel structure, the greater
the possible emission level of PAHs becomes. Mastral et al (1998) explained that incomplete
combustion of coal fuel results in the emission of unburned fragments consisting mainly of
aromatics from the coal structure. Following the PAH formation scheme mentioned above,
these unburned aromatics can readily undergo pyrosynthesis with other radicals or other
aromatic rings, thereby building higher molecular weight substances at low combustion
temperatures. Coal contains more aromatic structures than similar fuels. For instance,
bituminous coal consists of aromatic (roughly 45% of the total mass) and aliphatic (ca. 30%)

24

fractions (Yan et al., 2004). Likewise, biomass fuels - in particular wood - also contain high
percentages of aromatic ring components (as polyphenolic compounds in lignin). Phenol
compounds are potential precursors for PAHs (Sharma & Hajaligol, 2003). In addition, aromatic
fuels (represented by benzene) have been observed to emit PAH compounds up to 100 times
larger than those produced by aliphatic ones (acetylene) under the same burning conditions as
reviewed by Richter and Howard (2000). On the other hand, petroleum - which is largely made
up of aliphatic structures will produce low weight PAH16 compounds upon combustion as
compared to the other groups.
High temperature (plus oxidant level) decreases the overall emissions of PAHs (e.g.
Jenkins et al., 1996). Pyrogenic PAHs can be formed in a wide range of temperatures, stretching
from relatively low (ca. 300oC) to high (~1000oC), depending on how condensed the structure
of the precursors is. The more solid the precursors, the higher the temperature needed to crack
the precursors. Several studies have attempted to calculate a scale showing the optimal
conditions under which PAHs are produced. For example, a pyrolysis of cellulose (vegetation)
to produce PAHs occurs optimally between 300 650oC (McGrath et al., 2003). Combustion of
paper emits maximal PAHs at ~300oC (Yang et al., 2005). Neff's review (1973) showed that the
PAHs produced in the series of compounds from naphthalene to coronene are optimally
generated at 780oC. Bituminous coal combustion was shown to release an optimal emission of
PAHs at 800oC (Liu et al., 2000). McGrath et al. (2001) observed that increasing the burn
temperature from 800 to 850 oC lead to significant increases in PAHs resulting from burning
chlorogenic acid and cellulose; beyond this temperature the emission decreased. Khalfi et al.
(2000) observed that PAHs are optimally emitted from wood waste incinerators at 900 - 954 oC.
Jenkins et al. (1996) observed from the experimental burnings of biomass (wood and cereals)
that the magnitude of PAH emission depends also on the flame type. They found that fewer
PAHs are emitted in a vigorous flame, whereas the levels produced were much higher in both
smoldering stages and less robust flames.

25

10000

Emission Factor of PAHs

PAH16 (mg/Kg)

1000

100

10

Coal
Combustions

Biomass
Burnings

Petroleum
Combustions

Fig. 2.4. Emission factors for the 16 PAHs on the EPA's priority list. Results taken from the combustion
of various organic matter classified as coal, biomass or petroleum. Data were evaluated from various
literature sources (see Appendix 2.1. for details to the references).

The composition of PAH emissions serves to characterize the sources. Combustion


processes favor production of unsubstituted compounds as compared to their alkyl homologues
(Lima et al., 2005). The PAH16 represent the most common unsubstituted PAHs which are
produced by such processes. Therefore, they have been quite often examined by and employed
in environmental studies and assessments. Combustion of diesel fuel in modern vehicles
generates high levels of lighter PAHs (~300 Da). It emits no heavier PAHs (>300 Da, e.g.
coronene), unlike gasoline (Riddle et al., 2007a). It is due to this that modern diesel vehicles
have been equipped with more advanced technology, enabling high combustion temperatures
during engine operation which hinder the formation and aid in the breakdown of heavier PAHs.
In contrast, older vehicle technologies produced higher amounts of PAHs (Riddle et al., 2007b).
Increasing engine temperatures during the combustion cycle fosters the production of lower
molecular weight compounds. Liu et al (2000) observed an increase in the relative composition
of 2- and 3-ring structures as operating temperatures increased from 783oC to 843oC, even
though the total overall magnitude of PAHs was significantly reduced.
In order to better understand the source characteristics of PAHs with regards to their
PAH16 composition, various literature data is presented which evaluates their relative individual
and ring-group compositions. The relative composition of a particular compound from a given
source is evaluated by normalizing the individual concentration from the corresponding
PAH16. Fig. 2.5 and Fig. 2.6 show the relative composition (median values) of the common
sources of pyrogenic PAH16. The dashed line represents the composition pattern which

26

characterizes different sources. In addition, the literature data from Sumatra's peatland burning
episode in 2005 (mean values, after See et al., 2007) is also presented for a comparison. These
figures are particularly important, since they provide information on the conditions in the area
where this study took place.
The results shows that the compositions of PAH16 stemming from pyrogenic sources is
relatively similar with respect to high levels (>30%) of naphthalene (NAPH) and low levels
(<5%) of high molecular weight compounds (Fig. 2.5a). This suggests that all biomass, coal and
petroleum combustions emit a significant amount of naphthalene. This is particularly true in the
case of petroleum combustion, in which NAPH comprises roughly 65% of the emitted PAH16.
However, the ring-group composition shows a different pattern for petroleum and biomass-coal
combustion, particularly in their relative compositions of 2 and 3 ring compounds (Fig. 2.6.a).
The decreased levels of 3-ring groups from petroleum combustion is due to a lack of
acenapthylene (ACYN) and acenaphthene (ACEN) produced. In contrast, the high levels of 3ring PAHs for biomass and coal combustion are mainly caused by phenanthrene (PHEN) and
acenaphthylene (ACYN). Within the biomass group, wood burnings emit more ACYN than
grass burnings (Fig. 2.5b). Therefore, the relative compositions of the 2- and 3-ring groups can
probably be used to distinguish between petroleum and biomass-coal combustion simply by
comparing the mass ratios of these ring groups. As far as ring-groups are concerned (Fig. 2.6a),
the patterns emitted by biomass and coal combustion are quite similar, thus making it quite
difficult to differentiate between them. Therefore, the composition of individual compounds
such as fluorene (FLU), fluoranthene (FLA) and pyrene (PYR) and most of the other high
molecular weight compounds can be used as additional clues during separation, due to their
unique signatures and patterns. For example, the ratio of the relative composition of FLA to
PYR for biomass burnings is three-fold higher than that of coal combustion. Therefore, any
mass ratio between those compounds can be developed to help provide clues for source
differentiation.
In comparison to biomass burnings, Fig. 2.5b shows the relative compositions of
individual PAH compounds between two distinct locations. Dumai and Pekanbaru (the capital
city of Riau province) were affected by the tremendous volumes of smoke stemming from
peatland burnings in 2005. It shows that Sumatran peatland burnings emit less NAPH and
ACYN. The relative composition of the PAH16 between Dumai and Pekanbaru is also
significantly different. This is particularly true for the composition of PHEN, ANTH, PYR, as
well as benzo(b)fluoranthene (BbFLA), benzo(k)fluoranthene (BkFLA), benzo(a)pyrene (BaP),
dibenzo(a,h)anthracene (DANTH), and indeno(1,2,3-c,d)pyrene (IPYR). Although there is only
a brief discussion on the composition of PAH16 provided here by the authors, the evidence
suggests that PAHs derived from the peatland smoke experienced alterations due to

27

decomposition or additions from local sources. This data can therefore be a useful aid in
comparison for PAH assessments in aquatic systems of those particular studied areas.

A
Relative Composition (%)

35

Pyrogenic PAHs

30

40

Biomass
Burnings
Coal
Combustions
Petroleum
Combustions

25
20
15
10
5

Relative Composition (%)

65,1%

Biomass Burnings

35
30

Wood Burnings

25
20

Grass Burnings

15
10
5
0

C
Relative Composition (%)

35

Sumatra Peatland Burnings in 2005

30

Dumai

25

Pekanbaru

20
15
10
5
0

Fig. 2.5. Relative composition of 16 individual PAHs of the EPA priority list from (A) common
pyrogenic sources including coal, biomass and petroleum combustion (median values) (see Appendix 2
for references); (B) a subset of biomass burning sources comprising wood and grass burnings; and (C)
Sumatra peatland burnings (mean values, after *See et al., 2007). Note: the y-axis for graph A and B is
different.

B
80

Pyrogenic PAHs

70

Biomass
Burnings
Coal
Combustions
Petroleum
Combustions

60
50
40
30
20
10
0

Relative Composition (%)

Relative Composition (%)

80

Sumatra Peatland Burnings in 2005*

70

Dumai
Pekanbaru

60
50
40
30
20
10
0

2 rings

3 rings

4 rings

5 rings

6 rings

2 rings

3 rings

4 rings

5 rings

6 rings

Fig. 2.6. Relative composition of the ring groups of the 16 PAHs on the EPA priority list from (A)
common pyrogenic sources including coal, biomass and petroleum combustion (median values) (see
Appendix 2 for references); (B) Sumatra peatland burnings (mean values, after *See et al., 2007).

28

2.2.3.

Petrogenic PAHs
PAHs are substantial components of crude oil and its petroleum products. They are

considered to be the toxic fraction. Components of petroleum hydrocarbons are generally


grouped into four class: the saturates (n- and branched-chain alkanes and cycloparaffins), the
aromatics (mono-, di-, and polynuclear aromatic compounds containing alkyl side chains and/or
fused cycloparaffin rings), the resins (pyridines, quinolines, carbazoles, thiophenes, sulfoxides,
and amides), and the asphaltenes (extended polyaromatics, naphthenic acids, suifides,
polyhydric phenols, fatty acids, and metalloporphyrins) (Sugiura et al., 1997; Leahy & Colwell,
1990). The other significant components of the aromatic hydrocarbons found in petroleum are
monoaromatic (single-ring) compounds including Benzene, Toluene (or methylbenzene),
Ethylbenzene and Xylene (all isomers of dimethylbenzene), or so-called BTEX. PAHs and
BTEX are two aromatic components which are generally used to identify and characterize
sources of petroleum, in particular PAH compounds (Wang et al., 1999).
The aromatic fraction varies among different types of crude oil, but it can reach up to
50% of the total weight of the petroleum. Neff (1979) published a review showing that the
aromatic content of mineral oils varies from 7 % 34 %. Ryder et al. (2002) reported that
aromatic values ranged from 18% to 41% for BP crude oil samples. Heavy oil fuel transported
by the sunken tanker Prestige consisted of 50% aromatic hydrocarbons (Saco-lvarez et al.
2008). However, PAHs normally make up a more modest part (up to 20%) of the overall
aromatic fraction. Requejo et al. (1996) calculated that PAHs constituted somewhere between
1% and 20% of the total C12+ aromatic fraction in various marine crude oils. Stated otherwise, if
we assume PAHs to be roughly 20% of the aromatic fraction (which is about 50% of the
petroleum), the total PAH fraction in the oil would only be about 10% of the mass. However,
PAHs are in fact found to be largely variable in different sorts of petroleum, reaching values of
up to 13%. For instance, Neff (1979) found that the total tri- to hexacyclic PAHs varied only
from 0.2 - 7.4%, while Requejo et al (1996) estimated that PAHs constitute up to 12.9% of oil
originating from three marine source rocks: carbonates, marine shales and fluvio-deltaic oils.
Therefore, any oil spills which occur would add significant amounts of PAHs to the aquatic
environment.
In contrast with pyrogenic origins, the composition of petrogenic PAHs is notably
characterized as having an abundance of alkyl-homologue PAHs (substituted) as compared to
the parent (un-substituted) forms (Youngblood & Blumer, 1975; Lafamme & Hites, 1978; Neff,
1979). This abundance of alkylated PAHs is due to lower formation temperatures during the
catagenic process, which takes place over geological periods. Alkyl-substituted PAHs may
comprise 80 90% of the total PAHs in crude oil (Saravanabhavan et al., 2007) and be used for
chemically fingerprinting the source of PAHs in oil spills (e.g. Boehm et al., 1997). Wang et al.
(1999) reviewed oil spill identification attempts, showing that the ratios between the total of 3-6

29

ring parent PAHs and the total alkylated PAH homologues were 5%. This data was drawn
from PAH analysis performed on over 60 different crude oils and petroleum products, including
Arabian crude oils, California oils, UK Brent oil, Alaska Cook Inlet 1-3, Iranian Heavy oil,
Russia Komi, Norway Statfjord Oil, Terra Nova, jet fuel, diesel, Bunker C and many others.
Fig. 2.7 shows a typical bell curve distribution for alkyl homologues relative to their
parent compounds in several types of oil (adopted from Wang et al., 1999). The alkyl group
substitutes range mostly from one (methyl-) to four (tetraalkyl-) carbon atoms, often denoted as
C1-C4. In most cases, the degree of alkylation is presented as the sum of all isomers
corresponding to a given level of alkylation (total C1, total C2, and so on). For example, total
C2-fluorene is the sum of all possible isomers including 1-Ethylfluorene, 1,2-Dimethylfluorene,
1,5-Dimethylfluorene, and 1,6-Dimethylfuorene. Alkyl group substitution is usually used to
distinguish the petroleum sources for PAHs, including (C1-C4)-naphthalenes, (C1-C3)fluorenes, (C1-C4)-phenanthrenes, (C1-C3)- dibenzothiophenes, C1-fluoranthene/pyrene, and
(C1-C4)-chrysenes (Wang et al., 1999). However, the level of alkylation is often represented
differently in different studies. Some apply only a few, limited alkyl homologues, which
effectively turn out to be the same as total C1, C2 and so on. Due to this, the degree of
alkylation should be carefully interpreted to avoid mistakes in classification.
The composition of the PAH16 compounds is quite different between crude oil and oil
spills from tanker accidents (Fig. 2.8). The main differences are that the crude oils are
dominated by NAPH, while 46 ring compounds in crude compose only relatively small
percentages. Wang et al. (1999) found that naphthalene and its alkyl-homologues (C1-C4)
comprised up to 86% of total PAHs in Diesel No.2 and up to 99% in Jet B fuel. In contrast, oil
spills experience increasing compositions of primarily 4-ring compounds. It is most likely that
oil spills evidence contamination by pyrogenic PAHs. For example, Wang et al (2004)
demonstrated a clear signature of pyrogenic PAH input in the Detroit oil spills of 2002. These
pyrogenic inputs were attributed to combustion and motor lubrication processes. The lube oil
tested in the samples was waste lube oil.

30

homologues Bell Shape distribution

Fig. 2.7. An example of a bell-shaped distribution (red dashed line): alkylated homologues PAHs and
other EPA priority PAHs in ASMB crude, three oil products, and a tarball sample from British Columbia
(adopted and modified from Wang et al., 1999).
B

Relative Composition (%)

30
25

55,5%

Petrogenic PAHs
Oil Spills

20
15
10
5
0

Crude Oil**

80
Relative Composition (%)

70
60

Oil Spills
Crude Oil**

50
40
30
20
10
0
2 rings

3 rings

4 rings

5 rings

6 rings

Fig. 2.8. Relative composition of (A) the individual and (B) the ring groups of the 16 PAHs of the EPA
priority from petrogenic sources including oil spills and crude oils. The crude oil data represents mean
values (after ** Requejo et al., 1996).

31

2.2.4.

Source Apportionment
This section elaborates the source apportionments of anthropogenic PAHs based on

molecular weight ratios and isomeric ratios. Source apportionment of anthropogenic PAHs in
the aquatic environment is very challenging. This is due to that fact that anthropogenic PAHs,
particularly those of unsubstituted compounds, coexist with substances from various sources.
Also, the relative composition of the individual PAHs experience transformation processes such
as photo-oxidation and oxidation leading to changes in pattern from sources to pools (Soclo et
al., 2000). However, source apportionment is possible due to specific characteristics of
particular molecular weight isomer ratios as a fingerprint from two main anthropogenic
petrogenic and pyrogenic sources, as reviewed by Yunker et al. (2002). PAHs ratios have
been widely applied (e.g. Budzinski et al., 1997; De Luca et al., 2005). But, the interpretation in
most studies is limited to infer sources as to major categories: petrogenic and pyrogenic.
2.2.4.1. Molecular Weight Profile
As described earlier, petrogenic PAHs are generally characterized high proportions of
low molecular weight (LMW), typically 2- and 3-ring compounds. In contrast, pyrogenic PAHs
tend to feature higher levels of high molecular weight (HMW), 4- to 6-ring substances. In
response to these generic profiles, the mass ratio of LMW/HMW has been widely introduced
as a benchmark for distinguishing petrogenic from pyrogenic PAH sources (Neff, 1979; Soclo
et al., 2000; De Luca et al., 2005). Therefore, a LMW/HMW ratio with a value greater than
1 indicates petrogenic origins. However, this ratio should be cautiously interpreted, since the
emission sources of petroleum and biomass burnings produce high LMW profiles (Fig. 2.6).
The emission source profile can be changed during the atmospheric phase due to both
photo-degradation and transformational processes before the PAHs ever reach the acceptor
substrates. For example, gas-phase PAHs react rapidly with OH radicals in the present of NOx
gases in the atmosphere. This transformation leads to a PAH atmospheric lifetime of roughly 1
12 hours total (Atkinson et al., 1987). Behymer & Hites (1985) performed a photolysis
experiment with particulate PAHs. They found that the degradation of particle-borne PAHs also
depends on the type of substrates to which they were adsorbed. The lifetime of PAHs adsorbed
onto fly ash is greater than 29 hours, and black carbon acting as a substrate might extend this
lifetime to >1000 hours. Such natural particles retard degradation effectively and may facilitate
a long journey. The longer the atmospheric residential time, the greater the profile shifts which
occur towards HMWs. Therefore, PAH profiles in remote areas (those far away from the
source) tend to favor HMWs (Fernndez et al., 2002). Conversely, PAH profiles for regions
close to the original source closely mirror the profiles for specific emission sources given above.
In aquatic environments, we should expect suspended particulate matter to possess signatures
which are to a large extent similar to those of the surrounding atmospheric profiles. But

32

sediment profiles can also be expected to differ somewhat from those of SPM (Fernandes &
Sicre, 1999). They will tend to favor HMWs due to the complex sedimentary degradation
processes which take place. LMW compounds are very susceptible to biodegradation due to
their high solubility. However, HMWs are strongly associated with carrier sediment particles, a
relationship which can to a large extent retard microbial attack of particles adsorbed upon their
surfaces. As a direct consequence, an increase in the ratio of LMW/HMW would be
expected in the sediment.
2.2.4.2. Diagnostic Mass Ratio of Isomers
Isomers of unsubstituted PAHs have different kinetic and thermodynamic characteristics
due to varying structural stability. Compounds with a linear structure (e.g. anthracene,
benzo(a)anthracene, benzo(a)pyrene, dibenzo(a)anthracene) are less stable than corresponding
isomers containing angular, branched structures. Therefore, greater proportions of less stable
compounds, e.g. anthracene over phenanthrene (MW=178), are produced during combustion
when compared to other isomers. Budzinski et al. (1997) demonstrated that the ratio of
PHEN/ANTH was temperature dependent, with its value decreasing as the temperature
increased. At 300, 700 and 1000 K, the P/A values were 49, 8.3, and 5.5, respectively. In
contrast, lower thermal production (e.g. during petroleum maturation) lead to higher
PHEN/ANTH ratios. Phenanthrene and anthrene are therefore highly salient compounds when
separating pyrogenic from petroleum sources. A similar hypothesis has been proven for FLA
and PYR. With respect to this, pairs of parent PAH isomers have commonly been used for the
purpose of source apportionment, i.e. compounds with molecular weights of 178 (PHEN,
ANTH), 202 (FLA, PYR), 228 (BaA, CHRY) and 276 (BPERY, IPYR). Yunker et al. (2002)
reviewed those mass groupings and proposed the ratios ANTH/(ANTH+ PHEN),
FLA/(FLA+PYR), BaA/(BaA+CHRY) and IPYR/(IPYR+BPERY) as the best potential parent
ratios for discerning between natural and anthropogenic sources.

PAHs from combustion

sources have typical values of ANTH/(ANTH+PHEN) > 0.1; FLA/(FLA+PYR) >0.5, and
BaA/(BaA+CHRY) > 0.35. In contrast, PAHs associated with petroleum sources like crude oil
typically have values of the same isomeric ratios of <0.1, <0.4, and <0.2, respectively (Table
2.1). These mass ratios consist of the mass of the less stable compound divided by the total
isomer mass for a particular molecular weight.

33

Table 2.1. Diagnostic mass ratios for identification of petrogenic and pyrogenic sources
LMW/HMW

ANTH/(ANTH+PHEN)
MW 178

FLA/(FLA+PYR)
MW 202

Identification
Values a
Petrogenic
>1
<0.1
<0.4
Combustion
<1
>0.1
>0.5
Mixture
petrogenic/
combustion
Petroleum
0.4 0.5
combustion
a
The ratio of MW178, 202, 228, 276 after Yunker et al. (2002)

BaA/(BaA+CHRY)
MW228

<0.2
>0.35

IPYR/(IPYR+BPERY)
MW 276

<0.2
>0.5

0.2 0.35
0.2 0.5

For a better overview of the application of those values, various ratios of PAH isomers
from common classes stemming from different sources are presented here. The emission of
fossil fuel combustion and biomass burning as well as crude oil and spillage events, for
example, are evaluated from various literature sources (see Appendix 2). In this evaluation,
different data from the literature are classified into the four major groups described earlier,
including crude oil and spillage (petrogenic sources), coal and petroleum combustion,
wood/bamboo burnings and biomass and peat burnings (pyrogenic sources).
The ratios of MW178, 202, 228 and 276 for those major groups of sources from the
literature generally lies within the expected range of petrogenic and pyrogenic values. This
suggests that these ratios can be used to distinguish petrogenic versus pyrogenic sources (Fig.
2.9 & 2.10). For the petrogenic sources, the ratios with MW178, 202, and 228 are more
sensitive than MW 276. It is due to the fact that the concentration of HMW PAHs in crude oil
and spillages are relatively low. Therefore, differences between the isomers must not be large in
order to lower the ratio quite drastically. Likewise, the ratios for MW 178, 202 and 228 clearly
separate the values of wood/bamboo and biomass burnings from those found for petrogenic
sources. This also suggests that these values are acceptable for partitioning pyrogenic and
petrogenic sources. However, there are several data which do not follow suit, as seen in Fig. 2.9.
First, in the crude oil and spillage groups the ratios of MW202 and 228 for Detroit oil spills
(after Wang et al., 2004) are higher than the same ratios of the other crude oil and spills. They
even lie in the pyrogenic data region. Wang et al. (2004) explained the pyrogenic signature of
the spill by attributing it to combustion and motor lubricant processes, stating that the lube oil in
the spill samples was waste lube oil. Second, these combustions produced PAHs whose the
ratios also fall in the region of petrogenic sources, for example, coal (MW202) and diesel
(MW178) combustion, paper combustion (MW 178 and 228), and the burning of corn grasses
(MW 178). These disagreements between apportionment values indicate that application should
be carefully interpreted with respect to those sources. Because there may be diverse sources of
PAHs in the aquatic compartment, these ratios should therefore be complementarily applied.

34

To achieve this, the best way to represent the comprehensive analysis of the abovementioned ratios is through the use of cross-plot graphs, which allow for optimal discrimination
ability between both petrogenic and pyrogenic sources and within the pyrogenic sources
(petroleum vs. wood/biomass burnings). One possible combination is a cross plot between
MW202 (x-axis) to the other remaining ratios (y-axis). The primary consideration here is that
compounds with a molecular mass of 202 have been proven to possess a large thermal range of
stability with respect to their formation by heat (Yunker et al., 2002). The axes are arbitrary.
Fig. 2.11 shows the cross-plot graphs of the given isomer ratios are a useful tool for
apportioning petrogenic vs. pyrogenic sources. We can see that each of the ratio cross-plots
(MW202 vs. MW178, 228, and 276) shows a relatively clear segregation between oil (left
quadrant) and combustion sources (right quadrant), although a tendency still is likely for the
values to lie in the mixture area. Interestingly, a clear segregation between fossil fuel
combustion and biomass burning is given by the plot of the ratios of MW202 vs. MW178.
However, the application of such cross-plots must be comprehensively and carefully interpreted
within the context of the specific environmental conditions found at a given site.

35

Alaskan North Slope


Crude

1,00

Crude Oil and Spillages


0,90
0,80

Oil spills "Oil Tanker


Enrika"

0,70

Oil spills "Enrika"


Beach Sample1

0,60

Oil spills "Enrika"


Beach Sample2

0,50

Detroil Oil Spill_1

0,40
Detroil Oil Spill_2
0,30
0,20

Detroil Oil Spill_3

0,10

Pyrogenic

0,00
MW 178

1,00

MW 202

MW 228

MW 276

Coal and Petroleum Combustions

Petrogenic

Coal- Cement Industry

0,90

Bituminous Coal

0,80

Charcoal

0,70

Coal Briquette

0,60

Coal Domestic

0,50

0,30

HO + natural gas, Ind


Boiler
Diesel, Industrial
Boilers
Fuel Oil - HOP

0,20

Diesel 2 - HDV

0,10

Heavy oil, Industrial


Boilers
light-duty Diesel
vehicles
Pyrogenic

0,40

0,00
MW 178

MW 202

MW 228

MW 276

Fig. 2.9. Mass ratio distribution of specific PAH isomers from various (A) petrogenic (crude oil and
spillages) and (B) pyrogenic of coal and petroleum combustion. The lines represent reference values used
for apportionment petrogenic (dashed lines) and pyrogenic (solid lines) sources (after Yunker et al.,
2002). The ratio values were calculated from many sources (see: Fig. 2.4 including the references).

36

1,00

Almond

Wood/Bamboo Burnings

Walnut

0,90

Chinese clay woodstoves


Pine-domestic heating

0,80

Fir

0,70

Eucalyptus wood 1

0,60

Wood, Domesting Heating

0,50

Oak Wood
Lao trad. woodstoves

0,40

Thai bucket woodstoves

0,30

Cambodian trad. Woodstoves


Vietnam trad. Woodstoves

0,20

Eucalyptus wood 2

0,10

Paper
Bamboo

0,00
MW 178

MW 202

MW 228

MW 276

Pyrogenic
Petrogenic

1,00

Sugarcane

Biomass and Peat Burnings

0,90

Pampas grass

0,80

Mixed ryegrass

0,70
Rice grasses
0,60
Corn grasses

0,50
0,40

Agricultural Debris

0,30

Peat Burning Dumai

0,20

Peat Smoke Pekanbaru

0,10

Pyrogenic

0,00
MW 178

MW 202

MW 228

MW 276

Petrogenic

Fig. 2.10. Mass ratio distribution of specific PAH isomers from other various pyrogenic sources of (A)
wood/bamboo and (B) biomass and peat burning sources. The lines represent reference values used for
apportionment petrogenic (dashed lines) and pyrogenic (solid lines) sources (after Yunker et al., 2002).
The ratio values were calculated from many sources (see: Fig. 2.4 including the references).

37

Petroleum

Petroleum
Combustion

Grass/Wood/Coal
Combustion

1,0

IPYR/(IPYR+BPERY)

0,9

Grass/Wood/Coal
Combustion

0,8
0,7
0,6
0,5

Petroleum
Combustion

0,4
0,3
0,2

Petroleum

0,1
0,0
0,0

0,1

0,2

0,3

0,4

0,5

0,6

0,7

0,8

0,9

1,0

1,0
0,9
BaA/(BaA+CHRY)

0,8

Combustion

0,7
0,6
0,5
0,4
0,3

Mixed Sources

0,2
0,1

Petroleum

0,0
1,0

0,0

0,1

0,2

0,3

0,4

0,5

0,6

0,7

0,8

0,9

1,0

ANTH/(ANTH+PHEN)

0,9
0,8
0,7
0,6

Combustion

0,5
0,4
0,3
0,2
0,1

Petroleum

0,0
0,0

0,1

0,2

0,3

0,4

0,5

0,6

0,7

0,8

0,9

1,0

FLA/(FLA+PYR)

Fig. 2.11. Common cross plots of the mass ratio of 202 vs. 178, 228, 276 and the determination endmember values (dashed line) for source apportionment (after Yunker et al., 2002). The literature values
from various petrogenic (blue triangles) and pyrogenic sources: coal and petroleum combustion (red
circles); wood/bamboo burnings (gray circles); biomass and peat combustion (green circles) are plotted to
ascertain the extent of which the plots compromise the variability of environmental data. These cross
plots show that combination of MW 178, 202, and 228 plots are able to generally separate substances
with petrogenic (crude oil) and pyrogenic origins.

38

2.3.

Distribution in Aquatic Compartments


PAHs evidence widespread distribution in the hydrosphere of rivers, estuaries and coastal

areas (e.g. Gschwend & Hites, 1981; Mitra et al., 1999; Golomb et al., 2001; Woodhead et al,
1999) where the sources are mostly concentrated. This also holds true for far-distance and
remote aquatic systems such as mountain lakes (e.g. Fernndez et al., 1996, 2002; Vilanova et
al., 2001) and the Southern Ocean (e.g. Valero-Navarro et al 2007; Mazerra et al 2000). In the
aquatic system itself, PAHs were observed in most compartments either as truly dissolved
substances or as being associated with dissolved organic matter. This section briefly elaborates
upon PAH distribution in the main aquatic compartments of sediment, SPM and the water
solution, with a particular focus on rivers, estuaries and coastline areas.
2.3.1.

Surface Sediment and Grain Size Fractions


It is widely accepted that sediment layers record diverse environmental changes, which

include PAH loads both over short-term and geologic periods of time. Sediment therefore plays
an extremely important role in pollution assessment. Surface sediments (ca. 20 m depth) in
rivers, estuaries and coastal waters are especially dynamic, since remobilization and
resuspension occur constantly due to physical processes. Surface sediments in such dynamic
environments are therefore considered not only to be a sink (sequestration) but also a source
(desorption) of PAHs.
A sediment's capacity to concentrate and retain elements and hydrophobic organic
compounds results primarily from two sets of physical (grain size and type) and chemical
(organic composition and content) properties (Horowitz & Elrick, 1987). For example, a fine
grain size provides a larger amount of total surface area per unit mass for adsorption and is often
associated with a large organic matter fraction (a natural geosorbent for most hydrophobic
compounds). Luthy et al. (1997) proposed a schematic conceptual model for the sequestration of
hydrophobic organic contaminants by the sediment (Fig. 2.12). In this concept, sediment is
composed of various aggregates/particles of different origins and sizes, such as mineral
particles, combustion residues and plant debris. The capacity for sequestration therefore refers
to a combination of specific interactions of substrates with organic pollutants (i.e. binding
energy and rates of adsorption/desorption). These interactions include both particles and nondissolved liquids in the aqueous medium, such as oils, tars, and solvents adhered to or trapped in
sediment pores. The sequestration of hydrophobic organic pollutants in sediments is a
combination of factors including diffusion limitation, sorption and partitioning. If the
distribution of individual PAH molecules in sediment is controlled by an equilibrium governing
sorption to the solid surface, we might expect that the fine (or mud) fraction of the bulk
sediment should accumulate significantly high levels of pollutants due to its large surface area

39

and high organic carbon content (Karickhoff et al., 1979; Maruya et al., 1996; Tolosa et al.,
2004).

Fig. 2.12. A conceptual model for the sequestration of hydrophobic organic contaminant by geosorbents.
NAPL= nonaqueous-phase liquids; SOM= sorbent organic matter. Sorption mechanisms include (A)
absorption into amorphous or soft natural organic matter or NAPL; (B) absorption onto condensed or
hard organic polymeric matter or combustion residue (e.g. soot); (C) adsorption onto water-wet organic
solvent (e.g. soot); (D) adsorption onto exposed water-wet mineral surfaces (e.g. quartz); (E) adsorption
onto microvoids or microporous mineral (e.g. zeolites) with porous surface at water saturation <100%
(adopted from Luthy et al., 1997).

PAHs in surface sediments are, in fact, heterogeneously distributed among the various
particle-size fractions (e.g. Maruya et al., 1996; Budzinski et al., 1997; Tolosa et al., 2004; Prahl
& Carpenter, 1983; Simpson et al., 1998; Wang et al 2001; Rockne et al., 2002; Ahrens &
Depree, 2004). PAH enrichment is not found solely in the mud fraction. It has also been
observed in the sand fraction, which can also carry significant loads of various PAHs. For
example, Rockne et al. (2002) found that PAH levels tend to increase with increasing grain size
in sediment samples taken from Piles Creek and Newton Creek in the New York/New Jersey
harbor area. Similar trends were reported by many other studies (e.g. Ahrens & Depree, 2004;
Oen et al., 2006; Simpson et al., 1998). Therefore, in this study we chose to investigate the PAH
distributions in two general groups of sediment sizes as classified using the Wentworth-Udden
scale: sand (coarse) 2 mm 63m; and mud (fine) <63m, which includes the silk and clay
fractions.
Sedimentary organic matter (SOM) controls the distribution of PAH in those fractions.
The PAH sorption capacity of SOM depends both on its structure and composition (Grathwohl,
1990; Huang & Weber, 1997; Johnson et al., 2001). Structure and composition vary among

40

sources e.g. peat, fragmented plant materials, black carbon, and kerogen (Grathwohl, 1990). But
SOM is essentially the product of diverse geochemical alterations, ranging from biopolymer
precursors (e.g. carbohydrate, protein, lipids, lignin, tannin and pigments) to geopolymers (e.g.
fulvic, humic, humin substances and kerogen) through complex diagenesis processes. During
diagenesis, SOM experiences compositional changes in polarity and aromatic carbon content
which controls its reactivity with hydrophobic organic compounds (Garbarini & Lion, 1986;
Gauthier et al., 1987; Luthy et al., 1997; Chiou et al., 1998). Therefore, a correlation between
PAH and SOM is often helpful for pinpointing the significant role of organic matter (e.g.
Karickhoff et al., 1979; Means et al., 1980; Kim et al., 1999; Guinan et al., 2001; Maskaoui et
al., 2002; Viguri et al., 2002). Such a correlation applied to sediment size fractions can help
elaborate the preferential particle associations of PAHs (e.g. Prahl and Carpenter, 1983).
Coarse and fine sediment fractions contain different types of organic matter which affect
PAH distributions. Increasing PAH content for the coarse sediment fraction has been
acknowledged, due to the presence of organic particles that have a high affinity for PAH
sorption. This is despite the fact that the OM fraction normally represents only a tiny portion of
the total sediment mass. Ghosh et al. (2003) petrographically examined carbonaceous particles
(coal, coke, charcoal, pitch, cenospheres, and wood). These particles are typically in the size
range of 250m1mm and comprise only about 5-7% of the total mass, however, they account
for 90% of adsorbed PAHs. Oen et al. (2006) emphasized that the presence of decomposed
vegetable debris and shiny black particles dramatically increases the PAH content of the sand
fraction. Since combustion processes leave behind soot and black carbon materials as residue, it
is entirely possible that PAHs have already been adsorbed onto those particles before they ever
reach the sediment. Therefore, examining the PAH content of the various sediment size
fractions can provide important information on the mode of PAH transportation into the aquatic
environment. On the other hand, humic substances are organic matter specifically associated
with the fine fraction. This is due to the fact that the sorption of humic substances (normally in
the form of dissolved organic matter, DOM) onto the fine fraction is typically a direct function
of uptake onto its exceedingly large surface area. However, PAHs may be reluctant to associate
with fine fraction-OM in the presence of combustion-derived OM (e.g. Prahl & Carpenter,
1983). Furthermore, different sources of SOM in the fractions have also been recognized. Evans
et al. (1990) illuminated two different types of organic materials which were responsible for
bimodal distribution of PAHs in coarse (>250m 2mm) and fine (<63 m) fractions. First,
fragmented plant materials are assumed to be responsible for high levels of organic matter (OM)
in the coarse fraction. This is in addition to combustion-associated particles such as coal, soot or
black carbon. Second, condensed organic matter (humic substances) is mostly associated with
the fine fractions. Therefore, PAH enrichment in the sand fraction might be an indication of a
strong anthropogenic source for PAHs. An investigation of

these PAH-sized fraction

41

associations would provide us with significant additional information about PAH delivery
modes into aquatic environments, whether as a result of combustion-derived particle association
or from the sorption equilibrium on the surface of fine sediments.
2.3.2.

Suspended Particulate Matter and Water


Suspended particulate matter (SPM) is the main carrier by which most terrestrially-

derived materials - including anthropogenic pollutants - are transferred from land to aquatic
environments. It also provides a fundamental link between chemical constituents and the water
column, bed sediments and food chain (Turner & Milward, 2002; Suzumura et al., 2004). The
term SPM refers to all particulate matter with different natures and origins, but is operationally
defined as those materials retained by a filter with a specific pore size. Therefore, the definition
of SPM can operationally vary between studies. However, a maximum pore size of 0.7 m
(GF/F) is often employed for differentiating between the particulate and dissolved phases from
river, estuary and coastal waters for pollution and geochemistry studies (e.g. Zhou et al., 1998;
Luo et al., 2006; Suzumura et al., 2004; Boldrin et al., 2005; Gebhardt et al., 2004).
PAHs move from terrestrial environments into the oceans via river, estuary, and coastal
pathways. The concentrations of PAHs and SPM have been shown to be positively correlated
(e.g. Fernandes et al. 1997). This means that PAH loads entering the ocean are closely linked to
the overall SPM load emerging from riverine, estuarine and coastal drainage basins. Any change
in the hydrologic cycle (e.g. increased precipitation or extreme floods caused by climate
change) can therefore modify and possibly intensify the temporal load of PAHs. The
implications of flood events for PAH loading have been studied. For example, Witt and Siegel
(2000) observed a significant flux (two orders of magnitude higher than normal) of PAHs into
the Baltic Sea, stemming from municipal and industrial areas in the Oder River basin as a
consequence of a 1997 flood event. Sicre et al. (2008) calculated that 90% of the annual load of
particulate PAHs flowing from the Rhne River (France) into Mediterranean Sea took place
during flood episodes in 1994. These facts are extremely important for tropical rivers,
particularly in Sumatra where overall precipitation is high, and floods are occurring more
frequently due to climate changes. Thus, the transport of pollutants via Sumatran waterways
into the ocean should be expected to be significant.
Interactions between hydrophobic organic pollutants and SPM depend on the composition
of SPM in the water column. This is due to the fact that different types of particles embody
various binding capacities for specific, hydrophobic pollutants. In general, SPM in riverine,
estuarine and coastal waters represents a composite of lithogenous, hydrogenous, biogenic and
anthropogenic particles. Lithogenous particles are inorganic materials derived from the
weathering of rocks and other substances in the Earth's crust, which are composed mainly of
quartz and other aluminosilicate minerals. Hydrogenous matter is generated in-situ by chemical

42

processes, resulting in such materials as humic substances, carbonates and both iron and
manganese oxides. They occur either as coatings or discrete phases. Biogenic particles include
those stemming from microorganisms, plankton and the decaying remains of macroorganisms
and terrestrial plant debris. Bio-particles can also refer to those derived from proteins,
carbohydrates, lipids and pigments. Anthropogenic particles consist mainly of combustion byproducts such as dust and fly ash, but also include other widely varied synthetic materials such
as plastic, tar, solvents and surfactants (Turner & Millward, 2002). However, SPM particles are
often divided into inorganic and organic particles based on their chemical properties. Particulate
inorganic material (PIM) includes lithogenous matter (minerals and insoluble salts), whereas
particulate organic materials (POM) are composed of a rather broad mixture of hydrogenous,
biogenic and anthropogenic particles. PIM and POM are straightforward measures for the
overall SPM composition in a sample. The latter also acts as an effective sorbent for
hydrophobic organic pollutants.
The molecular composition of PAHs in particulate matter is somewhat different from that
of those found in sediments. Luo et al. (2006) concluded that SPM samples taken from the Pearl
River estuary contained large amounts of both 2- and 3-ring PAHs. On the other hand, these
sediments were also characterized by high levels of 5- and 6-ring PAHs. Similar findings have
been reported for other riverine and estuarine systems (e.g. Witt, 1995; Shi et al., 2005). These
patterns were driven by the fact that the SPM was continually receiving fresh PAH inputs, either
from the atmospheric deposition of combustion by-products or from direct oil spills, which are
predominantly characterized by low molecular weight compounds (see 2.1.2). However, this is
not always the case in every situation. PAH profiles in SPM samples represent the local
conditions where the transfer of high molecular weight molecules into the aquatic environment
occurs. This can also be mainly derived from land-water interactions. For instance, Witt &
Siegel (2000) observed that the distribution of PAHs resulting from Oder River floods were
characteristic of high-combustion profiles, which are predominated by HMW compounds. On
the other hand, the concentration of LMW PAHs in the water column and sediments was subject
to great variation due to differing degradation processes (photonic and microbial). Therefore,
examination of a specific PAH profile in SPM yields clues to the fate of PAHs in a particular
environmental setting.
2.3.3.

Water Solution as dissolved PAHs


In addition to SPM association, PAH compounds can alternately remain in solution as

truly dissolved substances or also be bound by

dissolved organic matter. Differentiation

between those two aqueous fractions is important for particular purposes like bioavailability or
toxicity studies (Hawthorne et al., 2007). However, for the general assessment of the
distribution of PAH in natural waters, separation of those two aqueous fractions is not

43

substantial. This is due to the fact that the fate of truly dissolved PAHs is appreciably controlled
by the existence of dissolved organic matter (DOM) as a geosorbent. Schlautman & Morgan
(1993) observed in an experiment that PAH-DOM binding could be completed within a
timeframe of only 3 minutes. As a consequence, the free fraction in natural waters is unstable
and is readily associated with DOM. Thus, it is quite challenging in terms of analytical
techniques to effectively separate these two aqueous fractions. And this is despite the
availability of various analytical techniques such as fluorescence quenching, purging or
sparging techniques, solid-phase microextracion (SPME), equilibrium dialysis, solubility
enhancement, ultrafiltration, size exclusion chromatography, and liquid-liquid extraction
(Burkhard, 2000 and references therein). This study simply considers these two fractions of
PAHs as the "dissolved phase", which operationally defines dissolved PAHs as those passing
through the GF/F. The most popular extraction techniques for determining dissolved PAHs are
solid phase extraction (SPE) and stir bar microextraction (SBME) (Falcon et al., 2004;
Fernndez-Gonzles et al., 2007; Poerschmann et al., 1997).
The composition of unsubstituted PAHs in the dissolved phase depends on the solubility
and hydrophobicity (Kow) of a given compound. The former decreases as molecular weight
increases and the latter increases with increasing molecular weight (see Appendix 1). With
regard to a compounds solubility, we can expect a predominance of low molecular weight
PAHs (2 and 3 rings) as compared to HMW compounds in the absence of DOM, salinity and
pH effects. The opposite also holds true. Furthermore, angular compounds are more soluble
than corresponding linear isomers, such as phenanthrene (1.18 mg/L) compared to anthracene
(0.08 mg/L) or fluoranthene (0.26 mg/L) compared to pyrene (0.14 mg/L). With regards to Kow,
we might assume that the presence of DOM should increase the concentration of HMW PAHs
in the dissolved phase. However, interactions between PAHs and DOM prove themselves to be
DOM source-dependent (Liu & Amy, 1993). Therefore, examining the concentration and
composition of materials in the dissolved phase can reveal the relevance of particular
environmental settings on the fate of particular PAHs.

2.4.

The fate of PAHs in the water: a partitioning concept and the role of
natural organic matter
The fate of hydrophobic organic pollutants in aquatic systems has been widely studied,

including which factors control their distribution in two immiscible phases (particulate and
water solution). This is due to the fact that transfer from the water solution onto particulate
matter can possibly act to retard these molecules from degradation, volatilization and desorption
(Mackay & Powers, 1987).

44

One way to understand the fate of the hydrophobic chemicals is to evaluate their partition
coefficients (e.g. Hellou et al., 2005; Ko & Baker, 1995; Wang et al., 2001; Zhou et al., 1999).
This partitioning involves complex dynamic sorption interactions between the molecules of
solutes, solvents and sorbents, which act in concert. Weber Jr. et al. (1991) explained that the
partitioning of a given chemical into two phases is a result of its relative affinities for sorbate
molecules and both the solvent and sorbent phases competing for thermodynamic balance.
Therefore, partition represents an equilibrium condition between molecules in two phases. Thus,
the concentration of given compounds in the solid phase depends on its concentration in the
water solution, which is often described as the sorption isotherm.
There are three mathematical models that have been widely called upon to quantify and
interpret the partition of hydrophobic organic contaminants between the aqueous and solid
phases, namely the Linear, Langmuir, and Freundlich models as reviewed by Voice & Weber Jr
(1983), Weber Jr. et al. (1991), and more recently by Huang et al. (2003). The Linear model is
built upon the hypothesis that the sediment organic matter has no limitation of sites and surface
space for sorption. It assumes that sedimentary organic matter will therefore act as a gel- or
liquid-like phase. The Langmuir model was originally developed for the adsorption of gases
onto solids, and generally assumes that (1) the energy of adsorption is constant and independent
of surface coverage, (2) adsorption occurs only on localized sites and there is no interaction
between adsorbed molecules and (3) the maximum adsorption possible is that of a complete
monolayer (Voice & Weber Jr., 1983). However, this model does not properly fit the sorption
patterns of hydrophobic molecules on soil and sediments (Huang et al., 2003). The Freundlich
model is the most commonly used model, due to the fact that it provides information about the
heterogeneity of a given sorbents sorptivity. Fig. 2.13 shows three types of Freundlich isotherm
linearity parameters (n < 1, n = 1, n > 1), where "n" is an indicator of the site energy
heterogeneity. A value of n < 1, a concave down shape curve, means the presence of specific
sorption sites, which limits sorption as the solute concentration increases. The case n = 1
indicates that the sorptive site is homogeneous, thus sorption is a linear function of increasing
solute concentration in the aqueous phase. The case n > 1, a concave up shape curve, suggests
that the sorptivity of the solid phase increases as the sorbate concentration increases.

45

Fig. 2.13. Three types of observed relationship between concentration of a chemical in the sorbed state,
Cs, and the dissolved state, Cw. All can be fit with a relationship of the form Cs = K Cwn where K and n
are constants (adopted from Schwarzenbach et al., 2003).

Due to high hydrophobicity, concentration of PAHs in the water is limited, and


determined by the affinity of particulate organic matter as a geosorbent (e.g. Weber et al.,
1991). It is widely accepted that within very low concentration (ppb or less), distribution of
PAHs between suspended particulate matter and water solution can be estimated from linear
sorption isotherms (Hwang et al., 2006). Equilibrium partition of PAHs onto the SPM is
actually determined by complex molecular interactions that determine the sorption of PAHs
onto the organic phase (Schlautmann & Morgan 1993; Christl & Kretzschmar, 2001; Weber et
al., 1991; Goss & Schwarzenbach, 2001). Therefore, to understand the extent of PAH
distribution affected by the particulate organic matter of given phase, the partition coefficient
(Kd = Cs / Cw in mg/L) is often normalized to the particulate organic carbon content (foc)
according to Koc = Kd / foc . Koc is so-called the organic carbon normalized partition coefficient.
Variation of Koc suggests a difference in the relative affinity of PAH for the SPM in water
systems.

46

III.

METHODS OF ANALYSIS

3.1.

Introduction
This analytical method was intended to determine the 16 PAHs (see Chapter 1.1. for the

compounds) in three environmental compartments: sediment size-fractions, suspended


particulate matter (SPM) and water solution as dissolved PAHs, using wide-pore octadecyl (C18
) reverse phase column with high performance liquid chromatography (RP-C18-HPLC) detected
by ultraviolet (UV) and programmable fluorescence detectors (FLD). The use of HPLC
UV/FLD has been well-recognized for PAHs instead of gas chromatography mass spectrometer
(GC/MS) (e.g. Grope, 2001; ICES, 1997; Wise et al., 1990). Even, EPA Method 610 suggests
that HPLC is the best used for the 16 PAHs of the EPA priority list as GC/MS can hardly
resolve four pair of isomer compounds: Anthracene and phenanthrene; chrysene and
benzo(a)anthracene; benzo(b)fluoranthene and benzo(k)fluoranthene; dibenzo(a,h)anthracene
and indeno(1,2,3-cd)pyrene.
The methods involved (1) sample collection and treatments including sediment grain-size
fractionation; (3) extraction with organic solvent percolation for sediment and SPM, and solid
phase extraction (SPE) for dissolved PAHs; (4) work-up procedures including clean-up,
concentration and matrix-exchange; and finally (5) determination of PAHs using HPLC
UV/FLD. The method performance was subject to quality control procedures using spiked
perdeuterated PAHs, procedural blank and spiked reference standards.

3.2.

Sample Collection and Treatments

3.2.1.

Surface Sediment and Size Fractionation


Surface sediment samples were collected directly using sediment grab from a small

vessel, and immediately homogenized with a stainless scoop. Any relatively large foreign
objects such as plant debris (stick or leaves), stones (rocks) or any other synthetic wastes were
discharged. The homogenized sediments were then placed in pre-combusted (heated at 250oC
for 5 hours) aluminum jars, and closed to avoid any possible contamination. The samples were
kept cool (ca. 4oC) during transportation to the laboratory, where they were then frozen at -20oC
until further analysis. Due to PAHs undergo photo-degradation under relatively adequate
oxidant and ultraviolet (UV) radiation (e.g. Sebat et al., 2001), direct exposure to sunlight and
possible strong UV sources were avoided during handle, storage and transportation.
Sediment size fractionation were carried out by wet sieving to render two general
fractions according to the Udden-Wentworth Scale: (1) the coarse (or so-called sand) fraction
with particle size diameter between 2 mm and 63 m; and (2) the fine (or mud) fraction with
particle size of < 63 m. The fine particle was obtained by centrifuging the remaining material

47

passing through the 63-m sieve at 1500 rpm for 20 minutes. Those sediment fractions were
then stored frozen until analysis.
3.2.2.

Suspended Particulate Matter (SPM)


Surface waters for SPM and dissolved PAH samples were collected with the Niskin

bottle, and placed in a pre-cleaned glass bottle, sealed, and stored until further treatment in the
laboratory. Surface water was defined as that of the surface and down up to 1 meter depth. Up
to 5-L water was sampled from the river and the estuary, and up to 10-L water was sampled
from the coast. Water filtration was undertaken on-board and in the laboratory.
SPM was sampled by filtering the samples through pre-combusted (at 400 oC for 4 hours)
Whatman GF/F (0,7 m in diameter) in triple sets of cleaned 250-ml glass containers. A
stainless-steel filtration unit with a vacuum pump was used for the collection. Prior to filtration,
the water was shaken to homogenize the particle concentration. Up to 5-L water samples was
filtered for river and estuarine water samples. Due to high density of particulate matter
particularly from black water Siak and its estuary, more than one filter was therefore employed
for one station. However, the use of filter in parallel was confirmed with blank filter values. For
coastal water, the volume of filtration was increased up to 10 L depending on the amount of
suspended particles obtained.
The filters were then removed from the unit, wrapped into aluminum foil, and stored
frozen at ca. -4oC (freezer temperature) to avoid microbial degradation until further analysis.
The filtrate (water passing through the filter) was collected for dissolved PAH sample. The
volume of filtered water was recorded in order to calculate the concentration of the suspended
matter in a given volume of water. However, the content of PAHs was calculated on a dryweight basis.
3.2.3.

Solid phase extraction (SPE) for pre-concentration of dissolved PAHs


Solid phase extraction (SPE) is one of high-performance methods in a sample preparation

for extracting and separating target components from liquid samples including PAH, which has
been increasingly used for environmental analysis (e.g. Burkhardt et al., 2005; Urbe & Ruana,
1997). In comparison to classical liquid-liquid extraction (LLE), in modern analytics SPE
provides several pre-concentration advantages with respects to low solvent consumption, time
saving, high selectivity due to various phase available, high reproducibility and recovery, as
well as automatization (e.g. Rossi & Zhang, 2000; Sargenti & McNair, 1998). In general, SPE
works as follows: (1) the target analytes in the given aqueous samples will retain in the selected
sorbent, (2) interfering (undesired) components will be phased out, and (3) the retained target
analytes will be then discharged from the sorbent through an elution with appropriate solvent.

48

The molecular interaction between the PAHs as target analytes and the sorbent
stationary phase of the SPE has been generally acknowledged as non-polar interactions taking
place between alkyl group of the sorbent and the analytes. The most typical alkyl groups
developed for the non-polar interaction are C18- or C8- (e.g. Macherey-Nagel GmbH, Duren,
Germany). However, environmental sample matrix, in particular humic substances, also plays a
significant role in determining the reproducibility and recovery of the analytes. Jeanneau et al
(2007) observed negative effects of humic substances in the natural water sample which is
virtually a clogging phenomenon that induces a competition between organic macromolecules
and organic micropollutant for the available sorption site. Therefore, for rich-humic substance
water like the Siak water sample (Baum et al., 2007) we applied a specific dual phase
aminopropyl and octadecyl- (NH2/C18) - cartridge for the extraction of PAHs. The polar group
of NH2 -modified silica is intended to provide surface interaction with polar compounds of the
complex macromolecules of humic substances. The C18 group is then to retain PAHs. SPE
Cartridge used in this study was Chromabond NH2/C18 (6 ml, 500 mg/1000 mg), obtained
from Macherey-Nagel GmbH (Dueren, Germany). The extraction was undertaken using a solid
phase extraction vacuum manifolds (VisiprepTM, SUPELCO).
Prior to extraction, the cartridges were sequentially conditioned with 10 ml hexane, 20 ml
methanol, and 20 ml Milli-Q water. To run the cartridge dried during conditioning process was
avoided. Slight film of water was left above the phase, and the cartridge was ready to be
aspirated. 500 ml to 1 L filtrated water sample was prepared. Methanol was added 2-5% of the
filtrates volume. Adding organic solvents (as a surfactant) such as methanol or 2-propanol to
the sample prior to extraction would partially overcome the clogging problem, as well as
increase the solubility of PAH (Marce & Borrull, 2000). Surrogate deuterated standards (d10fluoranthene and d12-perylene) were spiked to accomplish procedural efficiency and data
correction from any losses during extraction and work-up procedure. Then, the prepared water
sample was aspirated through the cartridge at flow rate of ca. 5 ml/min. The flow rate was
controlled by adjusting the pressure of the manifolds. After all of the filtrate was aspirated, the
cartridge was then again washed with Milli-Q water. Finally slight film (ca. 0.5 ml) of the
remaining water was left on the top of the cartridge. Then, the column was firmly topped with
foil, sealed with a paraffin film, and stored cool until analysis. Sample was collected in
duplicate.

49

3.3.

Determination of Polycyclic Aromatic Hydrocarbons using High


Performance Liquid Chromatograph coupled with Ultraviolet and
Fluorescence Detectors (HPLC UV/FLD)

3.3.1.

Soxhlet Extraction of sediment and SPM


The method was intended to do extraction for wet/moist sediments and dried-filters with

solvent percolation technique. It is based on the intimate contact between organic solvents and
the particles of sediment or suspended particulate matter, which facilitates desorption of PAHs
from particles and diffusion into the organic solvents. Sediment size-fractions were extracted in
two cycles using a combination of two systems of solvent: (1) acetone (water-miscible, polar) to
extract the water and PAHs from the sediment; and (2) less polar mixture of acetone/hexane
(1/9 ; v/v) to complete the desorption of PAHs out of the particles. Although different organic
solvents have their potential capacity to extract PAHs, these particular solvents have been
widely proposed (e.g. Grope, 2001; ICES method, 1997). The efficiency of extraction systems
as well as the method procedures were tested using spiked surrogate standards (see section 3.4.).

Sediment Extraction
Up to 10 g of the sediment fractions were prepared in the pre-cleanned (acetone/hexane,
1/1 : v/v at 120oC for 6 hours) cellulose extraction thimbles (25 x 60 mm obtained from
Whatmann), and spiked with three surrogate perdeuterated PAH standards: d10-phenanthrene,
d10-fluoranthene and d12-perylene. Then, the sediment was plugged with pre-cleaned glass
cotton to avoid spill out. The extraction was performed by means of SoxTec HT6 (Soxhletmodified extractor) at effective heating temperature of 120oC 140oC for about 4 6 hours.
The use of SoxTec extractor is beneficial for time and solvent consumption since it can perform
6 different extractions in parallel and require less solvent amount (ca. 50 ml) compared to a
traditional Soxhlet extractor (> 100 ml). Operationally speaking, before the heating-up the filled
thimbles were attached to the condenser at the Rinsing position where the thimbles hang. The
solvent container were then inserted into the extractor, and tightly clamped into the condenser.
After making sure all the connections were tight, the thimbles were moved into the Boiling
position where the thimble immersed in and the sediment came in contact with the solvents.
The extraction time was allocated for maximum of 6 hours (boiling position). At the first cycle,
extraction with acetone was performed for ca. 3 hours. Then, the thimbles were disconnected
with the solvent (at rinsing position). The solvent cups were gently removed from the condenser
and replaced with other solvent system (i.e. hexane/acetone). The second extraction was
performed for another 3 hours following the procedure as the first solvent system. Thereafter,
the extracts were combined, and proceed into the next step of working-up procedures (section
3.3.2). To check nothing is left behind after that given extraction time, the extracted sediment

50

was experimentally extracted with aceton/hexane (1/9) for 4 hours. The results confirmed no
peaks.
The result is presented on a dry-weight basis (weight/weight). Therefore other portion
(aliquot) of each fraction was prepared for dry/wet coefficient for moist-sediment weight
correction at the same time as for analytical ones. The sample aliquots were dried in the oven at
60 80oC until dried condition achieved (no change in weight after 3 minute on the balance).
Sediment dry-weight (d.w.) was corrected by the ratio of dry/wet weight of each aliquot. Water
portion in sediment ranged between ca. 1% - 3%.
SPM Extraction
Unlike sediment, extraction system for SPM-borne PAHs was using only a mixture of
acetone/hexane (1/9 ; v/v). The filter was air-dried in a clean-fumed cupboard for 24 hours prior
to extraction. The extraction efficiency was confirmed with the surrogate standards. The
accuracy of extraction system was tested with PAH standards spiking method to the blank filter
(see section 3.4.). The extraction procedure for the filter was principally similar to those of
sediment. The extraction time was ca. 6 hours at boiling position.
3.3.2.

Extract Working-Up
Following extraction, the extract working-up procedure includes concentration (or

volume reduction), clean-up from polar co-extracting compounds, and sample matrix exchange
to acetonitrile forHPLC analysis.
3.3.2.1. Concentration using rotary evaporation
Concentration (or volume reduction) was undertaken two times using rotary evaporator
(ROTAVAPOR-M, Bchi). The first concentration was carried out after extraction, and the
second one applied after column chromatography clean-up process. Instead of the PAH analytes
and other co-extracting compounds, the first extract from sediment contained a non-azeotrope
mixture of acetone and water solvents. The second extract has an azeotrope mixture of
acetone/hexane. A mixture of acetone (boiling point, BP, 56.2oC) and hexane (BP, 68.8oC)
forms an azeotrope mixture with boiling point at 49.8oC with 68% / 32% acetone/hexane by
weight in its vapor. The extracts were at the end combined. The concentration was carried out at
the cooling water rate ca. 1.5 liter/minute for two steps. First, the evaporation was undertaken
at a combination of pressure (400 500 mbar under vacuum pump) and water bath temperature
(40oC). Second, the evaporation was accomplished by applying the pressure and bath
temperature of 500-600 mbar and 45o 60oC. The evaporation ended with hexane (ICES, 1997).
The final volume attained ca. 1 - 2 ml.
The second evaporation was applied for a non-azeotrope mixture of (3/7 : v/v)
dichloromethane (BP, 40oC) and hexane after cleaning-up process. The evaporation was carried

51

out for two steps: (1) evaporation for dichloromethane with the pressure of 250 400 mbar and
the bath temperature of 30 - 35oC, and (2) evaporation of hexane with pressure of 500 600
mbar and the bath temperature of 50-60oC. The final volume attained ca. 1 ml. The loss of
analytes during concentration was confirmed with the surrogate standard for procedural
efficiency (section 3.4.).
3.3.2.2. Drying and Alumina/Silica (Al2O3 / SiO2) column for clean-up of polar mixtures
Following the first concentration process, the extracts were subject to clean-up process
using an Al2O3 / SiO2 column. This column was tapped with a drying agent of sodium sulfate
(Na2SO4) due to co-extracted water. Prior to use, anhydrous granular Na2SO4 was heated at
250oC for overnight, and stored at 150oC until use. One to three grams of Na2SO4 was placed on
the top of aluminum oxide phase in a column used for clean-up.
Clean-up procedure becomes a critical part in the sample work-up as any other coextracted compounds (interferences) may render difficulties in identification and lead to error in
quantification. Surface sediment and suspended particulate matter may contain a great variety of
polar and non-polar compounds which is inevitably co-extracted such as sulfur-containing
compounds and pigments (ICES, 1997; Wise et al., 1995). A combination of 1:2 (w/w)
deactivated (10%) Al2O3 / (3%) SiO2 column was used to clean-up the polar extract (ICES,
1997; Smedes & de Boer, 1997). Solvents used to eluate the extract were hexane and a mixture
of (3/7 : v/v) dichloromethane/hexane. A test fractionation was carried out by spiking 16 PAHs
standard in hexane matrix. The elution was divided by four fractions. Each of them was eluted
with 20 ml. The first fraction was eluted with hexane; the second and third fractions were eluted
by dichloromethane/hexane (3/7 : v/v). The fourth fraction was again eluted by hexane. The
PAHs were mainly detected in the second fraction, and slightly in the third fraction. The cleanup column was checked first with blank run, and with spiked 16 PAHs mixture standard. One
column was prepared for one batch of extraction which contained 6 samples, and conditioned
between the samples by dichloromethane/hexane as much as two times of the columns volume
(ca. 40 ml).
3.3.2.3. Nitogen (N2) stream and Matrix Exchange
The clean extract was again gently reduced to ca. 1 ml by the evaporator (see 3.3.2.1).
Further concentration was carried out using a gentle stream of N2 to dryness. Finally, 1000 l of
acetonitrile was added to the sample for HPLC analysis.
3.3.3.

Elution of the SPE Cartridges for dissolved PAHs


Dissolved PAHs retained in SPE cartridges were eluted with 4 ml of dicloromethane

(DCM) (repeated three times) at flow rate of 5 ml/min. Prior to elution, SPE cartridge was dried
under vacuum and gently stream of N2. The eluates were reduced to a dry state, then 300 400L of acetonitrile was added to them for HPLC analysis.

52

3.3.4.

PAH determination: High Performance Liquid Chromatography with ultraviolet


and fluorescence detectors (HPLC UV/FLDs)

3.3.4.1. Baseline separation with reverse phase octadecyl (RP-C18) column chromatography
Baseline separation and quantification of the PAHs were performed using a reverse phase
wide-pore octadecyl -RP-C18- column (250 mm x 46 mm, 5 m, 300 ) obtained from
BAKERBOND, J.T. Baker Inc (Phillipsburg, New Jersey, USA) with a high performance
solvent delivery pump (LKB Bromma 2249 Gradient Pump) for HPLC. The sample injection
loop was 20 l. The mobile phase or elution system was a mixture of water (Milli-Q) and
acetonitrile (CH3CN, HPLC grade), and set in a normal gradient elution mode. The normal
gradient mode means that solvent A (water), B (acetonitrile, ACN) and C (any others if
necessary) can be mixed at any ratio giving the sum of percentage equal to 100% (%A + %B +
%C = 100%). The elution program was a combination of isocratic elution and binary gradient of
water (A) and acetonitrile (B) that set off from 55% to 100% ACN with 1 ml/min flow rate for
45 minutes plus 5 minutes to conditioning (equilibrium) for the next run (Table 3.1.). Optimum
separation was exercised for PAH reference standard containing a mixture of the same 16 PAHs
compounds (16 PAH Mix 61), as well as for perdeuterated PAH standards. All reference
standards were obtained from Dr. Ehrenstorfer GmbH, Augsburg, Germany. Instead of
acenaphthene and fluorene which were not baselinely resolved, all the remaining target PAH
compounds were completely separated (Fig. 3.1). All HPLC run was performed at room
temperature (mean standard deviation = 24oC 1.1oC, N = 49, measured from June to Sept
08).
Table 3.1. Elution Mode (flow rate: 1 ml/min.)
Time Program
(minute)
00.0 10.0
10.0 30.0
30.0 40.0
40.0 45.0
45.0 50.0

A
Water (%)
45
45 0
0
0 45
45

B
Acetonitrile (%)
55
55 - 100
100
100 - 55
55

Remarks
Isocratic (at start time)
linear gradient (2.25% ACN/min.)
isocratic
linear gradient (9% ACN/min.)
isocratic (at stop time)

3.3.4.2. Detection Systems


Two different detectors were operated: (1) an ultraviolet (UVD) detector (2151 Variable
Wavelength Monitor, LKB Bromma) with a wavelength of 254 nm for acenaphthylene, and (2)
a Hawlett-Packard 1046A programmable fluorescence detector (FLD) for the other 15 PAH
compounds with three shifts of excitation/emission ( ex/ em) wavelengths (Table 3.2.). The
elution order of the compounds was confirmed by injected individual PAH reference standard.
Electronic features of the HP 1046A FLD were set to get optimum detection including the peak

53

amplification factor, PMTGAIN = 11; the flash frequency, LAMP = 1, 1.25 W / 55 Hz; and data
reduction interval, RESPONSETIME = 4, 1000 msec.

Table 3.2. Selective pairs of fluorometric wavelength and retention time (minute) of the PAHs
and the surrogate standards with the given elution mode.
No.

Analyte


Naphthalene, NAPH
Acenaphthylene, ACYN
Acenaphthene, ACEN
Fluorene, FLU
Phenanthrene, PHEN
Anthracene, ANTH
Fluoranthene, FLA
Pyrene, PYR
Benzo(a)anthracene, BaA
Chrysene, CHRY
Benzo(b)fluoranthene, BbFLA
Benzo(k)fluoranthene, BkFLA
Benzo(a)pyrene, BaP
Dibenzo(a,h)anthracene, DANTH
Benzo(g,h,i)perylene, BPERY
Indeno(1,2,3-c,d)pyrene, IPYR

250 / 341
250 / 341
250 / 341
254 / 400
254 / 400
254 / 400
254 / 400
254 / 400
254 / 400
254 / 400
254 / 400
254 / 400
254 / 400
246 / 488

Retention Time (minute)


UVD 254 nm (N=17)
FLD (N=31)
Mean SD
RSD
Mean SD
RSD
(%)
(%)


7,72 0,04
0,49
7,88 0,03
0,41
9,09 0,05
0,52
11,7 0,07
0,63
11,8 0,06
0,48
12,4 0,09
0,72
12,6 0,07
0,55
14,9 0,12
0,79
15,2 0,10
0,66
17,4 0,12
0,68
17,6 0,10
0,58
19,7 0,11
0,57
19,9 0,10
0,48
21,0 0,11
0,52
21,2 0,10
0,48
25,7 0,12
0,48
25,8 0,13
0,49
26,6 0,14
0,52
26,7 0,14
0,52
29,6 0,15
0,50
29,8 0,14
0,48
30,7 0,13
0,62
31,3 0,16
0,51
31,2 0,17
0,53
32,3 0,16
0,49
34,5 0,19
0,56
34,6 0,18
0,53
35,0 0,18
0,52
35,1 0,17
0,48
36,1 0,23
0,62
36,3 0,19
0,54

Surrogate Standards
d10-phenanthrene
d10-fluoranthene
6methylchrysene
d12-perylene

250 / 341
254 / 400
254 / 400
254 / 400

N = 10
13,6 0,23
18,9 0,15
27,6 0,15
28,9 0,20

Target Analytes

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

Wavelength
ex / em
nm
250 / 341

1,73
0,78
0,54
0,68

N = 10
13,7 0,24
19,1 0,15
27,6 0,15
29,0 0,19

1,76
0,80
0,54
0,67

54

100

100

Elution program: from 55% ACN to 100% ACN

55
5 min

10

15

20

25

30

35

40

BaP

250/341

254/400

IPYR

BPERY
DANTH

PERY D12

6 Methylchrysene

CHRY
FLA

FLA D10

PYR

PHEN

PHEN D10

ACYN

FLU

NAPH

ACEN

BaA

BbFLA

ANTH

BkFLA

246/488

Fig. 3.1. Comparison of chromatograms for 16 PAHs between UV 254 (red) and FLD (black) with the
deuterated PAHs (blue).

3.3.4.3. Identification and Quantification of PAHs


Retention time (RT) of the peaks of the 16 PAHs of the reference standard was used for
analytes peak identification. The standard was injected at the beginning of daily measurement.
The precision (confirmed by standard deviation) of the RT of the 16 PAH standards throughout
the analyses can be seen in the table 3.2. In general, the precision of the retention time between
0.03 and 0.2 minutes. The surrogate standards were also used for confirmation of the shift in the
retention time since PHEN D10, FLA D10, PERY D12 coming out just before PHEN, FLA,
and BbFLA (Fig. 3.2).
Quantification was made by integrating the peak area of the analytes using an external
standard method with five to six point calibrations. The repeatability of the peak area of each
analyte measured by relative standard deviation of repeated measurement of the working
reference standard extends from 3 to 10% (Table 3.3.). The linearity range of detector response
was exercised in the range of 0.0025 ng/l to 1 ng/l for FLD, and from 0.2 ng/l to 2 ng/l for
UVD 254 nm. The range of linearity follows the range of concentration of linearity suggested
by the work of Grope (2001). Dilution system of the 16 PAHs Mix61 standards was given in the

55

Appendix 3.1. Limit of detection (LoD) are provided in the table 3.4. It was obtained from

250/341

254/400

BPERY
IPYR

PERY D12
BbFLA

BaA
CHRY

FLA D10
FLA
PYR

ANTH

PHEN D10
PHEN

ACEN
FLU

NAPH

BkFLA

DANTH

6 Methylchrysene

BaP

repeated analysis of a sample with a very low content of the analytes (after Huber, 2003).

246/488

Fig. 3.2. An example of peaks identification using retention time of the standards (black) and
confirmation on the shift in the retention time using the deuterated standards. The chromatograms were
derived from fluorescence detector. The red chromatogram was an extract of mud fraction of sediment at
S142 Siak estuary.

Table 3.3. Repeatability of the peak area of the PAHs measured for the reference standards
(Cal#6, see Appendix 3.1.).
No.

PAHs

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene

Peak Area (N = 10)


Mean
SD
RSD (%)
74653
21481
118573
75650
58494
232070
8006
38689
114742
138477
279267
299455
54373
116668
38410
118573

2454
1095
10219
3083
4960
19156
409
2215
6803
11213
15598
15958
5374
8261
3581
10219

3,29
5,10
8,62
4,08
8,48
8,25
5,10
5,73
5,93
8,10
5,59
5,33
9,88
7,08
9,32
8,62

56

Table 3.4. Limit of detection (LoD) for the PAHs

3.4.

Conc. of
LoD (ng/l)

No.

PAHs

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene

0,01
0,08
0,02
0,01
0,01
0,001
0,02
0,002
0,001
0,001
0,01
0,001
0,001
0,003
0,004
0,002

Absolute
amount
(pg)
266
1569
441
154
115
16
453
45
25
16
120
20
14
63
79
35

Quality Controls
These quality control of method procedures involved measurements of routine procedural

blank, procedural efficiency, and instrumental performance. The procedural blank was
performed for every two batches of the sample extraction. Procedural or method efficiency for
the sample analysis was evaluated by the recovery of surrogate deuterated PAHs: d10phenanthrene, d10-fluoranthene, and d12-perylene. Those deuterated PAH standards were
respectively surrogating 2- and 3-ring PAHs, 4-ring PAHs, and 5- and 6-ring PAHs. The
summary of the recovery of each surrogate was given in the table 3.5. One working standard (16
PAH Mix61) was injected at the beginning of daily measurement for the instrumental
performance check i.e. precision, repeatability, and random errors due to changes in the
equipment conditions during the period of study.
Table 3.5. Recovery of spiked deuterated PAHs to the samples for procedural efficiency and
reproducibity. Data were driven from ultraviolet detectors (see Appendix 3.2 for detail).
Compounds

Sediment
N = 51
Mean SD

SPM
N = 32
Mean SD

Water
N = 10
Mean SD

Phenanthrene D10
Fluoranthene D10
Perylene D12

82.8 15.2
97.5 14.6
94.6 12.6

95.2 10.1
96.8 10.2
95.7 8.90

91.76.96
88.89.88

57

Procedural efficiency was also checked with 16-PAH standard spiked method. 16 PAHs
were spiked into the thimble blank for sediment and SPM, and Milli-Q water for dissolved
PAHs. These spiked standards then proceeded through the step of procedural analysis. For the
dissolved PAHs, the standard was spiked into 1000 ml of Milli-Q water plus 1% Methanol
(solubility enhancement). The spiked waters were then aspirated through the SPE cartridges,
and performed the step of analysis. The results were shown in the table 3.6 and 3.7. The
recovery of spiked standards ranged from 63% (BaP) to 105% (ANTH) for the thimbles, and
from 49% (BaP) to 96% (PHEN) for the water.
Table 3.6. Recovery of spiked PAH standards (16 PAH Mix 61) into the blank thimbles for
sediment and SPM analysis. The data was calculated from the peak area of the ultraviolet
detector response.
Compounds
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene

Spiked Mass
(ng)
500
1000
500
100
50
50
100
50
50
50
100
50
50
100
100
50

1
111
101
108
98
93
96
87
76
85
93
93
114
35
93
94
96

Recovery Test (%)


2
3
89
101
109
99
96
83
94
89
98
89
119
101
98
94
138
95
107
82
101
85
95
93
91
91
49
104
90
100
83
119
94
123

Mean
100
103
96
94
93
105
93
103
91
93
94
99
63
94
98
104

Table 3.7. Recovery of spiked PAH standards into the blank Milli-Q water for dissolved PAH
analysis. The data was calculated from the peak area of the ultraviolet detector response.
Compounds
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene

Spiked Mass
(ng)
30000
60000
30000
6000
3000
3000
6000
3000
3000
3000
6000
3000
3000
6000
6000
3000

1
47
92
84
102
98
108
97
93
101
94
102
74
58
106
104
100

Recovery Test (%)


2
3
82
88
96
99
91
95
100
103
93
97
106
112
86
95
83
93
68
75
64
66
63
66
44
48
42
46
53
53
52
53
56
59

Mean
72
95
90
101
96
109
93
90
81
75
77
55
49
71
70
72

58

References (Chapter I III)


Achten, C., Hofmann, T., 2009. Review: Native polycyclic aromatic hydrocarbons (PAH) in coals a
hardly recognized source of environmental contamination. Science of the Total Environment 407,
2461-2473.
Ahrens, M.J., Depree, C.V., 2004. Inhomogenous distribution of polycyclic aromatic hydrocarbons in
different size and density fractions of contaminated sediment from Auckland Harbour, New
Zealand: an opportunity for mitigation. Marine Pollution Bulletin, 48, 341-350.
Appel, J., Bockhorn, H., Frenklach, M., 2000. Kinetic modelling of soot formation with detailed
chemistry and physics: laminar premixed flames of C2 hydrocarbons. Combustion and Flame, 121,
122-136.
Atkinson, R., Arey, J., Zielinska, B., Pitts, Jr. J.N., Winer, A.M., 1987. Evidence for the transformation of
polycyclic organic matter in the atmosphere. Atmospheric Environment, 21(10), 2261-2264.
ATSDR, Chemical and Physical Information, in: Toxicological Profile for Polycyclic Aromatic
Hydrocarbons (PAHs), ATSDR, Atlanta, Georgia, USA, 1995, pp. 209-221
(http://www.atsdr.cdc.gov/toxprofiles/tp69-c3.pdf).
Baars, B-J., 2002. The wreckage of the oil tanker Erika human health risk assessment of beach
cleaning, sunbathing and swimming. Toxicology Letters, 128, 55-68.
Barbour, E.K., Sabra, A.H., Shaib, H.A., Berckley, A.M., Farajalla, N.S., Zurayk, R.A., Kassaify, Z.G.,
2008. Baseline data of polycyclic aromatic hydrocarbons correlation to size of marine organisms
harvested from a war-induced oil spill zone of the Eastern Mediterranean Sea. Marine Pollution
Bulletin 56, 770-797.
Baum, A., Rixen, T., Samiaji, J., 2007. Relevance of peat draining rivers in central Sumatra for the
riverine input of dissolved organic carbon into the ocean. Estuarine, Coastal and Shelf Science 73,
563 570.
Baumard, P., Budzinski, H., Garrigues, P., 1998. Polycyclic aromatic hydrocarbons in sediments and
mussels of the western Mediterranean sea. Environmental Toxicology and Chemistry 17(5), 765776.
Baumard, P., Budzinski, H., Garrigues, P., Dizer, H., Hansen, P.D., 1999a. Polycyclic aromatic
hydrocarbons in recent sediments and mussels (Mytilus edulis) from the Western Baltic Sea:
occurrence, bioavailability and seasonal variations. Marine Environmental Research, 47, 17-47.
Baumard, P., Budzinski, H., Garrigues, P., Narbonne, J.F., Burgeot, T., Michel, X., Bellocq, J., 1999b.
Polycyclic aromatic hydrocarbon (PAH) burden of mussels (Mytilus sp.) in different marine
environments in relation with sediment PAH contamination, and bioavailability. Marine
Environmental Research, 47: 415-439.
Behymer, T.D., Hites, R.A., 1985. Photolysis of polycyclic aromatic hydrocarbons adsorbed on simulated
atmospheric particulates. Environmental Science and Technology, 19, 1004-1006.
Berner, R.A., 1980. Early Diagenesis: A Theoritical Approach, 241 pp. Princeton University Press.
Bihari, N., Fafandel, M., Hamer, B., Kralj-Bilen, B., 2006. PAH content, toxicity and genotoxiciy of
coastal marine sediments from the Rovinj area, Northern Adriatic, Croatia. Science of the Total
Environment, 366, 602-611.
Binelli, A., Provini, A., 2004. Risk for human health of some POPs due to fish from Lake Iseo.
Ecotoxicology and Environmental Safety, 58, 139-145.
Bockhorn, Henning, 1994. Soot formation in combustion: mechanisms and models. Springer series in
chemical physics, 59. 596 pp.
Boehm, P.D., Douglas, G.S., Burns, W.A., Mankiewicz, P.J., Page, D.S., Bence, A.E., 1997. Application
of petroleum hydrocarbons chemical fingerprinting and allocation techniques after the Exxon
Valdez oil spill. Marine Pollution Bulletin, 34(8), 599-613.
Bofetta, P., Jourenkova, N., Gustavsson, P., 1997. Cancer risk from occupational and environmental
exposure to polycyclic aromatic hydrocarbons. Cancer Causes and Control, 8: 444-472.
Boldrin, A., Langone, L., Miserocchi, S., Turchetto, M., Acri, F., 2005. Po River plume on the Adriatic
continental shelf: dispersion and sedimentation of dissolved and suspended matter during different
river discharges rates. Marine Geology 222-223, 135-158.
BPDAS Indragiri Rokan., 2004. Physical condition of Siak River water (in Indonesian language). Paper
presented in the 1st SPICE Workshop Cluster 3.1., Pekanbaru, Indonesia.
Brasseur, C., Widart, S., Muller, M., Maghuin-Ragister, G., Scippo, M-L., 2007. Alteration of the
estrogen hormone pathway in hepatic cells after exposure to polycyclic aromatic hydrocarbons.
Toxicology Letters 172(1), S38.
Budzinski, H., Jones, I., Bellocq, J., Pierard, C., Garriques, P., 1997. Evaluation of sediment contaminant
by polycyclic aromatic hydrocarbons in the Gironde estuary. Marine Chemistry 58, 85-97.

59

Burkhardt, M.R., Zaugg, S.D., Burbank, T.L., Olson, M.C., Iverson, J.L., 2005. Pressurized liquid
extraction using water/isopropanol coupled with solid-phase extraction cleanup for semivolatile
organic compounds, polycyclic aromatic hydrocarbons (PAH), and alkylated PAH homolog groups
in sediment. Analytica Chimica Acta 549, 104-116.
Butler, J.D., Crossley, P., 1981. Reactivity of polycyclic aromatic hydrocarbons adsorbed on soot
particles. Atmospheric Environment, 15, 91-94.
Cachot, J., Geffard, O., Augagneur, S., Lacroix, S., Le Menach, K., Peluhet, L., Couteau, J., Denier, X.,
Devier, M.H., Pottier, D., Budzinski, H., 2006. Evidence of genotoxicity related to high PAH
content of sediments in the upper part of Seine estuary (Normandy, France). Aquatic Toxicology
79, 257-267.
Cao, Z., Wang, Y., Ma, Y., Xu, Z., Shi, G., Zhuang, Y., Zhu, T., 2005. Occurrence and distribution of
polycyclic aromatic hydrocarbons in reclaimed water and surface water of Tianjin, China. Journal
of Hazardous Materials A122, 51-59.
Carls, M.G., Short, J.W., Payne, J., 2006. Accumulation of polycyclic aromatic hydrocarbons by
Neocalanus copepods in Port Valdez, Alaska. Marine Pollution Bulletin, 52, 1480-1489.
Chester, R., 2003. Marine Geochemistry, 2nd Edition. Blackwell Science Ltd, 506 pp.
Chiang, K-C., Chio, C-P., Chiang, Y-H., Liao, C-M., 2009. Assessing hazardous risks of human exposure
to temple airborne polycyclic aromatic hydrocarbons. Journal of Hazardous Materials 166, 676685.
Chiou, C.T., McGroddy, S.E., Kile, D.E., 1998. Partition characteristics of polycyclic aromatic
hydrocarbons on soils and sediments. Environmental Science & Technology 32, 264-269.
Christl, I., Kretzschmar, R., 2001. Relating ion binding by fulvic and humic acids to chemical
composition and molecular size. 1. Proton Binding. Environmental Science and Technology, 35,
2505-2511.
Conde, F.J., Ayala, J.H., Afonso, A.M., Ganzalez, V., 2005. Emissions of polycyclic aromatic
hydrocarbons from combustion of agriculture and sylvicultural debris. Atmospheric Environment,
39: 6654 6663.
de Abrantes, R., de Assuno, J.V., Pesquero, C.R., 2004. Emission of polycyclic aromatic hydrocarbons
from light-duty diesel vehicles exhaust. Atmospheric Environment 38, 1631-1640.
De Luca, G.D., Furesi, A., Micera, G., Panzanelli, A., Piu, P.C., Pilo, M.I., Spano, N., Sanna., 2005.
Nature, distribution and origin of polycyclic aromatic hydrocarbons (PAHs) in the sediments of
Olbia harbor (Northern Sardinia, Otaly). Marine Pollution Bulletin 50, 1223-1232.
Deng, H., Peng, P., Huang, W., Song, J., 2006. Distribution and loadings of polycyclic aromatic
hydrocarbons in the Xijiang River in Guangdong, South China. Chemosphere, 64, 1401-1411.
Dobbins, R.A., Fletcher, R.A., Benner Jr. B.A., Hoeft, S., 2006. Polycyclic aromatic hydrocarbons in
flames, in diesel fuels, and in diesel emissions. Combustion and Flame, 144, 773-781.
Fabbri, D., Vassura, I., Sun, C.-G., Snape, C.E., McRae, C., Fallick, A.E., 2003. Source apportionment of
polycyclic aromatic hydrocarbon in a coastal lagoon by molecular and isotopic characterisation.
Marine Chemistry, 84, 123-135.
Falcon, M.S.G-., Lamela, C.P-., Gandara, J.S-., 2004. Startegies for the extraction of free and bound
polycyclic aromatic hydrocarbons in runn-off water rich in organic matter. Analyt. Chim. Acta,
508: 177 183.
Fang, M., Zheng, M., Wang, F., To, K.L., Jaafar, A.B., Tong, S.L., 1999. The solvent-extractable organic
conpounds in the Indonesia biomass burning aerosols characterization studies. Atmospheric
Environment 33, 783-795.
Fernandes, M.B., Sicre, M.-A., Boireau, A., Tronczynski, J., 1997. Polyaromatic hydrocarbon (PAH)
distributions in the Seine river and its estuary. Marine Pollution Bulletin, 34(11): 857-867.
Fernandes, M.B., Sicre, M.-A., Broyelle, I., Lorre, A., Pont, D., 1999. Contamination by polycyclic
aromatic hydrocarbons (PAHs) in French and European rivers. Hydrobiologia, 410, 343-348.
Fernndez, P., Grimalt, J.O., Vilanova, R.M., 2002. Atmospheric gas-particle partitioning of polycyclic
aromatic hydrocarbons in high mountain regions of Europe. Environmental Science and
Technology, 36, 1162-1168.
Fernndez, P., Vilanova, R., Grimat, J.O., 1996. PAH distribution in sediments from high mountain lakes.
Polycyclic Aromatic Compounds 9(1), 121-128.
Fernndez-Gonzles, V., Concha-Graa, E., Muniategui-Lorenzo, S., Lpez-Maha, P., Prada-Rodrguez,
D., 2007. Solid-phase microextraction-gas chromatography-tandem mass spectrometric analysis of
polycyclic aromatic hydrocarbons towards the European Union water directive 2006/0129 EC.
Journal of Chromatography A, 1176, 48-56.
Frenklach, M., 2002. Reaction mechanism of soot formation in flames. Physical Chemistry Chemical
Physics, 4, 2028-2037.

60

Garbarini, D.R., Lion, L.W., 1986. Influence of the nature of soil organics on the sorption of toluene and
trichloroethylene. Environmental Science and Technology 20, 1263-1269.
Gauthier, T.D., Seltz, W.R., Grant, C.L., 1987. Effects of structural and compositional variations of
dissolved humic materials on pyrene Koc values. Environmental Science and Technology, 21(3),
243-248.
Gebhardt, A.C., Gaye-Haake, B., Unger, D., Lahajnar, N., Ittekkot, V., 2004. Recent particulate organic
carbon and total suspended matter fluxes from the Ob and Yenisei Rivers into the Kara Sea
(Siberia). Marine Geology 2007, 225-245.
Gevao, B., Jones, K.C., Hamilton-Taylor, J., 1988. Polycyclic aromatic hydrocarbon (PAH) deposition to
and processing in a small rural lake, Cumbria UK. The Science of the Total Environment, 215,
231-242.
Ghosh, U., Zimmerman, J.R., Luthy, R.G., 2003. PCB and PAH speciation among particle types in
contaminated harbor sediments and effects on PAH bioavailability. Environmental Science &
Technology 37, 2209-2217.
Glarborg, P., 2007. Hidden interactions - trace species governing combustion and emissions. Proceedings
of the Combustion Institute, 31, 77-98.
Golomb, D., Barry, E., Fisher, G., Varanusupakul, P., Koleda, M., Rooney, T., 2001. Atmospheric
deposition of polycyclic aromatic hydrocarbons near New England coastal waters. Atmospheric
Environment 35, 6245-6258.
Goss, K-U., Schwarzenbach, R.P., 2001. Linear free energy relationships used to evaluate equilibrium
partitioning of organic compounds. Environmental Science and Technology, 35 (1), 1-9.
Grathwohl, P., 1990. Influence of organic matter from soils and sediments from various origins on the
sorption of some chlorinated aliphatic hydrocarbons: iplication on Koc correlations. Environmental
Science and Technology 24, 1687-1693.
Grope, N., 2001. Natrliche und anthropogene organische Spurenstoffe in kstennahen
Meeressedimenten: Gehalte und Verteilung von polyaromatischen Kohlenwasserstoffen. PhD
Dissertation. Pp. 329.
Gschwend, P.M., Hites, R.A., 1981. Fluxes of the polycyclic aromatic compounds to marine and
lacustrine sediments in the northeastern United States. Geochimica et Cosmochimica Acta, 45,
2359-2367.
Guinan, J., Charlesworth, M., Service, M., Oliver, T., 2001. Source and geochemical constraints of
polycyclic aromatic hydrocarbons (PAHs) in sediments and mussels of two Northern Irish Sealoughs. Marine Pollution Bulletin 42(11), 1073-1081.
Haitzer, M., Hss, S., Traunspurger, W., Steinberg, C., 1998. Effects of dissolved organic matter (DOM)
on the bioconcentration of organic chemicals in aquatic organisms, a review. Chemosphere, 37(7),
1335-1362.
Hawthorne, S.B., Azzolina, N.A., Neuhauser, E.F., Kreitinger, J.P., 2007. Predicting bioavailability of
sediment polycyclic aromatic hydrocarbons to Hyalella asteca using equilibrium partitioning,
supercritical fluid extraction, and pore water concentration. Environmental Science and
Technology, 41, 6297-6304.
Heemken, O.P., satchel, B., Theobald, N., Wenclawiak, B.W., 2000. Temporal variability of organic
micropollutants in suspended particulate matter of the River Elbe at Hamburg and the River Mulde
at Dessau, Germany. Archieve Environmental Contamination Toxicology 38, 11-31.
Hellou, J., Steller, S., Leonard, J., Langille, M.A., Tremblay, D., 2005. Partitioning of polycyclic
aromatic hydrocarbons between water and particles compared to bioaccumulation in mussels: a
harbour case. Marine Environmental Research, 59, 101-117.
Hellou, J., Steller, S., Leonard, J., Langille, M.A., Tremblay, D., 2005. Partitioning of polycyclic
aromatic hydrocarbons between water and particles compared to bioaccumulation in mussels: a
harbour case. Marine Environmental Research, 59, 101-117.
Hites, R.A., Laflamme, R.E., Farrington, J.W., 1977. Sedimentary polycyclic aromatic hydrocarbons: The
historical record. Science, 198, 829-831.
Horowitz, A.J., Elrick, K.A., 1987. The relation of stream sediment surface area, grain size and
composition to trace element chemistry. Applied Geochemistry, 2, 437-451.
Hu, Y., Bai, Z., Zhang, L., Wang, X., Zhang, L., Qingchan, Y., Zhu, T., 2007. Health risk assessment for
traffic policemen exposed to polycyclic aromatic hydrocarbons (PAHs) in Tianjin, China. Science
of the Total Envrionment 382, 240-250.
Huang, W., Peng, P., Yu, Z., Fu, J., 2003. Review: Effects of organic matter heterogeneity on sorption
and desorption of organic contaminants by soils and sediments. Applied Geochemistry 18, 955972.

61

Huang, W., Weber, W.J., Jr. 1997. A distributed reactivity model for sorption by soils and sediments. 10.
Relationships between desorption, hysteresis, and the chemical characteristics of organic domains.
Environmental Science and Technology 31, 2562-2569.
Huber, W., 2003. Basic calculation about the limit of detection and its optimal determination.
Accreditation and Quality Assurance 8, 213-217.
Hussain, M., Rae, J., Gilman, A., Kauss, P., 1998. Lifetime health rissk assessment from exposure of
recreastional users to polycyclic aromatic hydrocarbons. Archives of Environmental
Contamination and Toxicolology 35, 527-531.
Hwang , H-. M., Foster, G.D., 2006. Characterization of polycyclic aromatic hydrocarbons in urban
stormwater runoff flowing into the tidal Anacostia River, Washington, DC, USA. Environmental
Pollution 140, 416-426.
IARC, 1987. IARC Monographs on the evaluation of carcinogenic risks to humans, Suppl. 7, overall
evaluations of carcinogenicity: an updating of IARC Monographs volumes 1 to 42, Lyon, IARC
Press.
ICES, 1997. Determination of polycyclic aromatic hydrocarbons (PAHs) in sediments: Analytical
methods. In Report of the ICES Advisory Committee on the Marine Environment, 1997. ICES
Cooperative Research Report 222, 118-124.
IPCC 2000. Land Use, Land-Use Change and Forestry, A Special Report of the IPCC. Cambridge
University Press, 377 pp (available online: http://www.ipcc.ch).
Ishaq, R., Nf, C., Zebhr, Y., Broman, D., Jrnberg, U., 2003. PCBs, PCNs, PCDD/Fs, PAHs and ClPAHs in air and water particulate samples patterns and variations. Chemosphere, 50, 1131-1150.
Iwata, H., Tanabe, S., Sakai, N., Nishimura, A., Tatsukawa, R., 1994. Geographical distribution of
persistent organochlorines in air, water and sediments from Asia and Oceania, and their
implications for global redistribution from lower latitudes. Environmental Pollution, 85, 15-33.
Jeanneau, L., Faure, P., Jard, F., 2007. Influence of natural organic matter on the solid-phase extraction
of organic micropollutants application tot he water-extract from highly contaminated river sediment.
Journal of Chromatography A, 1173, 1-9.
Jenkins, B.M., Jones, A.D., Turn, S.Q., Williams, R.B., 1996. Emission factors for polycyclic aromatic
hydrocarbons from biomass burning. Environmental Science and Technology, 30, 2462-2469.
Jiang, C., Alexander, R., Kagi, R.I., Murray, A.P., 2000. Origin of perylene in ancient sediments andtis
geological significance. Organic Geochemistry, 31, 1545-1559.
Johnson, M.D., Huang, W., Weber, W.J., Jr. 2001. A distributed reactivity model for sorption by soils and
sediments. 13. Simulated diagenesis of natural sediment organic matter and its impact on
sorption/desorption equilibria. Environmental Science and Technology 35, 1680-1687.
Juhaz, A.L., Naidu, R., 2000. Bioremediation of high molecular weight polycyclic aromatic
hydrocarbons: a review of the microbial degradation of benzo[a]pyrene. International
Biodeterioration and Biodegradation 45, 57-88.
Kakareka, S.V., Kukharchyk, T.I., 2003. PAH emission from the open burning of agricultural debris. The
Science of the Total Environmental, 308: 257-261.
Kanaly, R.A., Harayama, S., 2000. Biodegradation of high-molecular-weight polycyclic aromatic
hydrocarbons by bacteria. Journal of Bacteriology, 182(8), 2059-2067.
Karickhoff, S.W., Brown, D.S. and Scott, T.A., 1979. Sorption of hydrophobic pollutants on natural
sediments. Water Research 13, 241-248.
Kayal, S., Connel, D.W., 1995. Polycyclic aromatic hydrocarbons in biota from the Brisbane river
estuary, Australia. Estuarine, Coastal and Shelf Science 40, 475-493.
Khalfi, A., Trouve, G., Delobel, R., Delfosse, L., 2000. Correlation of CO and PAH emission during
laboratory-scale incineration of wood waste furnitures. Journal of Analytical Application
Pyrolysis, 56, 243-262.
Kim, G.B., Maruya, K.A., Lee, R.F., Lee, J.-H, Koh, C.-H., Tanabe, S., 1999. Distribution and sources of
polycyclic aromatic hydrocarbons in sediments from Kyeonggi Bay, Korea. Marine Pollution
Bulletin, 38(1), 7-15.
Ko, F-C., Baker, J.B., 1995. Partitioning of hydrophobic organic contaminants to resuspended sediment
& plankton in the mesohaline Chesapeake Bay. Marine Chemistry, 49, 171-188.
Koelmans, A.A., Jonker, M.T.O., Cornelissen, G., Bucheli, T.D., Van Noort, P.C.M., Gustafsson, .,
2006. Black carbon: the reverse of its dark side, review. Chemosphere, 63, 365-377.
Kot-Wasik, A., D
browska, D., Namie nik, J., 2004. Photodegradation and biodegradation study of
benzo(a)pyrene in different liquid media. Journal of Photochemistry and Photobiology A:
Chemistry 168, 109-115.
Kowalewska, G., Konat-Stepowicz, J., Wawrzyniak-Wydrowska, B., Szymczak-Zyla, M., 2003. Transfer
of organic contaminants to the Baltic in the Odra Estuary. Marine Pollution Bulletin, 46, 703-718.

62

Ladesma, E.B., Kalish, M.A., Nelson, P.F., Wornat, M.J., Mackie, J.C., 2000. Formation and fate of PAH
during the pyrolysis and fuel-rich combustion of coal primary tar. Fuel, 79, 1801-1814.
Laflamme, R.E., Hites, R.A., 1978. The global distribution of polycyclic aromatic hydrocarbons in recent
sediments. Geochimica et cosmochimica Acta, 42: 289-303.
Laflamme, R.E., Hites, R.A., 1979. Tetra- and pentacyclic, naturally-occurring, aromatic hydrocarbons in
recent sediments. Geochimica et Cosmochimica Acta, 43, 1687-1691.
Langmann, B., 2007. A model study of smoke-haze influence on clouds and warm precipitation formation
in Indonesia 1997/1998. Atmospheric Environment 41, 6838-6852.
Law, R.J., Dawes, V.J., Woodhead, R.J., Matthiessen, P., 1997. Polycyclic aromatic hydrocarbons (PAH)
in seawater around England and Wales. Marine Pollution Bulletin 34(5), 306 322.
Leahy, J.G., Colwell, R.R., 1990. Microbial degradation of hydrocarbons in the Environment.
Microbiological Reviews 54(3), 305-315.
Lee, R.G.M., Coleman, P., Jones, J.L., Jones, K.C., Lohmann, R., 2005. Emission factors and importance
of PCDD/Fs, PCBs, PCNs, PAHs and PM10 from the domestic burning of coal and wood in the
U.K. Environmental Science and Technology, 39, 1436-1447.
Lehto, K-M., Vourimaa, E., Lemmetyinen, H., 2000. Photolysis of polycyclic aromatic hydrocarbons
(PAHs) in dilute aqueous solutions detected by fluorescence. Journal of Photochemistry and
Photobiology A: Chemistry, 136, 53-60.
Li, C-T., Mi, H-H., Lee, W-J., You, W-C., Wang, Y-F., 1999. PAH emission from the industrial boilers.
Journal of Hazardous Materials, A69, 1-11.
Liang, Y., Tse, M.F., Young, L., Wong, M.H., 2007. Distribution patterns of polycyclic aromatic
hydrocarbons (PAHs) in the sediments and fish at Mai Po Marshes Nature Reserve, Hong Kong.
Water Research, 41, 1303-1311.
Libes, S.M., 1992. An introduction to marine biogeochemistry. John Wiley & Sons, Inc.
Lima, A.L.C., Farrington, J.W., Reddy, C.M., 2005. Combustion-derived polycyclic aromatic
hydrocarbons in the environment a Review. Environmental Forensics, 6:2, 109-131.
Lipiatou, E., Saliot, A., 1991. Fluxes and transport of anthropogenic and natural polycyclic aromatic
hydrocarbons in the western Mediterranean Sea. Marine Chemistry 32, 51-71.
Liu, H., Amy, G., 1993. Modeling partitioning and transport interactions between natural organic matter
and polynuclear aromatic hydrocarbons in groundwater. Environmental Science and Technology,
27, 1553-1582.
Liu, K., Xie, W., Zhao, Z-B., Pan, W-P., Riley, J.T., 2000. Investigation of polycyclic aromatic
hydrocarbons in fly ash from fluidized bed combustion systems. Environmental Science and
Technology 34(11), 2273-2279.
Liu, W.X., Dou, H., Wei, Z.C., Chang, B., Qiu, W.X., Liu, Y., Tao, S., 2009. Emission characteristics of
polycyclic aromatic hydrocarbons from combustion of different residential coals in North China.
Science of the Total Environment 407, 1436-1446.
Liu, Y., Tao, S., Yang, Y., Dou, H., Yang, Y., Coveney, R.M., 2007. Inhalation exposure of traffic police
officers to polycyclic aromatic hydrocarbons (PAHs) during the winter in Beijing, China. Science
of the Total Environment, 383, 98-105.
Long, E.R., MacDonald, D.D., Smith, S.L., Calder, F.D., 1995. Incidence of adverse biological effects
within ranges of chemical concentration in marine & estuarine sediments. Environmental
Management, 19, 81 97.
Luo, X-J., Chen, S-J., Mai, B-X., Yang, Q-S., Sheng G-Y., Fu, J-M., 2006. Polycyclic aromatic
hydrocarbons in suspended matter and sedimens from the pearl River Estuary and adjacent coastal
areas, China. Environmental Pollution, 139: 9 20.
Luthy, R. G., Aiken, G.R., Brusseau, M.L., Cunningham, S.D., Gschwend, P.M., Pignatello, J.J.,
Reinhard, M., Traina, S.J., 1997. Sequestration of hydrophobic organic contaminants by
geosorbents. Environmental Science and Technology, 31, 3341-3347.
Mackay, D., Fraser, A., 2000. Bioaccumulation of persistent organic chemicals: mechanisms and models.
Environmental Pollution, 110, 375-391.
Mackay, D., Powers, B., 1987. Sorption of hydrophobic chemicals from water: a hypothesis for the
mechanism of the particle concentration effect. Chemosphere, 6(4), 745-757.
Maggi, C., Onorati, F., Lamberti, C.V., Cicero, A.M., 2008. The hazardous priority substances in Italy:
National riles and environmental quality standard in marine environment. Environmental Impact
Assessment Review, 28, 1-6.
Mai, B.-X., Fu, J.-M., Sheng, G.-Y., Kang, Y.-H., Lin, Z., Zhang, G., Min, Y.-S., Zang, E.Y., 2002.
Chlorinated and polycyclic aromatic hydrocarbons in riverine and estuarine sediments from Pearl
River Delta, China. Environmental Pollution, 117, 457-474.

63

Mnnist, M.K., Melin, E.S., Puhakka, J.A., Ferguson, J.F., 1996. Biodegradation of PAH misture by a
marine sediment enrichment. Polycyclic Aromatic Compounds, 11(1), 27-34.
Marce, R.M., Borrull, F., 2000. Solid-phase extraction of polycyclic aromatic compounds. Journal of
Chromatography A, 885, 273-290.
Marr, LC., Kirchstetter, T.W., Harley, R.A., Miguel, A.H., Hering, S.V., Hammond, S.K., 1999.
Characterization of polycyclic aromatic hydrocarbons in motor vehicle fuels and exhaust
emissions. Environmental Science and Technology, 33, 3091-3099.
Mart-Cid, R., Bocio, A., Llobet, J.M., Domingo, J.L., 2007. Intake of chemical contaminants through
fish and seafood consumption by children of Catalonia, Spain: Health risks. Food and Chemical
Toxicology, 45, 1968-1974.
Maruya, K.A., Risebrough, R.W., Horne, A.J., 1996. Partitioning of polynuclear aromatic hydrocarbons
between sediments from San Fransisco Bay and their porewater. Environmental Science and
Technology, 30, 2942-2947.
Maskaoui, K., Zhou, J.L., Hong, H.S., Zhang, Z.L., 2002. Contamination by polycyclic aromatic
hydrocarbons in the Jiulong River Estuary and Western Xiamen Sean, China. Environmental
Pollution 118, 109-122.
Mastral, A.M., Calln, M., Murillo, R., Garca, T., 1988. Assessment of PAH emissions as a function of
coal combustion variables in fluidized bed. 2. Air excess percentage (short communication). Fuel
77 (13), 1513-1516.
Mazzera, D., Hayes, T., Lowenthal, D., Zielinska, B., 1999. Quantification of polycyclic aromatic
hydrocarbons in soil at McMurdo station, Antarctica. The Science of the Total Environment, 229,
65-71.
McGrath, T.E., Chan, W.G., Hajaligol, M.R:, 2003. Low temperature mechanism for the formation of
polycyclic aromatic hydrocarbons from the pyrolysis of cellulose. Journal of Analytical and
Applied Pyrolysis 66, 51-70.
Means, J.C., 1995. Influence of salinity upon sediment-water partitioning of aromatic hydrocarbons.
Marine Chemistry, 51, 3-16.
Means, J.C., Wood, S.G., Hassett, J.J., Banwart, W.L., 1980. Sorption of polycyclic aromatic
hydrocarbons by sediments and soils. Environmental Science and Technology, 14, 1542-1528.
Menon, N.N., Menon, N.R., 1999. Uptake of polycyclic aromatic hydrocarbons from suspended oil borne
sediments by the marine bivalve Sunetta scripta. Aquatic Toxicology, 45, 63-69.
Miguel, A.H., Kirchstetter, T.W., Harley, R.A., 1998. On-road emissions of particulate polycyclic
aromatic hydrocarbons and black carbon from gasoline and diesel vahicles. Environmental Science
and Technology, 32, 450-455.
Mihelcic, J.R., Luthy, R.G., 1988. Degradation of polycyclic aromatic hydrocarbon compounds under
various redox condition in soil-water systems. Applied and Environmental Microbiology, 54(5),
1182-1187.
Miller, J.S., Olejnik, D., 2001. Photolysis of polycyclic aromatic hydrocarbons in water. Water Research,
35(1), 233-243.
Milliman, J.D., Farnsworth, K.L., Albertin, C.S., 1999. Flux and fate of fluvial sediments leaving large
islands in the East Indies. Journal of Sea Research, 41, 97-107.
Mitra, S., Dellapenna, T.M., Dickhut, R.M., 1999. Polycyclic aromatic hydrocarbon distribution within
lower Hudson River estuary sediments: physical mixing vs sediment geochemistry. Estuarine,
Coastal and Shelf Science, 49, 311-326.
Mondon, J.A., Nowak, B.F., Sodergren, A., 2001. Persistent organic pollutants in oysters Crassostrea
gigas and sand flathead Platycephalus bassensis from Tasmanian estuarine and coastal waters.
Mar. Poll. Bull., 42(2), 157-161.
Neff, J.M., 1979. Polycyclic aromatic hydrocarbon in the aquatic environment: sources, fates and
biological effects. Aplied Science Publisher Ltd, Essex, UK, 262 pp.
Oanh, N.T.K., Albina, D.O., Ping, L., Wang, X., 2005. Emission of particulate and polycyclic aromatic
hydrocarbons from selected cookstove fuel systems in Asia. Biomass and Bioenergy, 28, 579590.
Oanh, N.T.K., Reutergrdh, L.B., Dung, N.T., 1999. Emission of polycyclic aromatic hydrocarbons and
particulate matter from domestic combustion of selected fuels. Environmental Science and
Technology, 33, 2703-2709.
Oen, A.M.P., Cornelissen, G., Breedveld, G.D., 2006. Relation between PAH and black carbon contents
in size fractions of Norwegian harbor sediments. Environmental Pollution 141, 370-380.
Okay, O.S., Donkin, P., Peters, L.D., Livingstone, D.R., 2000. The role of algae (Isochrysis galbana)
enrichment on the bioaccumulation of benzo[a]pyrene and its effects on the blue mussel Mytilus
edulis. Environmental Pollution, 110, 103-113.

64

Ollivon, D., Garban, B., Chesterikoff, A., 1995. Analysis of the distribution of some polycyclic aromatic
hydrocarbons in sediments and suspended matter in the river Seine (France). Water, Air and Soil
Pollution, 81, 135-152.
Oros, D.R., Abas, M.R.B., Omar, N.Y.M.J., Rahman, N.A., Simoneit, B.R.T., 2006. Identification and
emission factors of molecular tracers in organic aerosols from biomass burnings: Part 3. Grasses.
Applied Geochemistry, 21, 919-040.
Oros, D.R., Ross, J.R.M., 2005. Polycyclic aromatic hydrocarbons in bivalves from the San Francisco
estuary: Spatial distribution, temporal trends, and sources (1993-2001). Marine Environmental
Research, 60, 466-488.
Page, S.E., Siegert, F., Rieley, J.O., Boehm, H-D. V., Jaya, A., Limin, S., 2002. The amount of carbon
released from peat and forest fires in Indonesia during 1997. Nature 420, 61-65.
Parameswaran, K., Nair, S.K., Rajeev, K., 2004. Impact of Indonesian forest fires during the 1997 El
Nino on the aerosol distribution over the Indian Ocean. Advances in Space Research 33, 10981103.
PEMPROV RIAU, 2005. Kebijakan pengelolaan daerah aliran sungai Siak. Presented at Workshop for
DAS SIAK Jakarta 11 August 2005, Agency for the Assessment and Application of Technology,
Indonesia.
Prez-Cadaha, B., Laffon, B., Psaro, E., Mndez, J., 2004. Evaluation of PAH bioaccumulation and
DNA damage in mussels (Mytilus galloprovincialis) exposed to spilled Prestige crude oil.
Comparative Biochemistry and Physiology, Part C, 138, 453-460.
Poerschmann, J., Zhang, Z., Kopinke, F-D., Pawliszyn, J., 1997. Solid phase microextraction for
determining the distribution of chemicals in aqueous matrices. Analytical Chemistry, 69, 597-600.
Poeton, T.S., Stensel, H.D., Strand, S.E., 1999. Biodegradation of polyaromatic hydrocarbons by marine
bacteria: Effect of solid phase on degradation kinetics. Water Research, 33(3), 868-880.
Prahl, F.G., Carpenter, R., 1983. Polycyclic aromatic hydrocarbons (PAH) phase associations in
Washington Coastal sediments. Geochimica et Cosmochimica Acta, 47, 1013-1023.
Quah, E., 2002. Transboundary pollution in southeast Asia. World Development 30(3), 429-441.
Rappaport, S.M., Waidyanatha, S., Serdar, B., 2004. Naphthalene and its biomarkers as measures of
occupational exposure to polycyclic aromatic hydrocarbons. Journal of Environmental Monitoring,
6, 413-416.
Ravindra, K., Sokhi, R., Van Grieken, R., 2008. Atmospheric polycyclic aromatic hydrocarbons: source
attribution, emission factors and regulation. Atmospheric Environment, 42, 2895-2921.
Requejo, A.G., Sassen, R., McDonald, T., Denoux, G., Kennicutt II, M.C., Brooks, J.M., 1996.
Polynuclear aromatic hydrocarbons (PAH) as indicators of the source and maturity of marine
crude oils. Organic Geochemistry, 24(10/11), 1017-1033.
Ribeiro, C.A.O., Vollaire, Y., Sanchez-Chardi, A., Roche, H., 2005. Bioaccumulation and the effects of
organochlorine pesticides, PAH and heavy metals in the Eel (Anguilla anguilla) at the Camargue
Natura Reserve, France. Aquatic Toxicology, 74, 53-69.
Richardson, B.J., Zheng, G.J., Tse, E.S.C., De Luca-Abbott, S.B., Siu, S.Y.M., Lam, P.K.S., 2003. A
comparison of polycyclic aromatic hydrocarbon and petroleum hydrocarbon uptake by mussels
(Perna viridis) and semi-permeable membrane devices (SPMDs) in Hong Kong coastal waters.
Environmental Pollution 122: 223-227.
Richter, H., Howard, J.B., 2000. Formation of polycyclic aromatic hydrocarbons and their growth to soot
a review of chemical reaction pathways. Progress in Energy and Combustion Science, 26, 565608.
Ricking, M., Koch, M., Rotard, W., 2005. Organic pollutants in sediment cores of NE-Germany:
Comparison of the marine Arkona Basin with freshwater sediments. Marine Pollution Bulletin, 50,
1699-1705.
Riddle, S. G., Jakober, C.A., Robert, M.A., Cahill, T.M., Charles, M.J., Kleeman, M.J., 2007a. Large
PAHs detected in fine particulate matter emitted from ligh-duty gasoline vehicles. Atmospheric
Environment 41, 8658-8668.
Riddle, S. G., Robert, M.A., Jakober, C.A., Hannigan, M.P., Kleeman, M.J., 2007b. Size distribution of
trace organic species emitted from light-duty gasoline vehicles. Environmental Science and
Technology 41, 7464-7471.
Rockne, K.J., Shor, L.M., Young, L.Y., Taghon, G.L., Kosson, D.S., 2002. Distribution sequestration and
release of PAHs in weathered sediment: the role of sediment structure and organic carbon
properties. Environmental Science and Technology, 36, 2636-2644.
Rossi, D.T., Zhang, N., 2000. Automating solid-phase extraction: current aspects and future prospects.
Journal of Chromatography A, 885, 97-113.

65

Rybicki, B.A., Nock, N.L., Savera, A.T., Tang, D., Rundle, A., 2006. Polycyclic aromatic hydrocarbonDNA adduct formation in prostate carcinogenesis. Cancer Letter, 239, 157-167.
Ryder, A.G., Glynn, T.J., Feely, M., Barwise, A.J.G., 2002. Characterization of crude oils using
fluorescence lifetime data. Spectrochimica Acta Part A, 58, 1025-1037.
Sabat, J., Bayona, J.M., Solanas, A.M., 2001. Photolysis of PAHs in aqueous phase by UV irradiation.
Chemosphere 44, 119-124.
Saco-lvarez, Bellas, J., Nieto, ., Bayona, J.M., Albaiges, J., Beiras, R., 2008. Toxicity and
phototoxicity of water-accomodated fraction from Prestige fuel oil and marine fuel oil evaluated
by marine bioassays. Science of the Total Envrionment, 394, 275-282.
Samanta, S.K., Singh, O.V., Jain, R.K., 2002. Polycyclic aromatic hydrocarbons: environmental pollution
and bioremediation. TRENDS in Biotechnology, 20(6), 243-248.
Sargenti, S.R., McNair, H.M., 1998. Comparison of solid-phase extraction and supercritical fluid
extraction for extraction of polycyclic aromatic hydrocarbons from drinking water. Journal of
Microcolumn Separation, 10(1), 125-131.
Schauer, J.J., Kleeman, M.J., Cass, G.R., Simoneit, B.R.T., 2001. Measurement of emission from
pollution sources. 3. C1-C29 organic compounds from fireplace combustion of wood.
Environmental Science and Technology, 35, 1716-1728.
Schlautman, M.A., Morgan, J.J., 1993. Effects of aqueous chemistry on the binding of polycyclic
aromatic hydrocarbons by dissolved humic materials. Environmental Science and Technology, 27,
961-969.
Schwarzenbach, R.P., Escher, B.I., Fenner, K., Hofstetter, T.B., Johnson, C.A., von Gunten, U., Wehrli,
B., 2006. The Challenge of micropollutants in aquatic systems. Science, 313, 1072-1077.
Sebat, J., Bayona, J.M., Solanas, A.M., 2001. Photolysis of PAHs in aqueous phase by UV irradiation.
Chemosphere 44, 119-124.
See, S.W., Balasubramanian, R., Rianawati, E., Karthikeyan, S., Streets, D.G., 2007. Characterization and
source apportionment of particulate matter  2.5 m in Sumatra, Indonesia during a recent peat
fire episode. Environmental Science and Technology, 41, 3488-3494.
Sharma, R.K., Hajaligol, M.R., 2003. Effect of pyrolysis conditions on the formation of polycyclic
aromatic hydrocarbons (PAHs) from polyphenolic compounds. Journal of Analytical and Applied
Pyrolysis 66, 123-144.
Shemer, H., Linden, K.G., 2007. Photolysis, oxidation and subsequent toxicity of a mixture of polycyclic
aromatic hydrocarbons in natural waters. Journal of Photochemistry and Photobiology A:
Chemistry, 187, 186-195.
Shi, Z., Tao, S., Pan, B., Fan, W., He, H.C., Zuo, Q., Wu, S.P., Li, B.G., Cao, J., Liu, W.X., Xu, F.L.,
Wang, X.J., Shen, W.R., Wong, P.K., 2005. Contamination of rivers in Tianjin, China by
polycyclic aromatic hydrocarbons. Environmental Pollution, 134: 97-111.
Shou, M., Ktausz, K.W., Gonzalez, F.J., Gelboin, H.V., 1996. Metabolic activation of the potent
carcinogen dibenzo(a,h)anthracene by cDNA-expressed human cytochromes P450. Archives of
biocgemistry and biophysics, 38(1), 201-207.
Sicre, M-A., Fernandes, M.B., Pont, D., 2008. Poly-aromatic hydrocarbon (PAH) inputs from the Rhne
River to the Mediterranean Sea in relation with the hydrological cycle: impact of floods. Marine
Pollution Bulletin, 56, 1935-1942.
Silliman, J.E., Meyers, P.A., Eadie, B.J., 1998. Perylene: an indicator of alteration processes or precursor
materials?. Organic Geochemistry, 29(5-7), 1737-1744.
Simpson, C.D., Harrington, C.F., Cullen, W.R., 1998. Polycyclic aromatic hydrocarbon contamination in
marine sediments near Kitimat, British Columbia. Environmental Science and Technology, 32,
3266-3272.
Singh, R., Sram, R.J., Binkova, B., Kalina, I., Popov, T.A., Georgieva, T., Garte, S., Taioli, E., Farmer,
P.B., 2007. The relationship between biomarkers of oxidative DNA damage, polycyclic aromatic
hydrocarbon DNA adducts, antioxidant status and genetic susceptibility following exposure to
environmental air pollution in humans. Mutation Research, 620, 83-92.
Smedes, F., de Boer, J., 1997. Determination of chlorobiphenyls in sediments analytical methods. Trend
in Analytical Chemistry, 16(9), 503-517.
Smith, L.E., Denissenko, M.F., Bennett, W.P, Li, H., Amin, S., Tang, M-S., Pfeifer, G.P., 2000.
Targetting of lung cancer mutational hotspots by polycyclic aromatic hydrocarbons. Journal of the
National Cancer Institute 92(10), 803-811.
Soclo, H.H., Garrigues, PH., Ewald, M., 2000. Origin of polycyclic aromatic hydrocarbons (PAHs) in
coastal marine sediments: case studies in Cotonou (Benin) and Aquitaine (France) areas. Marine
Pollution Bulletin 40(5), 387-396.

66

Stolyhwo, A., Sikorski, Z.E., 2003. Polycyclic aromatic hydrocarbons in smoked fish a critical review.
Food Chemistry 91, 303-311.
Sugiura, K., Ishihara, M., Shimauchi, T., Harayama, S., 1997. Physicochemica properties and
biodegradability of crude oil. Environmental Science and Technology, 31, 45-51.
Suzumura, M., Kokubun, H., Arata, N., 2004. Distribution and characteristics of suspended particulate
matter in a heavily eutrophic estuary, Tokyo Bay, Japan. Marine Pollution Bulletin 49, 496-503.
Syvitski, J.P.M., Peckham, S.D., Hilberman, R., Mulder, T., 2003. Predicting the terrestrial flux of
sediment to the global ocean: a planetary perspective. Sedimentary Geology, 162, 5-24.
Telli-Karako, F., Tolun, L., Henkelmann, B., Klimm, C., Okay, O., Schramm, K.-W., 2002. Polycyclic
aromatic hydrocarbon (PAHs) and polychlorinated biphenyls (PCBs) distributions in the Bay of
Marmara sea: Izmit Bay. Environmental Pollution, 119, 383-397.
Thorsen, W.A., Cope, W.G., Shea, D., 2004. Bioavailability of PAHs: Effects of soot carbon and PAH
source. Environmental Science and Technology, 38, 2029-2037.
Turner, A., Millward, G.E., 2002. Suspended particles: their role in estuarine biogeochemical cycles.
Estuarine, Coastal and Shelf Science, 55, 857-883.
Urbe, I, Ruana, J., Application of solid-phase extraction discs with a glass fiber matrix to fast
determination of polycyclic aromatic hydrocarbons in water. Journal of Chromatography A, 778,
337-345.
Valero-Navarro, A., Fernndez-Snchez, J.F., Medina-Castillo, A.L., Fernndez-Ibez, F., SeguraCarretero, A., Ibez, J.M., Fernndez-Gutirrez, A., 2007. A rapid, sensitive screening test for
polycyclic aromatic hydrocarbons applied to Antarctic water. Chemosphere, 67, 903-910.
Venkatesan, M.I., 1988. Occurrence and possible sources of perylene in marine sediments a Review.
Marine Chemistry, 25, 1-27.
Vigan, L., Farkas, A., Guzzella, L., Roscioli, C., Erratico, C., 2007. The accumulation levels of PAHs,
PCBs, and DDTs are related in an inverse way to the size of a benthic amphipod
(Echinogammarus stammeri Karaman) in the River Po. Science of the Total Environment, 373,
131-145.
Vilanova, R.M., Fernndez, P., Martnez, C., Grimalt, J.O., 2001. Polycyclic aromatic hydrocarbons in
remote mountain lake waters. Water Research, 35(16), 3916-3926.
Voice, T., Weber Jr, W.J., 1983. Sorption of hydrophobic compounds by sediments, soils and suspended
solids-I (Theory and Backgraound). Water Research 17(10), 1433-1441.
Wakeham, S. G., Schaffner, C., Giger, W., 1980b. Diagenetic polycyclic aromatic hydrocarbons in recent
sediments: structural information obtained by high performance liquid chromatography. Physics
and Chemistry of the Earth, 12, 353-363.
Wakeham, S.G., Schaffner, C., Giger, W., 1980a. Polycyclic aromatic hydrocarbons in recent lake
sediment-II. Compounds derived from biogenic precursors during early diagenesis. Geochimica et
Cosmochimica Acta 44, 415-429.
Wang, X. -C, Zhang, Y.-X., Chen, R.F., 2001. Distribution and partitioning of polycyclic aromatic
hydrocarbon (PAHs) in different size fractions in sediments from Boston Harbor, United States.
Marine Pollution Bulletin 42(11), 1139-1149.
Wang, Z., Fingas, M., Lambert, P., Zeng, G., Yang, C., Hollebone, B., 2004. Characterization and
identification of the Detroit River mystery oil spill (2002). Journal of Chromatography A 1038,
201-214.
Wang, Z., Fingas, M., Page, D.S., 1999. Oil spill identification. Journal of Chromatography A 843, 369411.
Weber Jr, W.J., Leboeuf, E.J., Young, T.M., Huang, W., 2001. Contaminant interaction with geosorbent
organic matter: insights drawn from polymer sciences. Water Research, 35(4), 853-868.
Weber Jr, W.J., McGinley, P.M., Katz, L.E., 1991. A distributed reactivity model for sorption by soils
and sediments. 1. Conceptual basis and equilibrium assessments. Environmental Science and
Technology 26, 1955-1962.
Wenzl, T., Simon, R., Kleiner, J., Anklam, E., 2006. Analytical methods for polycyclic aromatic
hydrocarbons (PAHs) in food and the environment needed for new food legislation in the
European Union. Trends in Analytical Chemistry, 25(7), 716-725.
Westerholm, R., Almn, J., Li, H., Rannug, U., Rosn, ., 1992. Exhaust emissions from gasoline-fuelled
light duty vehicles operated in diferent driving conditions: a chemical and biological
characterization. Atmospheric Environment, 26B(1): 79-90.
Williamson, K.S., Petty, J.D., Huckins, J.N., Lebo, J.A., Kaiser, E.M., 2002. Sequestration of priority
pollutant PAHs from sediment pore water employing semipermeable membrane devices.
Chemosphere, 49, 717-729.

67

Wise, S.A., Hilpert, L.R., Byrd, G.D., May, W.E., 1990. Comparison of liquid chromatography with
fluorescence detection and gas chromatography/mass spectrometry for the determination of
polycyclic aromatic hydrocarbons in environmental samples. Polycyclic Aromatic Compounds, 1(1),
81-89.
Wise, S.A., Schantz, M.M., Benner, B.A., Hays, M.J., Schiller, S.B., 1995. Certification of polycyclic
aromatic hydrocarbons in a marine sediment standard reference material. Analytical Chemistry,
67, 1171-1178.
Witt, G., 1995. Polycyclic aromatic hydrocarbons in water and sediment of the Baltic Sea. Marine
Pollution Bulletin, 31(4-12), 237-248.
Witt, G., Siegel, H., 2000. The consequence of the Oder Flood in 1997 on the distribution of polycyclic
aromatic hydrocarbons (PAHs) in the Oder River Estuary. Marine Pollution Bulletin, 40(12),
1124-1131.
Woodhead, R.J., Law, R.J., Matthiessen, P., 1999. Polycyclic aromatic hydrocarbons in surface sediments
around England and Wales, and their possible biological significance. Mar. Poll. Bull. 83(9): 773790.
Xia, X.H., Yu, H., Yang, Z.F., Huang, G.H., 2006. Biodegradation of polycyclic aromatic hydrocarbons
in the natural waters of the Yellow River: Effects of high sediment content on biodegradation.
Chemosphere, 65, 457-466.
Xue, W., Warshawsky, D., 2005. Metabolic activation of polycyclic and heterocyclic aromatic
hydrocarbons and DNA damage: A review. Toxicology and Applied Pharmacology, 206, 73-93.
Yamashita, N., Kannan, K., Imagawa, T., Villeneuve, D.L., Hashimoto, S., Miyazaki, A., Giesy, J.P.,
2000. Vertical profile of polychlorinated Dibenzo-p-dioxins, Dibenzofurans, Naphthalenes,
Biphenyls, Polycyclic aromatic hydrocarbons, and alkylphenols in a sediment core from Tokyo
Bay, Japan. Environmental Science and Technology, 34, 3560-3567.
Yan, J-H., You, X-F., Li, X-D., Ni, M-J., Yin, X-F., Cen, K-F., 2004. Performance of PAH emission from
bituminous coal combustion. Journal of Zhejiang University Science 5(12), 1554-1564.
Yang, H-H., Chien, S-M., Lo, M-Y., Lan, J.C-W., Lu, W-C., Ku, Y-Y., 2007. Effects of biodiesel on
emissions of regulated air pollutants and polycyclic aromatic hydrocarbons under engine durability
testing. Atmospheric Environment, 41: 7232-7240.
Yang, H-H., Jung, R-C., Wang, Y-F., Hsieh, L-T, 2005. Polycyclic aromatic hydrocarbons emission from
joss paper furnaces. Atmospheric Environment, 39, 3305-3312.
Yang, H-H., Lee, W-J., Chen, S-J., Lai, S-O., 1998. PAH emission from various industrial stacks. Journal
of Hazardous Materials 60, 159-174.
Youngblood, W.W., Blumer, M., 1975. Polycyclic aromatic hydrocarbons in the environment:
homologous series in soils and recent marine sediments. Geochimica et cosmochimica Acta 39,
1303-1314.
Yunker, M.B., Macdonald, R.W., Goyette, D., Paton, D.W., Fowler, B.R., Sullivan, D., Boyd, J., 1999.
Natural and anthropogenic inputs of hydrocarbons to the Strait of Georgia. The Science of the
Total Environment, 225, 181-209.
Yunker, M.B., Macdonald, R.W., Vingarzan, R., Mitchell, H., Goyette, D., Sylvestre, S., 2002. PAHs in
the Fraser River basin: a critical appraisal of PAH ratio as indicators of PAH source and
composition. Organic Geochemistry 33, 489-515.
Zakaria, M.P., Horinouchi, A., Tsutsumi, S., Takada, H., Tanabe, S., Ismail, A., 2000. Oil pollution in the
Straits of Malacca, Malaysia: Application of molecular markers for source identification.
Environmental Science and Technology, 34, 1189-1196.
Zhang, Y., Tao, S., 2009. Global atmospheric emission inventory of polycyclic aromatic hydrocarbons
(PAHs) for 2004. Atmospheric Environment, 43, 812-819.
Zhou, J.L., Fileman, T.W., Evans, S., Donkin, P., Llewellyn, C., Readman, J.W., Mantoura, R.F.C.,
Rowland, S., 1998. Fluoranthene and pyrene in the suspended particles and surface sediments of
the Humber Estuary, UK. Marine Pollution Bulletin, 36(8), 587-597.
Zhou, J.L., Fileman, T.W., Evans, S., Donkin, P., Readman, J.W., Mantoura, R.F.C., Rowland, S., 1999.
The partition of fluoranthene and pyrene between suspended particles and dissolved phase in the
Humber Estuary: a study of the controlling factors. The Science of the Total Environment 243,
305-321.
Zielinska, B., Sagebiel, J., Arnott, W.P., Rogers, C.F., Kelly, K.E., Wagner, D.A., Lighty, J.S., Sarofim,
A.F., Palmer, G., 2004. Phase and size distribution of polycyclic aromatic hydrocarbons in diesel
and gasoline vehicle emissions. Environmental Science and Technology, 38, 2557-2567.

68

IV.

DISTRIBUTION AND SOURCE OF POLYCYCLIC AROMATIC


HYDROCARBONS (PAHs) IN SURFACE SEDIMENTS FROM THE
SIAK RIVER, ITS ESTUARY AND THE ADJACENT COASTAL AREA
OF RIAU PROVINCE, INDONESIA

Muhammad Lukman*, Wolfgang Balzer* and Christine Jose**


*Marine Chemistry Working Group, Department of Biology/Chemistry, University of Bremen,
Leobener Strasse, 28359 Bremen, Germany
**Department of Chemistry, University of Riau, Jl. Simpang Baru Panam, 28293 Pekanbaru,
Riau, Indonesia

Abstract
The distribution and sources of polycyclic aromatic hydrocarbons (PAHs) were examined
in sediments from the Siak river, its estuary and the Riau coast of Sumatra, Indonesia. PAH
concentrations were determined by HPLC with UV and fluorimetric detectors in two size
fractions (sand: 2 mm 63m; mud: <63 m). The sum of the 16 PAH selected by EPA
(PAHs) ranged from 126 5474 ng g-1 d.w. In general, the coarse fraction (164 to 5474 ng g-1
d.w., median = 837 ng/g d.w.) contained circa twice as much PAH as the mud fraction (126 to
1314 ng g-1 d.w., median = 517 ng/g d.w.). The concentrations were higher than in similar
systems in Asia suggesting a link to land use change and peat burning practices. Organic carbon
contents varied greatly from 0.01% to 24% in the sand, but only slightly in the mud from 0.34%
to 3.70%. The composition of PAHs in both fractions was largely similar to each other with 3-,
4- and 5-ringed PAHs being most abundant. Molecular-weight ratios and isomer ratios for
source apportionment indicated a dominance of pyrogenic PAHs, but petrogenic origins showed
significant signatures in the vicinity of residential and industrial areas.

Keywords: PAHs; Sediment; Siak River; Estuary; Riau Coast

MANUSCRIPT PREPARED TO BE SUBMITTED

(TABLES MENTIONED IN THIS MANUSCRIPT ARE PROVIDED IN APPENDIX 4: p.131-133)

69

4.1.

Introduction
Polycyclic aromatic hydrocarbons (PAHs) are a class of toxic and carcinogenic organic

pollutants containing two or more fused aromatic rings, produced mainly through the burning of
biomass and the incomplete combustion of fossil fuels (Neff, 1979; IARC, 1987). In addition,
they constitute a substantial fraction crude oil and its petroleum products (Requejo et al., 1996;
Wang et al., 1999). Concern over these compounds worldwide is not only due to their
carcinogenic properties, but also to their ubiquity and persistence in the aquatic environments
which increase potential exposures. Human exposure is therefore a possibility through
consumption of contaminated food in which PAHs experienced bio-concentration and biomagnification, and inhalation (Hellou et al., 2005; Ribeiro et al 2005; Bai et al., 2009).
Land use change in Indonesia by peatland burnings particularly in Sumatra and
Kalimantan, is a major source of pollutants, particularly the emission of PAHs in addition to
CO2 causing climate change. The worst peatland fire episode occurred during a severe El Nio
event (1997/1998), which emitted up to 2.57 Gt Carbon in 1997 during the damage of ca. 6.8
Mha (or 34%) of Indonesian peatland (Page et al., 2002). In comparison, the global net emission
of CO2 from land-use change was recently estimated at 2.4 Gt y-1 and 50% of this flux came
from tropical Asia (IPCC, 2000). The environmental and health effects of haze and smoke have
been widely recognized ever since (e.g. Langmann et al., 2007; Fang et al., 1999; Parameswaran
et al. 2004). PAHs are inevitable products of peatland and biomass burnings. During another
severe fire episode in 2005 in Sumatra, See et al. (2007) observed ca. 561 ng m-3 of the total 16
PAHs on the US EPA priority pollutant list in the emission after peatland burnings around
Dumai. In comparison Pekanbaru (a city affected by the haze) showed a concentration of ca.
135 ng m-3. The latter was even higher than that (ca. 116 ng m-3) of polluted urban areas in
Beijing, China (Zhou et al., 2005). Thus, it can be expected that peatland and biomass burnings
in Sumatra over decades have produced a significant amount of PAHs which have polluted the
surrounding lands and aquatic systems. However, PAH investigation on the aquatic systems in
Sumatra has not been comprehensively undertaken.
Sediment is a well-known pollutant recorder which integrates not only temporal and
spatial pollutant loads of modern settings, but also envoys history of the past. The capacity of
sediments to concentrate and retain elements and hydrophobic organic compounds results from
two inter-correlated sets of physical (grain size) and chemical (organic and mineral
composition) properties (Horowitz & Elrick, 1987). The level of total PAH in the sediment
greatly varied ranging from a few ng g-1 up to hundreds of g g-1 for highly oil contaminated
intertidal sediments (e.g. Sauer et al., 1998). In most studies, PAH assessments have focused on
the bulk sediment content. However, it is becoming increasingly common to differentiate
between grain sizes for better understanding of the pollutant distribution. Furthermore, the

70

assessment of grain size content may help in designing remediation efforts (Ahrens & Depree,
2004).
Like many other pollutants, the PAHs distribute among grain size fractions (e.g. Maruya
et al., 1996; Budzinski et al., 1997; Tolosa et al., 2004; Prahl & Carpenter, 1983; Simpson et al.,
1998; Wang et al 2001; Rockne et al., 2002; Ahrens & Depree, 2004). Enrichment in a
particular fraction, e.g. in coarse (sand: 2mm 63m) or fine (mud: <63m), is strongly related
to specific PAH-geosorbent interactions reflecting the state of environmental conditions.
Several studies showed that in most polluted areas PAHs are enriched in the coarse fraction and
especially in carbonaceous particles such as coal, charcoal, soot or black carbon and plant debris
(e.g. Prahl & Carpenter, 1983; Simpson et al., 1998; Wang et al 2001; Ghosh et al., 2000,
Rockne et al., 2002). In most cases, these particles constitute a minor part of the bulk dry weight
mass, but provide strong affinity for PAHs. Therefore, the PAH enrichment in the coarse
fraction suggests that PAHs and those carbonaceous particles are from the same sources. On the
other hand, the fine (silt/clay) fraction is expected to accumulate more PAHs due to a larger
surface area to mass ratio provided for adsorption (e.g. Karickhoff et al., 1979; Maruya et al.,
1996). Charlesworth et al. (2002) revealed the importance of the fraction <63m in assessing
the distribution of PAHs in areas distant from direct input, and vice versa.
This study aimed (1) to examine the PAH distribution between two size fractions of
surface sediments from the Siak river, its estuary and the adjacent Riau coastal areas of Riau,
Sumatra, Indonesia; (2) to appoint the possible sources; and (3) to understand the role of

sedimentary organic matter for each fraction in PAH distribution, for which the very
humic-rich condition of the Siak river system appear to be especially suited.
4.2.

Study Area and Methods

4.2.1.

Study Area & sampling locations


The study areas cover the Siak River from the upstream tributaries of the Tapung Kanan

and the Tapung Kiri down to the mouth of the estuary; it further includes the coast extending
from the oil harbour Dumai to the southern part of Selat Panjang where some oil ridges are
situated (Fig. 1). The Siak River and its estuary stretch out over 300 km and drain a large area
of lowland and peat swamps within various landscapes, including huge palm-oil plantations,
rainforests, small urban centres, the capital city of Pekanbaru, a pulp and paper factory, and oil
refineries near the Siak mouth. During 2004 to 2006, four expeditions were carried out to take
samples from a total of 27 stations.
With regards to potential PAH sources, the study areas are affected by frequent events of
dense smog from commonly practiced agricultural/biomass burnings and forest/swamp fires,
which are significant contributors to health problems in the region. (e.g. Davies &

Unam,

71

1999; Quah 2002). In addition, oil discharges from routine river-boat transports as well as from
oil industry related activities (production, transportation and disposals of residues) are
observable.

S269

S226

S227
S267

S228

S 266

Sumatra

S253
S230
S251
S 250

INDONESIA

S231
S138
S232

S142
S143
S145
S101
S42
S104 S116
S105
S20 S24

S35

Fig. 1. The map of the sampling stations


4.2.2.

Sample Collection and Fractionation


The samples were collected directly from a small vessel using a sediment grab sampler.

They were immediately homogenized with a stainless scoop before being placed in closed
aluminium jars. During homogenization, foreign objects such as big plant sticks, stones or any
other synthetic waste were removed. The samples were kept cool (ca. 4oC) during transported to
the laboratory, where they were frozen (-20oC) and stored until further treatment. Sizefractionation was carried out by wet sieving to get the sand/coarse fraction (2mm - 63 m) and
the mud/fine fraction (< 63 m after Udden-Wentworth scale).
4.2.3.

Analytical Methods

4.2.3.1. PAH Determination


The samples were analyzed for 16 PAHs priority pollutants selected by US EPA
including naphthalene (NAPH), acenaphthylene (ACYN), acenaphthene (ACEN), fluorene
(FLU), phenanthrene (PHEN), anthracene (ANTH), fluoranthene (FLA), pyrene (PYR),
benzo(a)anthracene

(BaA),

chrysene

(CHRY),

benzo(b)fluoranthene

(BbFLA),

benzo(k)fluoranthene (BkFLA), benzo(a)pyrene (BaP), dibenzo(a,h)anthracene (DANTH),


benzo(g,h,i)perylene (BPERY) and indeno(1,2,3-c,d)pyrene (IPYR). Baseline separation and

72

quantification of these analytes were performed on a reverse phase RP-C18 column (250 x 4.6
mm, 5m, BAKERBOND, J.T. Baker Inc) by using a high performance liquid chromatograph
(HPLC LKB 2249 Broma) with ultraviolet (2151 Variable Wavelength Monitor, LKB Bromma)
and programmable fluorescence detectors (Hawlett-Packard 1046A). The mobile phase was a
combination of isocratic and a linear binary gradient elution of acetonitrile(ACN)/water,
programmed from 55% to 100% ACN at a constant flow rate of 1 mL min-1, which set for a
total of 45 minutes.
Sample extraction and work-up procedure were as follows. 10 g of the sediment fractions
were spiked with three surrogate perdeuterated PAH standards: d10-phenanthrene (PHEN D10),
d10-fluoranthene (FLA D10) and d12-perylene (PERY D12), and extracted in Soxhlet-modified
extractors (SoxTec) with solvents suggested by the ICES Method (ICES, 1997) for 6 8 hours:
the first cycle was extracted by acetone followed by a mixture of acetone/hexane (1/9 ; v/v) in
the second cycle. The extracts were then combined and reduced to ca. 1 mL by a rotary
evaporator. A 1:2 (w/w) deactivated Al2O3 (10%) / SiO2 (3%) column plus anhydrous Na2SO4
on the top to remove the co-extracted water, was used for cleaning-up the extract. During the
clean-up process the extracts were eluted with 40 mL of degassed 3/7 (v/v)
dichloromethane/hexane. The excess solvent of the clean extracts was removed by evaporator to
ca. 1 mL. Then, the extracts were subject to solvent exchange into ACN for HPLC analysis.
HPLC grade acetonitrile and analytical grade solvents (acetone, dichloromethane and
hexane) obtained from Fischer Scientific were used throughout the analytical procedures. A
certified mixture of 16 PAHs standards (purchased from Dr. Ehrenstorfer GmbH, Germany)
were used to identify and quantify the analytes. Detector limit of detection ranged from 0.4
ng/mL (BaP) to 12 ng/mL (NAPH) for fluorescence detection, and 38 ng/mL for ACYN with
UV detection. The three perdeuterated surrogate PAHs were used to accomplish the procedural
efficiency, reproducibility and data correction from the evaporative losses during extraction and
work-up procedures. The recovery ranged respectively from 40.6% to 114% (mean = 82.8%),
69.4% to 141% (m = 97.5%), and from 73.9% to 121% (m = 94.6%) for for PHEN D10, FLA
D10 and PERY D12 with their relative standard deviations (RSD) of 15.2%, 14.6% and 12.6%.
4.2.3.2. Determination of Sedimentary Organic Matter
Sedimentary organic carbon (OC) and organic nitrogen (ON) content were determined
using a Vario EL III CHNOS Elemental Analyser at 1800oC (during combustion) at the
GeoScience laboratory at the University of Bremen. 20 - 25 mg of the homogenized sediment
fraction were placed in a silver boat, treated with 30 L of 1 M HCl to remove inorganic carbon
and dried at 40oC. This process was repeated 2 3 times to make sure that all carbonates were
transformed into carbon dioxide. The relative standard deviation of the method was 5.9 %.

73

4.3.

Results & Discussion

4.3.1.

Geochemistry of sediment fractions


Organic carbon contents (OC) greatly varied from 0.01% to 24% by weight in the sand

fraction, but only slightly in the mud from 0.34% to 3.70% (Table 1, p. 131). In general, the
coarse fraction contained higher OC, but the fraction contributed less than 25% to the bulk
sediment dry-weight. It indicates the presence of organic-rich particles in the sand fraction such
as wood remnants, fragmented plant debris and black particles which were apparently abundant.
A great variation of OC in the sand fraction between locations suggests a relevance of local
inputs. Coarse estuarine sediments had higher OC than both river and coastal sediments,
particularly the station S142 (24%) was characterized by a considerable presence of black
particles. The highest OC in the river was observed at S42 (14%) between Pekanbaru city and
the Perawang industrial area. In contrast, the OC content in the mud fraction varied less between
locations and decreased toward the coast.
The C/N ratios (Table 1) suggest dissimilar sources of organic matter between the
fractions. The C/N end-member values of ca. 7 and higher than 20 are commonly used to define
marine and terrigenous origins, respectively (Meyer, 1994; Holtvoeth et al., 2005). Values
between 6-12 suggest a mixture dominated by marine-originated organic matter (Ruttenberg &
Go i, 1997). High percentages of terrestrial organic matter were found in the sand fractions
with median C/N ratio of 19.4, 12.8 and 31.8 for the river, estuary and the coast, respectively.
These median ratios were about two-fold higher than those of the fine fraction in the river
sediment ( 9.1) and in the coastal sediments (14.3). However, C/N ratios of both sand and mud
fractions of the estuarine sediments showed relatively similar values with the mean of 14.
Higher C/N values (>20) are commonly associated with terrestrial plants, thermal degradation
of biomass, or peat (Holtvoeth, 2004, Pillon et al., 1985). On the other hand, a lowered C/N
ratio in terrestrial OC is often tied to decomposition processes such as humification and
mineralization (e.g. Zech et al., 1997; Holtvoeth et al., 2005). Furthermore, a partial retardation
of the organic nitrogen remineralization could have resulted in relatively high ON contents in
the organic-rich sediment (e.g. Balzer et al., 1998). Comparable patterns of the C/N ratio
between the sand and mud fractions were confirmed from similar peatland system of the
Mahakam Delta in Borneo (Pillon et al., 1986). The lower C/N ratio in the fine fraction of the
Mahakam Delta sediments was related to high percentage of hydrolysable materials represented
by amino acid and ammonium which might be considered of animal origins rather than of
terrestrial higher plant sources. To sum up, low OC and C/N ratios in the mud fraction suggest
that the type of organic matter in the mud fraction was different from that of the sand fraction
which is associated with terrestrial plant debris, peat and thermal degradation of biomass (black

74

carbon). These sources were likely to reach the aquatic environment through soil washout or
land abrasion during high precipitation.
4.3.2.

Content and Distribution of PAH


The contents and distribution of the 16 PAHs (expressed as PAHs) in the sediments

from the three different environments are summarized in Table 2 (p. 132 see also Fig. 2). Owing
to a low mud fraction ( 2%), river stations of S104, S105 and S35 were analysed only for the
sand fraction. The PAHs in the sand fraction of the Siak river, estuarine and coastal sediments
range from 164 to 5474 (median=556) ng g-1 d.w., 208 3913 (425) ng g-1 d.w., and 594 2495
(1142) ng g-1 d.w., respectively. While in the mud sediment, corresponding contents varied from
319 to 1143 (521) ng g-1 d.w., 126 to 584 (468) ng g-1 d.w., 443 to 1314 (633) ng g-1 d.w. The
highest PAHs content (>5000 ng g-1 d.w.) in the sand fraction was found at S42 in the river,
followed by S138 in the estuary (~4000 ng g-1 d.w.), S253 and S269 in the coast (>2000 ng g-1
d.w.). In the mud fraction, the highest PAHs contents (1000-1300 ng g-1 d.w.) were observed
at S42 in the river and S269 and S226 in the coast. The mud fraction of the estuarine samples
with a mean content of 410 ng g-1 d.w. had intermediate PAHs concentrations, which were not
appreciably different between sampling locations. The PAHs were generally about two times
higher in the sand fraction. Sometimes, the PAHs concentrated up to four times higher in the
sand fraction e.g. at S42 (river), S250 (coast) and S232 (coast) , and even to seven times higher
at S138 (estuary) suggesting the existence of high affinity organic matter for PAHs associated
with the sand fraction. Enriched PAH contents in the coarse/sand fraction were also reported
from other riverine, estuarine and coastal systems (e.g. Oen et al., 2006; Wang et al., 2001;
Ahrens & Depree, 2004; Rockne et al., 2002; Yang et al., 2008). These studies figured out the
role of carbonaceous geosorbents such as coal-/wood-derived particles and black carbon in
sorbing more PAHs in the sand fraction.

75

6000

5000

Sand

PAHs (ng/g d.w.)

Mud
4000

Bulk

3000

2000

1000

S 24
S 101
S 20
S 104
S 105
S 35
S 116
S 42
S 145
S 143
S 142
S 138
S 134
S 252
S 125
S 250
S 251
S 269
S 226
S 227
S 228
S 267
S 266
S 253
S 230
S 231
S 232

River

Estuary

Coast

Fig. 2. Distribution of PAHs in sand and mud fractions of surface sediments from the Siak River, estuary
and the coastal areas. Stations are arranged downstream from the Siak tributaries over the estuary to the
coastal region.

A bulk PAH content (PAHBulk) was calculated for each station from the content for each
fraction and the proportion on the total sediment dry mass. PAHBulk of all aquatic systems
ranged from 145 to 1234 ng g-1 d.w.(Table 2). The coastal sediments turn out to contain the
highest PAHBulk with a median of 577 ng g-1 d.w., followed by the river and the estuary of 484
and 443 ng g-1 d.w., respectively. This calculation placed S42, S138, S253 and S269 as stations
having the highest PAHBulk (>1000 ng g-1 d.w.). In addition, these locations were also
identified as having the highest PAH concentration within the sand fraction alone. The spatial
distribution of PAHBulk as well as PAHs for the sand and the mud fractions did not show a
clear pattern between sampling locations towards the river mouth of the estuary area suggesting
various local inputs (Fig 2). However, the highest PAHBulk contamination was found in the
vicinity of urban and industrial centres (S42, S226, S269, and S253). In contrast, the lowest
PAHBulk contents were detected at the river upstream (S104) and in the estuary (S125).
Because most studies in the literature considered only the bulk assessment, the PAHBulk
could bring comparable information on their pollution level. Values of PAH >1000 ng/g d.w.
mostly represent chronically polluted industrialized areas and harbours (Baumard et al., 1998;
Tolosa et al., 2004). Therefore, the river sediment around the capital city of Pekanbaru (S42),
the coastal sediments in the vicinity of Dumai (S269, S226) and in the mouth of Selat Panjang
(S253) can be considered highly polluted, while the other locations show moderate pollution.

76

Direct comparison to those of other studies is somewhat difficult due to variation in numbers of
studied-PAHs (not all studies investigated all the 16 US EPA compounds), the analytical
performances and the background conditions. However, for a general idea of the PAH level,
relevant information from studies in several Asian and European countries are presented in
Table 3. The median values of PAH

Bulk

in surface sediments of Riau aquatic environments

were higher by a factor of two or more than those on the opposite side of the Malacca Strait
(Zakaria et al., 2002), the industrialized Gulf of Thailand (Boonyatumanond et al., 2006), and
the Pearl River Estuary plus its adjacent coastal areas of China (Luo et al., 2006). Higher PAH
in sediments from this study than in areas from the neighboring countries are assumed to be due
to greater inputs of PAHs from widespread, intense plantation and peat burnings. These sources
produced significant amount of PAH compared to other pyrogenic sources such as combustion
of fossil fuels which are typical sources of sedimentary PAH in urban and industrial
environments (e.g. Kakareka & Kukharchyk, 2003; Miguel et al., 1998).
4.3.3.

PAH & OC Relationship


The PAHs and the OC content in the sand fraction are generally correlated (R2 = 0.70,

Fig. 3). In this correlation, the S253 (coast), S269 (Dumai port), and S142 (estuary) are treated
as outliers due to high PAHs coupled with low OC and vice versa, and thus omitted for this
regression. In contrast, there was no correlation between PAHs and the OC content in the mud
fraction. These results suggest that organic matter (OM) associated with the sand fraction has a
stronger affinity for PAHs. On the other hand, in the fine fraction PAHs was independent on
the OC content suggesting that PAHs were reluctant to associate with the OM, or that the OM
content was relatively high for given PAHs content.
Furthermore, we figure out that there are actually two groups of the PAH-OC association
regardless of the fractions. First, the PAHs with <1% OC was not correlated (Fig. 4a), and
second, PAHs with >1% OC was linearly correlated (R2= 0.78, Fig. 4b). The first type seems
to affirm a specific PAH-organic particle interaction which has nothing to do with the OC
content. It might be that PAH-particles came from the same sources entering the aquatic
systems. The second type supports the idea that PAHs in both fractions was controlled by its
OM content and properties. But, the association was in favor of the coarse-attached OM. As no
PAH-OC correlation observed for the fine fraction (Fig. 3), this type emphasizes that the OM of
both fractions have different source and composition which control the PAHs sequestration. The
coarse fraction-OM was most probably attributed to those from combustion-derived particles or
black carbon, vascular plant debris and peat, while the fine fraction-OM was derived mainly
from humification. The reluctance of PAH to associate with the fine fraction-OM is probably
due to less aromatic fraction of the OM structure stemming from highly degraded peatland. As
associations of PAH with the OM of the fine fraction occurs mostly as a direct function of fine

77

particle surface area, it could be then assumed that the low C/N ratio of the fine fraction would
represent the condition of dissolved organic matter (DOM). Kalbitz & Geyer (2002) observed
that low C/N ratio of DOM is one characteristic property of a degraded peatland. Prahl and
Carpenter (1983) also observed evidences for a lack of correlation between PAHs and humic
substances that were enriched in the fine (<64m) fraction of the sediment samples from the
Washington coastal region, although no specific reason was mentioned.
To sum up, this PAH-OC association could suggest three modes of which the PAH
integrated into the aquatic systems. First, the PAH entered the aquatic environment favorably
binding to the OM with high affinity for PAH (i.e. black particles, vascular plant debris, or peat)
as found in S42, S116, S143, S138, S250, S232. Second, the PAH-OM came together from
similar sources, as present in the coarse fraction comprised of low OC content and low total
sediment weight. It was observed at the S230, S231, S269, and probably S226, and S253. Third,
the PAHs were unfavorable associated with the fine fraction due to uncondensed, less aromatic
fraction of the humic substances from the degraded peatland.

6000

5000

PAHs (ng/g d.w.)

Sand Fraction
4000

PAHs = 295.4*OC + 349.1


R2 = 0.7

3000
S253: the Coast
S269: Dumai Port
2000

1000

Estuarine S142
Sediment
0

10

15

20

25

OC (%)

Fig. 3. Correlation between PAHs and the organic carbon contents in the sand ( ) and mud () fractions
of the sediments from the Siak River, the estuary and the coast (N=24, R2 =0.70 for the sand fraction,  =
outliers).

78

PAHs (ng/g d.w.)

3000

PAH vs OC <1%

2500
2000
1500
1000
500
0
0.0

0.1

6000

0.2

0.3

0.4

0.5

PAH vs OC >1%

PAHs (ng/g d.w.)

5000
4000
3000
y = 351.24x - 128.98
R = 0.78

2000
1000
0
0.0

2.0

4.0

6.0

8.0

10.0

12.0

14.0

16.0

OC (%)
Fig. 4. Specific type of PAH-OC correlations for both fractions: (a) PAHs to low (<1%) OC content
suggesting that the PAH and the OC entered the aquatic systems from the same sources, and (b) PAHs
to high (>1%) OC indicating that the PAH sorption to the organic matter was a function of the OC
content and composition.

4.3.4.

Relative Composition of PAHs


The relative composition of PAHs was examined by ringed-group PAHs classified as 2

ring (NAPH), 3 rings (ACYN, ACEN, FLU, PHEN, ANTH), 4 rings (FLA, PYR, BaA, CHRY),
5 rings (BbFLA, BkFLA, BaP, DANTH), and 6 rings (BPERY, IPYR). The composition for
both the sand and the mud fractions was largely identical with 3 to 5-ringed PAHs being most
abundant. The 3-ringed PAH dominated the composition for both fractions ranging from 34.3%
to 72.7% in all aquatic environments, followed by 4-ringed and 5-ringed groups. The most
abundant compound of the 3-ringed group is acenaphthylene. On average, it made up to 30.3%,
28.9%, and 67.7% of the relative individual composition in the sand fraction from the river, the
estuary and the coasts, respectively. It made up 28.6%, 32.9%, and 68.9% of the relative
composition in the mud fraction for the respective areas.

79

The other prevalent groups were the 4- and 5-ringed PAHs. The relative composition of
both groups greatly varied among the stations in all environments ranging on average from
10.6% to 21.0% for the 4-ringed PAHs, and from 9.12% to 32.7% for the 5-ringed PAHs. These
groups had a tendency to concentrate around the urban centres of Pekanbaru and Dumai cities.
However, the Siak tributaries of Tapung Kanan and Tapung Kiri Rivers were dominated by the
5-ringed

PAH

suggesting

potential

repository

for

the

high

molecular

PAHs.

Dibenzo(a,h)anthracene is the most abundant compound of the 5-ringed PAH in almost all the
stations. This compound was particularly pronounced in the Siak tributaries, in the estuary
(S143, S138, S134) and the Siak Kecil (S250), and in the coast (S253), where it made up ca.
50% to 90% of the relative composition of the 5-ringed PAH. Since those stations commonly
surrounded by large plantation and peatland, we assumed that biomass and peat burnings
occurring in the surrounded land could be the source of this compound. If this is true, increased
contents of DANTH and other 5-ring PAHs could become signatures of peat-generated PAHs.
However, further investigations (field and laboratory) on profiles of peat-generated PAHs from
this area are needed. Widespread distribution of 4- and 5-ringed PAH in the areas strongly
suggests invasive pyrogenic non-point sources which were most likely introduced to the aquatic
environments through soil washout and atmospheric deposition.
4.3.5.

Source Apportionment
It is widely accepted that anthropogenic PAHs stem from two general sources: pyrogenic

and petrogenic, which in most cases co-exist in aquatic sediments. In identifying which source
is dominant, the apportionment for the PAH in many studies make use of specific characteristics
of low over high molecular weight ratio (LMW/HMW), and specific isomeric ratios such as
ANTH/(ANTH + PHEN), FLA/(FLA + PYR), BaA/(BaA+CHRY) (Neff 1979; Budzinski et
al., 1997; Yunker et al., 2002; De Luca et al., 2005; Soclo et al., 2000). Petrogenic origin from
maturation organic matters is typically marked by a high proportion of LMW (2 and 3 ring)
over HMW (4 to 6 rings) PAHs. For instance, naphthalene (2 rings) contributes more than half
of the total non-alkylated polyaromatic compounds in crude oil, followed by three-ring
substances, PHEN and FLU (Requejo et al., 1996).
The specific isomer ratios have often been used due to their molecular stability against
the increase temperature during pyrolysis. For example, phenanthrene is known to be
thermodynamically more stable than the kinetically-stable isomer anthracene (Budzinski et al.,
1997). The proportion of anthracene increases as processes involve higher temperatures. PAHs
from combustion sources have typical values for the ratio ANTH / (ANTH + PHEN) > 0.1;
FLA/(FLA-PYR) >0.5, and BaA/(BaA+CHRY) > 0.35. In contrast, PAHs associated with
petroleum e.g. crude oil, have typical values of those corresponding isomeric ratios of <0.1,
<0.4, and <0.2 respectively (Yunker et al., 2002). For instance, Alascan Crude Oil had for the

80

respective ratios values of 0.03, 0.26 and 0.10 (Requeojo et al., 1996). However, the boundaries
between the assigned values are not clear-cut, and any value falling between those determining
values is usually considered as mixtures of petroleum and combustion sources.
These ratios were applied in this study, and the assignment was carried out for both
fractions. The values of the isomeric ratios were cross-plotted to get a tendency of the data (Fig.
5). Our results showed that the ratio of LMW/HMW for most of the stations was < 1 clearly
indicating combustion sources, except for S104 and S105 in the upstream area of the River,
S125 in the river mouth and S251 the mouth of Siak Kecil, and S267 in the northern mouth of
the Bengkalis canal. The ratio of ANTH/(ANTH+PHEN) was respectively 0.1 0.05 (mean
standard deviation, n=9), 0.22 0.18 (n=8), and 0.2 0.2 (n=10) in the River, the estuary and
the coast confirming predominance of combustion sources. Likewise, the ratio of FLA/(FLAPYR) was respectively 0.48 0.24, 0.65 0.13, and 0.69 0.17. High values of this ratio (>
0.5) indicate typical biomass burnings such as grasses, wood, or agricultural debris. The ratio of
BaA/(BaA + CHRY) was 0.36 0.13, 0.36 0.15, and 0.47 0.17 for the River, the estuary
and the coast. These values were slightly larger than those of petroleum-combustion sources
(upper limit 0.35).
All ratios clearly showed that pyrogenic sources were dominant, mainly from biomass
(vegetation) with a small contribution from petroleum combustion (Fig. 5). These results could
be attributed to widespread and intensive agricultural burnings and forest/peat swamp fires that
occurs frequently in those areas particularly since the last severe outbreak of the El Nio.
Accordingly, pyrogenic PAHs, which were primarily formed during dry seasons and especially
during burn-episodes, were delivered to the river and the estuary through soil washout,
intensified during high rain and flood seasons. The widespread signature of pyrogenic PAHs
might also reflect the significance of atmospheric (wet/dry) deposition.
However, the signatures of petrogenic PAHs were also observed in the river, the estuary
and the coast. The river sediments are to a great extent contaminated by petrogenic sources,
which is particularly observable in the Siak upper tributaries (S101) and around the capital city
of Pekanbaru (S35, S116), the Siak Kecil River (S250), Dumai port (S269), and the mouth of
Bengkalis canal (S267). The petrogenic PAHs in the upstream river might stem from petroleum
discharges from many local small boats which were observable during the sampling. In general,
the petroleum contamination observed in the other stations could be attributed to oil discharges
from ships operation, regular transportation and cargo activities along the river as the study area
is one of the biggest oil-production sites in Sumatra. Strong petrogenic signatures were also
reported from estuarine and coastal sediments of the neighboring country, Malaysia (Zakaria et
al., 2002).

81

Petroleum

Biomass Combustion

1.0

0.8

Combustion

BaA / (BaA + CHRY)

0.9

0.7
0.6
0.5
0.4
0.3
0.2

Petroleum

0.1
0.0

0.6

Combustion

ANTH / (ANTH + PHEN)

0.7

0.5
0.4
0.3
0.2
0.1

Petroleum

0.0

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

FLA / (FLA + PYR)


Fig. 5. Cross plot of PAH isomeric ratios of the sand and mud sediment fractions from the Siak River (),
the estuary ( ) and the coast ( ).

4.4.

Conclusion
The sediments contained PAHs at moderate to high level. The sand/coarse fraction

generally contained higher PAHs and OC content than did the mud/fine fraction. A PAH-OC
linear correlation was shown only by the sand fraction. Enriched PAHs in the high-OC sand
fraction plus the linear correlation indicated the existence of particular organic matter with highaffinity for PAHs, such as black carbon, vascular plant, and peat. The spatial distribution
showed no clear pattern in distance to the river mouth. But, increased content of the PAHs was
centered at urban and industrial areas. The high molecular weight was widespread predominant
and the molecular ratios for source apportionment provide further evidences for pyrogenic
sources: biomass and petroleum combustions. PAHs were delivered to those aquatic system by
both land-water and air-water transports.

82

Acknowledgments
This study is part of the German-Indonesian SPICE Project Cluster 3.1., funded by the
Federal German Ministry of Education, Science, Research and Technology (BMBF, Bonn), and
supported by the German Academic Exchange Service (DAAD), and we thank you for all
supports. Also, we would like to extent our grateful to scientists and students from University of
Riau, Indonesia, and to SPICE German colleagues for their contribution during sampling
campaigns and discussion. Furthermore, we appreciate the help of our colleagues Dipl. Eng.
Immo Becker (Marine Science working group, Dept. of Biology/Chemistry, University of
Bremen) and Ms. Dominique Schobes for laboratory assistances, and Ms. Hella Buschhoff
(Department of Geoscience, University of Bremen) for organic carbon measurement. Last but
not least, we thank all reviewers for their constructive critics and comments.

References
Agarwal, T., Khillare, P.S., Shridhar, V., 2006. PAHs contamination in bank sediment of the Yamuna
River, Delhi, India. Environmental Monitoring and Assessment 123, 151-166.
Ahrens, M.J., Depree, C.V., 2004. Inhomogenous distribution of polycyclic aromatic hydrocarbons in
different size and density fractions of contaminated sediment from Auckland Harbour, New
Zealand: an opportunity for mitigation. Marine Pollution Bulletin 48, 34-350.
Bai, Z., Hu, Y., Yu, H., Wu, N., You, Y., 2009. Quantitative health risk assessment of inhalation
exposure to polycyclic aromatic hydrocarbons on citizens in Tianjin China. Bulletin
Environmental Contamination and Toxicology, 83: 151-154.
Balzer, W., Helder, W., Epping, E., Lohse, L., Otto, S., 1998. Benthic denitrification and nitrogen cycling
at the slope and rise of the N.W. European Continental Margin (Goban Spur). Progress in
Oceanography 42, 111-126.
Baumard, P., Budzinski, H., Garrigues, P., 1998. Polycyclic aromatic hydrocarbons in sediments and
mussels of the western Mediterranean sea. Environmental Toxicology and Chemistry 17(5), 765776.
Boonyatumanond, R., Wattayakorn, G., Togo, A., Takada, H., 2006. Distribution and origins of
polycyclic aromatic hydrocarbons (PAHs) in riverine, estuarine, and marine sediments in
Thailand. Marine Pollution Bulletin 52, 942-956.
Budzinski, H., Jones, I., Bellocq, J., Pierard, C., Garriques, P., 1997. Evaluation of sediment contaminant
by polycyclic aromatic hydrocarbons in the Gironde estuary. Marine Chemistry 58, 85-97.
Charlesworth, M., Service, M., Gibson, C.E., 2002. PAH contamination of western Irish Sea sediments.
Mar. Poll. Bull., 44: 1421-1434.
Davis, S.J., Unam, L., 1999. Smoke-haze from the 1997 Indonesian forest fires: effects on pollution
levels, local climate, atmospheric CO2 concentrations, and tree photosynthesis. Forest Ecology
and Management 124, 137-144.
De Luca, G.D., Furesi, A., Leardi, R., Micera, G., Panzanelli, A., Piu, P.C., Sanna, G., 2004. Polycyclic
aromatic hydrocarbons assessment in the sediments of the Porto Torres harbor (northern Sardinia,
Italy). Marine Chemistry 86, 15-32.
De Luca, G.D., Furesi, A., Micera, G., Panzanelli, A., Piu, P.C., Pilo, M.I., Spano, N., Sanna., 2005.
Nature, distribution and origin of polycyclic aromatic hydrocarbons (PAHs) in the sediments of
Olbia harbor (Northern Sardinia, Otaly). Marine Pollution Bulletin 50, 1223-1232.
Evans, K.M., Gill, R.A., Robotham, P.W.J., 1990. The PAH and organic content of sediment particle size
fractions. Water, Air, and Soil Pollution 51, 13-31.
Fang, M., Zheng, M., Wang, F., To, K.L., Jaafar, A.B., Tong, S.L., 1999. The solvent-extractable organic
conpounds in the Indonesia biomass burning aerosols characterization studies. Atmospheric
Environment 33, 783-795.
Ghosh, U., Gillette, J.S., Luthy, R.G., Zare, R.N., 2000. Microscale location, characterization, and
association of polycyclic aromatic hydrocarbons on harbor sediment particles. Environmental
Science & Technology 34, 1729-1736.
Hellou, J., Steller, S., Leonard, J., Langille, M.A., Tremblay, D., 2005. Partitioning of polycyclic
aromatic hydrocarbons between water and particles compared to bioaccumulation in mussels: a
harbour case. Marine Environmental Research 59, 101-117.
Holtvoeth, J., Kolonic, S., Wagner, T., 2005. Soil organic matter as an important contributor to late
quaternary sediments of the tropical West African continental margin. Geochimica et
Cosmochimica Acta 69(8), 2031-2041.

83

Horowitz, A.J., Elrick, K.A., 1987. The relation of stream sediment surface area, grain size and
composition to trace element chemistry. Applied Geochemistry, 2, 437-451.
IARC, 1987. IARC Monographs on the evaluation of carcinogenic risks to humans, Suppl. 7, overall
evaluations of carcinogenicity: an updating of IARC Monographs volumes 1 to 42, Lyon, IARC
Press.
ICES, 1997. Determination of polycyclic aromatic hydrocarbons (PAHs) in sediments: Analytical
methods. In Report of the ICES Advisory Committee on the Marine Environment, 1997. ICES
Cooperative Research Report 222, 118-124.
IPCC 2000. Land Use, Land-Use Change and Forestry, A Special Report of the IPCC. Cambridge
University Press, 377 pp (available online: http://www.ipcc.ch).
Kakareka, S.V., Kukharchyk, T.I., 2003. PAH emission from the open burning of agricultural debris. The
Science of the Total Environmental, 308: 257-261.
Kalbitz, K., Geyer, S., 2002. Different effects of peat degradation on dissolved organic carbon and
nitrogen. Organic geochemistry 33, 319-326.Agarwal, T., Khillare, P.S., Shridhar, V., 2006. PAHs
contamination in bank sediment of the Yamuna River, Delhi, India. Environmental Monitoring
and Assessment 123, 151-166.
Karickhoff, S.W., Brown, D.S. and Scott, T.A. (1979). Sorption of hydrophobic pollutants on natural
sediments. Water Research 13, 241-248.
Kim, G.B., Maruya, K.A., Lee, R.F., Lee, J.-H, Koh, C.-H., Tanabe, S., 1999. Distribution and sources of
polycyclic aromatic hydrocarbons in sediments from Kyeonggi Bay, Korea. Marine Pollution
Bulletin 38(1), 7-15.
King, A.J., Readman, J.W., Zhou, J.L., 2004. Dynamic behaviour of polycyclic aromatic hydrocarbon in
Brighton marina, UK. Marine Pollution Bulletin 48, 229-239.
Langmann, B., 2007. A model study of smoke-haze influence on clouds and warm precipitation formation
in Indonesia 1997/1998. Atmospheric Environment 41, 6838-6852.
Li, G., Xia, X., Yang, Z., Wang, R., Voulvoulis, N., 2006. Distribution and sources of polycyclic
aromatic hydrocarbons in the middle and lower reaches of the Yellow River, China.
Environmental Pollution 144, 985 993.
Luo, X-J., Chen, S-J., Mai, B-X., Yang, Q-S., Sheng G-Y., Fu, J-M., 2006. Polycyclic aromatic
hydrocarbons in suspended matter and sedimens from the pearl River Estuary and adjacent coastal
areas, China. Environmental Pollution 139, 9 20.
Magi, E., Bianco, R., Ianni, C., Di Carro, M., 2002. Distribution of polycyclic aromatic hydrocarbons in
the sediments of the Adriatic Sea. Environmental Pollution 119, 91-98.
Maruya, K.A., Risebrough, R.W. and Horne, A.J., 1996. Partitioning of polynuclear aromatic
hydrocarbons between sediments from San Fransisco Bay and their porewater. Environmental
Science and Technology 30, 2942-2947.
Maskaoui, K., Zhou, J.L., Hong, H.S., Zhang, Z.L., 2002. Contamination by polycyclic aromatic
hydrocarbons in the Jiulong River Estuary and Western Xiamen Sean, China. Environmental
Pollution 118, 109-122.
Meyer, P.A., 1994. Preservation of elemental and isotopic source identification of sedimentary organic
matter. Chemical Geology 114, 289-302.
Miguel, A.H., Kirchstetter, T.W., Harley, R.A., 1998. On-road emissions of particulate polycyclic
aromatic hydrocarbons and black carbon from gasoline and diesel vahicles. Environmental Science
and Technology, 32, 450-455.
Neff, J.M., 1979. Polycyclic aromatic hydrocarbon in the aquatic environment: sources, fates and
biological effects. Applied Science Publisher Ltd, Essex, UK, 262 pp.
Oen, A.M.P., Cornelissen, G., Breedveld, G.D., 2006. Relation between PAH and black carbon contents
in size fractions of Norwegian harbor sediments. Environmental Pollution 141, 370-380.
Page, S.E., Siegert, F., Rieley, J.O., Boehm, H-D. V., Jaya, A., Limin, S., 2002. The amount of carbon
released from peat and forest fires in Indonesia during 1997. Nature 420, 61-65.
Parameswaran, K., Nair, S.K., Rajeev, K., 2004. Impact of Indonesian forest fires during the 1997 El
Nino on the aerosol distribution over the Indian Ocean. Advances in Space Research 33, 10981103.
Pillon, P., Jocteur-Monrozier, L., Gonzalez, C., Saliot, A., 1986. Organic geochemistry of recent
equatorial deltaic sediments. Organic Geochemistry, 10, 711-716.
Prahl, F.G., Carpenter, R., 1983. Polycyclic aromatic hydrocarbons (PAH) phase associations in
Washington Coastal sediments. Geochimica et Cosmochimica Acta 47, 1013-1023.
Quah, E., 2002. Transboundary pollution in southeast Asia. World Development 30(3), 429-441.
Readman, J.W., Fillmann, G., Tolosa, I., Bartocci, J., Villeneuve, J.-P., Catinni, C., Mee, L.D., 2002.
Petroleum and PAH contamination of the Black Sea. Marine Pollution Bulletin 44, 48-62.

84

Requejo, A.G., Sassen, R., McDonald, T., Denoux, G., Kennicutt II, M.C., Brooks, J.M., 1996.
Polynuclear aromatic hydrocarbons (PAH) as indicators of the source and maturity of marine
crude oils. Organic Geochemistry 24(10/11), 1017-1033.
Ribeiro, C.A.O., Vollaire, Y., Sanchez-Chardi, A., Roche, H., 2005. Bioaccumulation and the effects of
organochlorine pesticides, PAH and heavy metals in the Eel (Anguilla anguilla) at the Camargue
Natura Reserve, France. Aquatic Toxicology 74, 53-69.
Rockne, K.J., Shor, L.M., Young, L.Y., Taghon, G.L., Kosson, D.S., 2002. Distribution sequestration and
release of PAHs in weathered sediment: the role of sediment structure and organic carbon
properties. Environmental Science and Technology 36, 2636-2644.
Ruttenberg, K.C., Go i, M.A., 1997. Phosphorus distribution, C:N:P ratios, and 13Coc in arctic,
temperate, and tropical coastal sediments: tools for characterizing bulk sedimentary organic
matter. Marine Geology 139, 123-145.
Sauer, T.C., Michel, J., Hayes, M.O., Aurand, D.V., 1998. Hydrocarbon characterization and weathering
of oiled intertidal sediments along the Saudi Arabian coast two years after the gulf war oil spill.
Environment International 24(1/2), 43-60.
Secco, T., Pellizzato, F., Sfriso, A., Pavoni, B., 2005. The changing state of contamination in the Lagoon
of Venice. Part 1: organic pollutants. Chemosphere 58, 279-290.
See, S.W., Balasubramanian, R., Rianawati, E., Karthikeyan, S., Streets, D.G., 2007. Characterization and
source apportionment of particulate matter  2.5 m in Sumatra, Indonesia during a recent peat
fire episode. Environmental Science and Technology, 41, 3488-3494.
Simpson, C.D., Harrington, C.F. and Cullen, W.R., 1998. Polycyclic aromatic hydrocarbon contamination
in marine sediments near Kitimat, British Columbia. Environmental Science and Technology 32,
3266-3272.
Soclo, H.H., Garrigues, PH., Ewald, M., 2000. Origin of polycyclic aromatic hydrocarbons (PAHs) in
coastal marine sediments: case studies in Cotonou (Benin) and Aquitaine (France) areas. Marine
Pollution Bulletin 40(5), 387-396.
Tolosa, I., de Mora, S., Sheikholeslami, M.R., Villeneuve, J.-P., Bartocci, J., Cattini, C., 2004. Aliphatic
and aromatic hydrocarbons in coastal Caspian Sea sediments. Marine Pollution Bulletin 48, 44-60.
Wang, X. -C, Zhang, Y.-X., Chen, R.F., 2001. Distribution and partitioning of polycyclic aromatic
hydrocarbon (PAHs) in different size fractions in sediments from Boston Harbor, United States.
Marine Pollution Bulletin 42(11), 1139-1149.
Wang, Z., Fingas, M., Page, D.S., 1999. Oil spill identification. Journal of Chromatography A 843, 369411.
Yang, Y., Ligouis, B., pies, C., Grathwohl, P., Hofmann, T., 2008. Occurrence of coal and coal-derived
particle-bound polycyclic aromatic hydrocarbons (PAHs) in a river floodplain soil. Environmental
Pollution 151, 121-129.
Yim, U.H., Hong, S.H, Shim, W.J., Oh, J.R., Chang, M., 2005. Spatio-temporal distribution and
characteristics of PAHs in sediments from Masan Bay, Korea. Marine Pollution Bulletin 50, 319326.
Yunker, M.B., Macdonald, R.W., Vingarzan, R., Mitchell, H., Goyette, D., Sylvestre, S., 2002. PAHs in
the Fraser River basin: a critical appraisal of PAH ratio as indicators of PAH source and
composition. Organic Geochemistry 33, 489-515.
Zakaria, M.P., Takada, H., Tsutsumi, S., Ohno, K., Yamada, J., Kouno, E., Kumata, H., 2002.
Distribution of polycyclic aromatic hydrocarbon (PAHs) in rivers and estuaries in Malaysia: a
widespread input of petrogenic PAHs. Environmental Science and Technology 36, 1907-1918.
Zech, W., Senesi, N., Guggenberger, G., Kaiser, K., lehmann, J., Miano, T.M., 1997. Factors controlling
humification and mineralization of soil organic matter in the tropics. Geoderma 79, 117-161.
Zhang, Z.L., Hong, H.S., Zhou, J.L., Yu, G., 2004. Phase association of polycyclic aromatic
hydrocarbons in the Minjiang River Estuary, China. The Science of the Total Environment 323,
71-86.
Zhou, J., Wang, T., Huang, Y., Mao, T., Zhong, N., 2005. Size distribution of polycyclic aromatic
hydrocarbons in urban and suburban sites of Beijing, China. Chemosphere 61: 792-799.
Zhou, J.L., Maskaoui, K., 2003. Distribution of polycyclic aromatic hydrocarbons in water and surface
sediments from Daya Bay, China. Environmental Pollution 121, 269 281.

85

V.

POLYCYCLIC AROMATIC HYDROCARBONS IN SURFACE


WATERS OF THE SIAK RIVER, ITS ESTUARY AND THE
COASTAL AREAS OF RIAU PROVINCE, INDONESIA:
DISTRIBUTION AND SOURCES

M. Lukman*, W. Balzer*, I. Becker*, C. Jose**, J. Samiaji**


*Marine Chemistry Working Group, Department of Biology/Chemistry, University of Bremen,
Leobener Strasse, 28359 Bremen, Germany
** University of Riau, Jl. Simpang Baru Panam, 28293 Pekanbaru, Riau, Indonesia

Abstract

Polycyclic aromatic hydrocarbons (PAHs) were studied both in water solution and in suspended
particulate matter (SPM) of the Siak river, including its estuary and the coastal areas of
Sumatra, Indonesia. Concentration of dissolved 16 PAHs according to the EPA selection
(PAHs) varied greatly in river, estuary and coastal water samples ranging from 129 to 5140 ng
L-1 (median = 824 ng L-1), 320 to 619 ng L-1 (m = 385 ng L-1), and 121 to 130 ng L-1 (m = 130
ng L-1), respectively. The PAHs in the SPM samples for those three aquatic systems ranged
from 1475 to 59050 ng g-1 (median = 5286 ng g-1), 156 to 7669 ng g-1 (m = 758 ng g-1), and 326
to 10234 ng g-1 (m = 1572 ng g-1), respectively. Two- to four-ring PAHs were predominant in
the water and particulate samples. PAH source apportionment was estimated using specific
isomer ratios in combination with the ratio of low and high PAH molecular weights. The
signatures of petrogenic and pyrogenic sources were found in all water systems, but biomass
burning signatures were more prominent. This might be evidence for palm-oil plantation and
peatland burnings which occurred frequently in the region since the last decades. The chemical
signatures of petroleum spillages were mainly observed around the River mouth and in the coast
north of the River, as well as close to the capital city of Pekanbaru, the industrial area Perawang
and Dumai city.

Keywords: PAHs; Suspended Particulate Matter; Water, Siak River; Estuary; Riau Coast

MANUSCRIPT PREPARED TO BE SUBMITTED

(TABLES MENTIONED IN THIS MANUSCRIPT ARE PROVIDED IN APPENDIX 5: p.134-138)

86

5.1.

Introduction
Contamination of polycyclic aromatic hydrocarbons (PAHs) in the river, estuary and

coastal waters has attracted high attention, not only due to their toxic and carcinogenic effects,
but also to their persistence, ubiquity and bioaccumulation. These multiply the potential risks
for human exposure (e.g. Shou et al., 1996; Bofetta et al., 1997; Hussain et al., 1998; Rybicki et
al., 2006; Singh et al., 2007). Therefore, in many countries allowable concentration of PAHs are
strictly regulated for food and drinking water. PAHs refer to a group of arene compounds built
up by two or more fused aromatic benzene rings in various planar configurations of molecular
structures containing no heteroatom (Neff, 1979). These compounds have both natural and
anthropogenic origins. The latter refer to those generated by combustion processes, and are the
most significant due to various possible sources, mainly 1) biomass burning, for example slashand-burn agriculture (e.g. Kakareka & Kukharchyk, 2003; Oanh et al., 2005), 2) forest/swamp
fires (e.g. Olivella et al., 2006; Vila-Escal et al., 2007), and 3) the combustion of fossil fuels
(e.g. de Abrantes et al., 2004, Zielinska et al., 2004). Furthermore, crude oil and its related byproducts (hereafter called petroleum) contain a significant aromatic fraction in which PAHs
constitute the toxic portion of crude oil (Requejo et al., 1996).
The Siak River in Riau province, Sumatra, Indonesia is pivotal for the local populations
for daily life and drinking water. Together with its estuary and the adjacent coastal areas, it is a
potent fishing ground supporting ca. 2.5 million urban and rural inhabitants along the basin.
They also become a main shipping lane for economic importance mainly oil industry. The Siak
aquatic system significantly discharges humic-enriched blackwater plumes into the Malacca
Strait causing ocean acidification (Baum et al., 2007; Siegel et al., 2009). There is currently no
information available on the concentrations and distribution of organic contamination,
particularly PAHs in these aquatic systems. On the other hand, relatively high level of PAHs
may be anticipated for this region owing to widespread common agricultural/biomass burnings
and severe outbreaks of forest/peat swamp fires ever since the El Nino 1997 episode in Sumatra
(e.g. Page et al., 2002). During a severe fire episode in 2005 in Sumatra See et al. (2007)
observed around Dumai (north of the Siak River) and in Riau the capital city of Pekanbaru that
the total 16 PAHs of the US Environmental Protection Agency (the 16 PAHs) priority pollutant
emitted from peatland burnings was ca. 561 ng m-3 and ca. 135 ng m-3, respectively. The latter
was even higher than typical values (ca. 116 ng m-3) of polluted urban areas of Beijing, China
(Zhou et al., 2005).
This study aims to examine the concentration, distribution and sources of PAHs in the
water solution and in the suspended particulate matter (SPM) of the Siak River, its estuary and
the Riau coastal areas. PAHs are restricted to the 16 PAHs, since these compounds have been
widely used in the framework of environmental assessment.

87

5.2.

Materials and Methods

5.2.1.

Study Areas and Sampling Locations


The Siak River stretch ca. 300 km, including several tributaries: the Tapung Kanan,

Tapung Kiri (the upstreams) and Mandau Rivers (Fig. 1). Along with its estuary, the Siak
catchment extents of about 1 million hectares, 10% of the total land area of Riau Province
(BPDAS, 2004), draining largely low-laying land characterized by large peatland, palm-oil
plantations and swamp forests. The Siak receives also discharges from many artificial canals.
The highest rainfall is usually between October and December (BPDAS, 2004).

S317
S229

S318

S225

S227

Dumai

Malacca Strait

S228
S324
S226

S267

S266

Bengkalis
Bukit Batu River

Sumatra

S124

S233
S230
S251

Siak Kecil River


S132, S133
S130, S126
S314

INDONESIA

S305

S316

S125
S128
S231
S134
S231
S139
Panjang
Strait

S 305
Upstream Segment

Perawang
S291

Tapung Kiri River


S17
S102
S205

S 301
S216

Siak Sri Indrapura


S307

S9
S 115
S 370

S10
S105
S272
S218

Pekanbaru
B

Fig. 1. Map of the sampling locations for water (triangles) and suspended particulate matter (cycles). The
Siak River was sampled for three segments (dash-line boxes): (A) upstream; (B) Pekanbaru; (C)
Perawang.

During the three-year study (2004 2006), a total of 30 sampling stations along the river,
the estuary and the coastal areas were selected for particulate and 10 stations for dissolved PAH
assessments. Due to the large spatial scale, we selected sampling locations which represent a
variety of anthropogenic pressures which follow. First, the Siak River was sampled for (i) the
upstream region of the Siak including the Tapung Kanan and Tapung Kiri Rivers, which are
representative of agricultural sections surrounded by large palm-oil plantations, ii) the city
region around the capital Pekanbaru, and (iii) the downstream region enclosing the industrial
area of Perawang and the population centre of Siak Sri-Indrapura. Secondly, the estuary zone

88

was sampled following the salinity gradient from low (0 10 practical salinity unit, psu), over
medium (10 25 psu) and to high (> 25 psu) salinity. Finally, the coastal area was sampled
from the oil-industry city of Dumai in the north to the coastal channel between Bengkalis and
Sumatra Islands, and to Panjang Strait in the south of the Siak river, where several offshore oilridges are located.
5.2.2.

Sample Collection and Treatments


Five to ten liters of surface water (from circa 1 meter depth) were collected with a Niskin

bottle, then poured in a pre-cleaned glass bottle, sealed, and stored until further treatment in the
laboratory. Within 8 hours after sampling, SPM was obtained by filtering one to five liters of
river or estuarine water through pre-combusted (at 400 oC for 4 hours) Whatman GF/F filter in
triple sets of 250-ml glass containers. The water was mixed thoroughly prior to filtration. For
coastal waters, the volume of filtration reached up to 10 L due to the low SPM concentration.
Then, the filters were wrapped in aluminum foil, and stored at -20 oC until further analysis. The
filtrate was pre-treated for dissolved PAH extraction (see 5.2.4.).
5.2.3.

Extraction & Work-Up Procedures for particulate PAH


Prior to extraction, the filters containing the SPM were air dried in a clean-bench for 24

hours. Surrogate deuterated PAH standards including d10-phenanthrene (PHEN D10), d10fluoranthene (FLA D10) and d12-perylene (PERY D12) were spiked onto the filter. The filters
were then Soxhlet-extracted with an SoxTec HT6 extraction system using 50 mL of
acetone/hexane (1/9; v/v) for 6 hours. The extracts were reduced to ca. 1 mL with a rotary
evaporator, and were subjected to a clean-up process in a 1:2 (w/w) deactivated alumina (10%) /
silica (3%) column. For elution the extract, 40 mL of a degassed 3/7 (v/v)
dichloromethane/hexane mixture was aspirated through the column. The extraction solvents and
the clean-up systems are taken from the ICES Method (ICES, 1997). The clean extract was
again reduced to ca. 1 mL by the evaporator, and further concentrated until it was dry under a
gentle stream of purified N2. Finally, 500 L of acetonitrile was added to the sample for HPLC
analysis.
5.2.4.

Solid Phase Extraction (SPE) system for dissolved PAH


Dissolved PAHs were obtained from the filtrate, by passing the filtered water through a

SPE cartridge (Chromabond NH2/C18) obtained from Macherey-Nagel GmbH (Dueren,


Germany). Methanol (2% of the filtrates volume) was added to 0.5 to 1 L of the filtrate in order
to enhance the recovery. The surrogate standards (FLA D10 and PERY D12) were spiked to
determine the procedural efficiency and reproducibility, and to enable the data correction from
any losses during extraction and work-up procedure. Prior to extraction, the cartridges were
sequentially conditioned with 5 mL hexane and 10 mL methanol, then washed with ca. 100 mL

89

Milli-Q water. Thin film of water was left above the phase to avoid drying. The sample was
then aspirated through the cartridge at flow rate of ca. 5 mL/min controlled by adjusting the
pressure. After all of the sample was emptied, the cartridge was again washed with Milli-Q
water, and a little bit of water was left on top of the cartridge to avoid drying. Then, the column
was firmly wrapped with foil, sealed with a paraffin film, and stored at -4 oC until analysis.
The steps of the work-up procedure for the retained PAHs on the cartridge were relatively
simple compared to those for the SPM. Prior to elution, the cartridges were gently dried by a
stream of N2. Then, the PAHs were eluted with 4 mL of dichloromethane (DCM) (repeated
three times) at flow rate of 5 mL/min. The use of DCM for elution of PAH for SPE system has
been suggested (Sargenti & McNail, 1998). The elutes were reduced to a dry state by by N2
blow down, then 300 L of acetonitrile was added prior to analysis with high performance
liquid chromatography (HPLC).
5.2.5.

Determination of PAH by HPLC UV/FLD


Baseline separation and quantification of the 16 PAHs were performed using high

performance liquid chromatography (HPLC LKB 2249 Broma) with a reverse phase RP-C18
HPLC column (Bakerbond 250 x 4.6 mm, 5m obtained from J.T. Baker Inc, Phillipsburg,
USA), detected by a couple of ultraviolet (UV) and fluorescence detectors (FLD). The 16
PAHs include naphthalene (NAPH), acenaphtylene (ACYN), acenapthene (ACEN), fluorene
(FLU), phenanthrene (PHEN), anthracene (ANTH), fluoranthene (FLA), pyrene (PYR),
benzo(a)anthracene

(BaA),

chrysene

(CHRY),

benzo(b)fluroanthene

(BbFLA),

benzo(k)fluoranthene (BkFLA), benzo(a)pyrene (BaP), dibenzo(a,h)anthracene (DANTH),


benzo(g,h,i)perylene (BPERY), and indeno(1,2,3-c,d)pyrene (IPYR). The mobile phase was a
combination of an isocratic and a linear binary-gradient elution of acetonitrile-ACN/water,
programmed from 55% to 100% ACN at a constant flow rate of 1 ml/min for 45 minutes.
The quality of the analytical method was tested and confirmed. A mixture of 16 certified
PAH standards and the three surrogate deuterated standards (obtained from Dr. Ehrenstorfer
GmbH, Augsburg, Germany) were used to identify and quantify the analytes. HPLC-grade
acetonitrile and analytical-grade quality solvents (acetone, dichloromethane and hexane)
obtained from Fischer Scientific were used throughout the analytical procedures. Procedural
efficiency experiments for the SPE system were carried out prior to the sampling sequences
using a spiked-mixture of 16 PAHs and Milli-Q quality water. The individual recovery rates
were greater than 74%, except for the naphthalene (47%) and benzo(a)pyrene (58%).
Procedural efficiency, reproducibility and data correction from losses during the extraction and
work-up procedures were confirmed by the surrogate deuterated PAHs (PHEN D10, FLA D10,
and PERY D12) which had been spiked into the sample prior to the extraction. The mean
percent recovery standard deviation of the surrogates for the SPE were 91.7% 6.98% (FLA

90

D10) and 88.8% 9.9% (PERY D12), while for the SPM were 95.3% 10.1%, 96.8% 9.8%,
and 95.7% 8.9% for PHEN D10, FLA D10 and PERY D12, respectively.

5.3.

Results and Discussion

5.3.1.

Dissolved PAHs
The concentration of total PAHs (PAHs) in the dissolved phase of the riverine,

estuarine and coastal waters ranged from 129 to 5140 ng L-1 (median = 824 ng L-1), 320 to 619
ng L-1 (m = 385 ng L-1), and 121 to 130 ng L-1 (m = 130 ng L-1), respectively (Table 1). The
PAHs decreased about three-folds towards the coastal waters. The concentration of PAHs in
the rivers were markedly different between the stations, but not in the estuary and the coast. The
highest concentration in river water was detected in confluence of the Siak River and the
Mandau River (S301), the industrial area of Perawang (S305), and of the capital city of
Pekanbaru (S291). The high concentration at S301 was party attributed to the input of the black
water of Mandau River which contained relatively high PAH (unpublished data). The Mandau
River drains the water mostly from the catchment areas surrounded by large peat and plantation
areas. Increased material transport from surrounding peatland to the Mandau River has been
acknowledged for dissolved organic matter (e.g. Baum et al., 2007). The PAH concentration at
all sampling sites was considered high in comparison to aquatic systems around industrial and
urban centres in some Asian and European cities (Table 3) such as rivers in Tianjin China (e.g.
Shi et al., 2005), Qiantang River China (e.g. Chen et al., 2007), Seine River France (e.g.
Fernandes et al., 1997). These Siak concentrations were orders of magnitude higher than waters
affected by PAH after forest fires (e.g. Olivella et al., 2006; Vila-Escal et al., 2007), and fewfolds above the concentration of PAHs in waters affected by oil spills (e.g. Gonzles et al., 2006
and references therein).
As to individual PAH, the relative composition of the PAHs in all water systems was
generally dominated by NAPH, ACYN, PHEN and FLA. These compounds are main
constituents of biomass burnings such as wood (e.g. Oanh et al., 2005), agricultural (e.g.
Kakareka, & Kukharchyk, 2003), forest fires (e.g. Vila-Escale et al 2007) and Sumatra peatland
burnings (See et al., 2007). However, high level of NAPH and FLA is also produced from
petroleum (diesel) combustion (e.g. de Abrantes et al., 2004), and was detected in water
polluted by oil spills (e.g. Baars et al., 2002). The ring composition pattern shows that 2-, 3- and
4-ring are dominant in all sampling stations, while 5- and 6-rings are less significant. NAPH
significantly enriched in the Pekanbaru (S291) which stem most probably from the city
wastewater discharges and ports. Increase in 2-ring PAHs is typically observed in the water
around industrial areas and ports e.g. Qiantang River in China (Chen et al., 2007).

91

5.3.2.

PAHs in the SPM


The PAHs in the SPM samples from the Siak River ranged from 1475 to 59050 ng g-1

d.w (Table 2a). Very high contents (>15000 ng g-1 d.w) were detected in March 2004 in all river
segments (Fig. 2). The content of the PAHs generally decreased ever since, except in the
upstream segment where increased content was observed in September 2004. Monthly samples
(see: inset Fig. 2) taken at the upstream segment in 2004 showed that the contents of PAHs
remained very high until the wet season (September). It suggests that the upstream segment was
a potential repository for the particulate PAHs. Since the upstream segment is surrounded by
vast palm-oil plantations having experienced frequent burnings, it is expected that the high
content of PAHs was attributed to the burnings. Likewise the dissolved PAHs, NAPH, ACYN,
PHEN, FLA and PYR generally dominated the relative composition of individual PAH in all
segments.

70000

Wet Season

Upstream
60000

60000

PAHs (ng/g d.w.)

PAHs (ng/g d.w.)

Dry Season

70000

50000
40000

50000
40000
30000
20000
10000
0

30000

Mar 04 May 04 Jun 04 Aug 04 Sep 04 Jul 05

20000
10000
0

Mar 2004

Sep 04
Upstream Tributaries

Jul 05
Pekanbaru

Mar 06
Perawang

Fig. 2. Seasonal distribution of particulate PAHs from Siak River waters. Inset: monthly
variation of the particulate PAHs from the upstream segment.
Fig. 3 illustrates the relative composition of the ring group PAH between dry and wet
seasons. It shows that the profile of the ring-group composition in the upstream part is similar
with 4-ring PAHs (i.e. FLA, PYR) being predominant. FLA and PYR accounted for ca. 48%
and ca. 24% of the individual composition. It confirms that PAHs stemmed from the same
sources throughout the year, most possibly from the burnings from palm-oil plantation and

92

peatland. The higher content in the wet season (2004) than in the dry season (2005) might
suggest that land runoffs (land-water interfaces), by which burning residues were flushed out to
the River, were more significant sources of particulate PAHs than atmospheric deposition (airwater interfaces). However, there is a clear seasonal modification of the ring group composition
in the city of Pekanbaru and the industrial area of Perawang. The 3-ring PAHs (ACYN and
PHEN) were very dominant in the wet season in both segments. ACYN and PHEN contribute
ca. 40% and 20% of the individual relative composition, respectively. Instead of biomass
burnings, the PAHs were possibly derived from observable oil discharges during the samplings.
Oil discharges to the aquatic systems are not strictly monitored. In the dry season the
composition profile was shared among 3, 4, 5 and 6 rings. It suggests many various sources of
PAHs, possibly including urban and industrial wastewater runoffs and urban atmospheric
fallouts in addition to smokes from the plantation and peatland burnings during 2005.

Relative Composition (%)

90

The Siak River: Dry Season (Jul 2005)

80

Upstream S205

70

Pekanbaru S272
Perawang S216

60
50
40
30
20
10
0

Relative Composition (%)

90

The Siak River: Wet Season (Sep 2004)

80

Upstream S102

70

Pekanbaru S105

60

Perawang S115

50
40
30
20
10
0

2 rings

3 rings

4 rings

5 rings

6 rings

Fig. 3. The relative composition of the ring groups of the PAHs in the Siak River during (a) dry and (b)
wet seasons. The dash lines illustrate the ring-group profiles.

93

In the estuary, the PAHs in the SPM samples vary also greatly ranging between 156 and
7669 ng g-1 d.w. with a median of 758 ng g-1 d.w. (Table 2b). The PAHs was lower by an
order of magnitude than those of the River. The highest level (7669 ng g-1 d.w.) was observed at
S134 around Sungai Apit in the mouth of the River. The high content was partly due to another
potential input from the Channel connecting the blackwater Siak Kecil River and the Siak River.
There is no clear trend of distribution towards the coast (Fig. 4). As to individual PAHs, the
relative composition of PAHs was shared among NAPH, ACYN, PHEN, FLA, and DANTH. In

12000
10000
8000
6000
4000
2000
0
S139
S140
S137
S134
S130
S126
S132
S124
S133
S125
S128
S225
S226
S227
S228
S229
S267
S266
S233
S251
S230
S231
S232

Particulate PAHs (ng/g d.w.)

general, the 4-ring PAHs are dominant, followed by the 3- and 5-ring PAHs (Fig. 5).

The Siak Estuary (Sep 04)

The Riau Coast (Jul 05)

Fig. 4. Spatial distribution of particulate PAHs in the Estuary and the Coast.

The PAHs in the coastal SPM was higher than those of the estuary extending from 326
to 10234 ng g-1 d.w., with a median of 1572 ng g-1 d.w. (Table 2c). The high values at S267 and
S266 were possibly due to an accumulation of PAH-rich particle plumes from blackwater
streams in the area including the Siak River, the Siak Kecil River and Bukit Batu River. In
addition, significant plumes of humic substances from those three blackwater streams into the
Malacca Strait have been recognized (Siegel et al., 2009). The relative composition of the ring
group PAHs was dominated by 4-ring PAH mainly FLA and PYR (Fig. 6). The high portion of
4-ring PAH were also shared by the estuarine SPM and the upstream segment of the River SPM.
It might indicate the common sources (i.e. palm-oil plantation and peatland burnings). The
profile of coastal plume and Malacca strait (at blue water stations) is slightly different in which
the relative composition of 6-ring PAHs increased in the Malacca Strait (Fig. 6a). It is most
likely due to diesel and gasoline combustion derived from vast ships along the Strait. Diesel and

94

gasoline combustion emit relatively high proportion of high molecular PAHs (e.g. Riddle et al.,
2007). NAPH was higher in the coastal than in the estuarine SPM. Increase in 2-ring PAH in
the coast might partly stem from oil discharges since the coastal areas are being the main
transport pathway for oil cargos. It suggests a chronic petroleum contamination around the river
mouth and the coast. Furthermore, a proportion of 2- to 3-ring compounds were more
pronounced along the coast north of the river mouth (including Bengkalis and Dumai Strait)
than the coast south of the river mouth or Panjang Strait (Fig. 6b). It suggests that the north of
the river mouth was more affected by petroleum sources.

A
45

Dry Season (July 2005)

Relative Composition (%)

40

The Estuary
The Coast

35

The Malacca Strait


(S229)

30
25
20
15
10
5
0

B
Relative Composition (%)

45
40

The Coast
Dry Season (July 2005)

Northern part
Southern part

35
30
25
20
15
10
5
0

2 rings

3 rings

4 rings

5 rings

6 rings

Fig. 5. The relative composition of the ring groups of the PAHs. The dash lines illustrate the ring-group
profiles. (a) the profiles in the estuary (median value, n=10), the coast (median value, n=11) and the
Malacca Strait (at S229), (b) the profiles in northern and southern (of the Siak River mouth) parts of the
coast.

On a volume basis, the concentration of particulate PAHs ranged from 62.5 to 694 ng
-1

L , 25.4 to 293 ng L-1, and 11.7 to 147 ng L-1 in the River, the estuary and the coast,

95

respectively. The concentration decreased towards the coastal waters similar to those of
dissolved PAHs (Fig. 6). But, the volume concentration of the particulate PAHs was lower by
up to one order magnitude compared to the dissolved PAHs. The decrease from the river to the
coast suggests entrapment and dilution effect of sea water. In comparison to other river, estuary
and coastal systems in Asia and Europe (Table 3), the content of PAHs in this study was
comparable to those rivers in industrial city of Tianjin, China (Shi et al., 2005), Seine River in
France (Fernandes et al., 1997), and Elbe River at Dessau (Heemken et al., 2000). Even, they
were higher than the Pearl River and Estuary, China (Luo et al., 2006).
800

Particulate 6PAHs (ng/L)

600

400

200

Siak River

Sak Estuary

Riau Coast

Fig. 6. Concentration of the particulate PAHs on volume base (ng/L) in Siak River (n=12), estuarine
(n=11) and coastal waters (n=12).

5.3.3.

Distribution Coefficient of PAHs between SPM and Water Solution


Distribution of PAHs between particulate and dissolved phases is driven by complex

particle-water interactions which involve physico-chemical properties of individual compounds


and the water properties especially particulate and dissolved organic matter as geosorbents for
hydrophobic pollutants (Schlautmann & Morgan 1993; Christl & Kretzschmar, 2001; Weber Jr.
et al., 1991). In order to understand factor controlling PAH phase distribution in the studied
areas, distribution coefficient (KD, mL/g), that refers to as the ratio of PAH content (ng/g) in the
SPM to its corresponding concentration (ng/mL) water solution, was examined for NAPH,
PHEN, FLA, BaP, BPERY as these compounds linearly correlated with PAH16, and represent
the ring groups. The KD values generally ranged from 2 to 5 on the logarithmic scale. The KD

96

value in the coastal water is the highest. The partition of PAH is expected to increase as salinity
increases (Brunk et al., 1997; Turner & Rawling, 2001; Tremblay et al., 2006). However, the KD
values in the estuary were lower than those of Siak River (Fig. 7a). It means that increased
salinity did not automatically increase the sorption coefficients of PAHs suggesting other
factors enhancing their solubility. Dissolved organic matter (DOM) in the water solution
enhances the solubility of PAHs (Liu & Amy, 1993; McGroddy & Farrington, 1995).

6.0

7.0
NAPH
PHEN

4.0

FLA

Log KOC

5.0

Log KD

Samples from Mar 2006 (mean values)

8.0

6.0
River (n=5)
Estuary (n=3)

5.0

Coast (n=3)

BaP

3.0

BPERY

Koc = Kow

4.0
3.0

2.0

River

Estuary

3.0

Coast

4.0

5.0

6.0

7.0

Log KOW

8.0
Log KOC = -0.137*DOC + 6.9
R = 0.62

log KOC

7.0
6.0
5.0
4.0

BaP

3.0
0.0

5.0

10.0

15.0

20.0

DOC (mg/L)

Fig. 7. (a) Distribution coefficient (KD) of selected PAHs in the Siak River, its estuary and the Riau coast
(mean log values); (b) a cross plot of organic-carbon normalized distribution coefficient (KOC) vs.
compounds octanol-water distribution coefficient (KOW, the values taken from Williamson, et. al., 2002),
a measure of hydrophobicity; (c) a cross plot of Log KOC vs. DOC for BaP suggesting the role of
dissolved organic matter in enhancing the solubility of PAHs. DOC data was provided by Dr. A. Baum
and Dr. T. Rixen from ZMT Bremen, Germany.

To understand the effect of particulate organic matter (POM) on the distribution, organic
carbon normalized partition coefficient (KOC = KD /fOC) is evaluated, where fOC is the SPM
fraction of organic carbon. The results showed that the KOC was variable, and showed no
correlation between KD and fOC. It means that the KD is not dependent on the content of the
organic matter, but on the quality. The POM in the SPM had different affinity for PAH. Many
studies pointed out that in natural water the KOC of PAH compounds is higher than its
hydrophobicity (KOW) due to strong sorption of PAHs onto the combustion-derived organic

97

matter such as soot or black carbon (e.g. Jonkers & Koelmans, 2002; Gustafsson et al., 1997).
However, although the Siak water systems are severely affected by biomass burnings, the KOC
values is not all exceed the KOW (Fig. 7b). It might suggest a significant role of DOM enriched
in colloidal fraction from leaching peatland that maintained the PAH in the dissolved phase as
shown by an inverse correlation between KOC and DOC (Fig. 7c). Also, it might suggest that
the PAHs stemmed from non-combustion sources (e.g. Zhu et al., 2008).
5.3.4.

Source apportionment
The identification of significant sources for anthropogenic PAH contamination plays a

critical role in water pollution control strategies, but contain a great challenge. This is due to the
fact that PAHs particularly the parent PAHs, have both petroleum and combustion sources,
which coincide in the aquatic environment. Therefore, we applied several ratios which are
commonly used to distinguish petroleum from pyrolytic sources for the particulate PAHs. Those
ratios can be classified in two categories: (i) the ratio of the sums of low-molecular to the sum
of high-molecular weight substances ( LMW/HMW); and (ii) the concentration ratios of
specific

parent

isomers

namely

ANTH/(ANTH+PHEN),

FLA/(FLA+PYR),

BaA/(BaA+CHRY) (Neff 1979; Budzinski et al., 1997; Yunker et al., 2002; de Luca et al.,
2005; Soclo et al., 2000). Petrogenic origin is typically characterized by a high proportion of
LMW PAHs particularly those of 2-3 ring compounds. Therefore, LMW/HMW >1 indicates
petroleum sources. The specific isomer ratios as mentioned earlier are perceptive to distinguish
different combustion from petroleum sources due to different stability of the isomers with
increasing temperature during pyrolysis. PAHs from combustion processes have typical value of
ANTH/(ANTH+PHEN) > 0.1, FLA/(FLA+PYR) > 0.5, and BaA/(BaA+CHRY) > 0.35. On the
other hand, petrogenic PAHs have typical values of those corresponding isomeric ratios of <
0.1, < 0.4, and < 0.2, respectively (Yunker et al., 2002). Any value falling between those
determining values is usually considered as mixtures of petroleum and combustion sources.
The LMW/HMW ration ranged from 0.192 to 4.84, 0.137 to 11.4, and 0.26 to 2.54 in
the river, the estuary and the coast, respectively. Those ratios clearly suggest that in all studied
areas the PAHs stemmed from both petrogenic and pyrogenic origins. The influence of
petrogenic sources was especially pronounced at Pekanbaru and Perawang in the River, and in
the northern part of the coast including the blue water stations. The signature of petroleum
sources in all sampling stations was also detected in the ANTH/(ANTH+PHEN) with the mean
ratio of 0.02 0.01 (mean SD), 0.02 0.01, and 0.03 0.01 in the River, the estuary and the
coast. The petrogenic PAHs were assumed to come from vessel discharges, port activities,
industrial and urban runoffs. Thin oil films on the water surface were often observed during the
sampling campaigns.

98

On the other hand, the ratios FLA/(FLA+PYR) and the BaA/(BaA+CHRY) in all studied
areas

show important pyrogenic sources, particularly biomass burning and petroleum

combustion. The FLA/(FLA+PYR) were 0.53 0.14 (mean SD), 0.49 0.18, and 0.39 0.10
in the River, the estuary and the coast. The respective values of the BaA/(BaA+CHRY) was
0.44 0.16, 0.38 0.21, and 0.48 0.11. As been applied in the many studies, cross plots of
FLA/(FLA-PYR) vs. ANTH/(ANTH+PHEN) and BaA/(BaA+CHRY) shows that pyrogenic
the dominant sources (Fig. 8). The ratios in the river water point more to biomass burning
signatures which were most possibly palm-oil plantation and peatland burnings, while mixtures
of sources between biomass burnings, petroleum combustion, and oil discharges were assigned
in the estuary and the coast.
Petroleum

Biomass Combustion

1.0
Peat Burning
Dumai

0.8

light-duty diesel
vehicles

Agricultural
debris

0.7
SPM after forest
fires

0.6

Combustion

Weat grasses

0.5
Rice grasses

0.4
Oil spills "Tanker Charcoal
Enrika"

0.3
0.2

Bituminous coal

Petroleum

BaA / (BaA + CHRY)

0.9

Estuary S137

0.1

Alaskan crude oil

Estuary S139

0.0
0.4

0.3

Combustion

Agricultural
debris

0.3

Charcoal
Oil spills
"Tanker Enrika"

0.2

Weat grasses

0.2

Rice grasses

Peat Burning
Dumai
Alaskan crude
oil

0.1
0.1

Petroleum

ANTH / (ANTH + PHEN)

Bituminous coal
0.4

light-duty diesel
vehicles

0.0

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

FLA / (FLA + PYR)

Fig. 8. Cross plots of isomer ratios for the source apportionment of PAHs in the Siak River (unfilled
triangles), the estuary (black squares), and the coast (black cycles) (discrimination values dash lines after Yunker et al., 2002). For comparison, literature data illustrates petrogenic and pyrogenic sources
(grey cycles). References: petrogenic sources: Alaskan Crude Oil (Requojo et al. 1996); Oil-spill Tanker
Enrika (Baars et al., 2002); pyrogenic sources: Bituminous coal (Liu et al 2009); Charcoal (Oanh et al.,
1999); Agricultural debris (Kakareka & Kukharchyk, 2003); Rice & Weat grasses (Jenkins et al.,1996);
light-duty diesel vehicle (de Abrantes et al 2004); Sumatra peatland burning-Dumai (See et al., 2007);
SPM after forest fire (grey triangle: Vila-Escale et al 2007).

99

5.4.

Conclusion
The Siak River, its estuary and the coastal waters of Riau Province are highly

contaminated by PAHs. Enrichment of PAHs was centred at waters around residential and
industrial areas. Two- to four-ring PAHs were dominant in the both dissolved and particulate
phases. Source apportionment indicated petroleum and biomass burning sources in all studied
areas. This reflects the decisive effects of biomass burnings and swamp/forest fires, as well as
oil discharges from vessel, port activities, urban and industry.

Acknowledgments
This study is part of the German-Indonesian SPICE Project Cluster 3.1., funded by the
Federal German Ministry of Education, Science, Research and Technology (BMBF, Bonn), and
supported by the German Academic Exchange Service (DAAD). We would like to show our
gratitude to scientists and students from University of Riau, Indonesia, the crew of RSV
Senangin, and to SPICE German colleagues for their contribution during sampling campaigns
and discussion. Last but not least, we thank all reviewers for their constructive critics and
comments.

References
Baumard, P., Budzinski, H., Garrigues, P., Sorbet, J.C., Burgeot, T., Bellocq, J., 1998.
Concentrations of PAHs (Polycyclic Aromatic Hydrocarbons) in various marine
organisms in relation to those in sediments and to trophic level. Marine Pollution
Bulletin, 36(12): 951-960.
Budzinski, H., Jones, I., Bellocq, J., Pierard, C., Garriques, P., 1997. Evaluation of sediment
contaminant by polycyclic aromatic hydrocarbons in the Gironde estuary. Marine
Chemistry, 58: 85-97.
Countway, R.E., Dickhut, R.M., Canuel, E.A., 2003. Polycyclic aromatic hydrocarbon (PAH)
distributions and associations with organic matter in surface waters of the York River,
VA estuary. Organic Geochemistry, 34: 209-224.
Gigliotti, C.L., Totten, L.A., Offenberg, J.H., Dachs, J., Reinfelder, J.R., Nelson, E.D., Glenn
IV, T.R., Eisenreich, S.J., 2005. Atmospheric concentrations and deposition of polycyclic
aromatic hydrocarbons to the Mid-Atlantic east coast region. Environmental Science and
Technology, 39: 5550-5559.
Tsapakis, M., Apostolaki, M., Eisenreich, S., Stephanou, E.G., 2006. Atmospheric deposition
and marine sedimentation fluxes of polycyclica aromatic hydrocarbons in the eastern
Mediterranean basin. Environmental Science and Technology, 40: 4922-4927.
Baars, B-J., 2002. The wreckage of the oil tanker Erika human health risk assessment of
beach cleaning, sunbathing and swimming. Toxicology Letters, 128, 55-68.
Baum, A., Rixen, T., Samiaji, J., 2007. Relevance of peat draining rivers in central Sumatra for
the riverine input of dissolved organic carbon into the ocean. Estuarine, Coastal and Shelf
Science 73, 563 570.
Bofetta, P., Jourenkova, N., Gustavsson, P., 1997. Cancer risk from occupational and
environmental exposure to polycyclic aromatic hydrocarbons. Cancer Causes and
Control, 8: 444-472.
BPDAS Indragiri Rokan., 2004. Physical condition of Siak River water (in Indonesian
language). Paper presented in the 1st SPICE Workshop Cluster 3.1., Pekanbaru,
Indonesia.

100

Brunk, B.K., Jirka, G.H., Lion, L.W., 1997. Effects of salinity changes and the formation of
dissolved organic matter coatings on the sorption of phenanthrene: Implication for
pollutant trapping in estuaries. Environmental Science and Technology, 31, 119-125.
Budzinski, H., Jones, I., Bellocq, J., Pierard, C., Garriques, P., 1997. Evaluation of sediment
contaminant by polycyclic aromatic hydrocarbons in the Gironde estuary. Marine
Chemistry 58, 85-97.
Chen, Y., Zhu, L., Zhou, R., 2007. Characterization and distribution of polycyclic aromatic
hydrocarbon in surface water and sediment from Qiantang River, China. Journal of
Hazardous Materials, 141: 148-155.
Christl, I., Kretzschmar, R., 2001. Relating ion binding by fulvic and humic acids to chemical
composition and molecular size. 1. Proton Binding. Environmental Science and
Technology, 35, 2505-2511.
de Abrantes, R., de Assuno, J.V., Pesquero, C.R., 2004. Emission of polycyclic aromatic
hydrocarbons from light-duty diesel vehicles exhaust. Atmospheric Environment 38,
1631-1640.
de Luca, G.D., Furesi, A., Micera, G., Panzanelli, A., Piu, P.C., Pilo, M.I., Spano, N., Sanna.,
2005. Nature, distribution and origin of polycyclic aromatic hydrocarbons (PAHs) in the
sediments of Olbia harbor (Northern Sardinia, Otaly). Marine Pollution Bulletin, 50:
1223-1232.
Fernandes, M.B., Sicre, M.-A., Boireau, A., Tronczynski, J., 1997. Polyaromatic hydrocarbon
(PAH) distributions in the Seine river and its estuary. Marine Pollution Bulletin, 34(11):
857-867.
Gonzlez, J.J., Vias, L., Franco, M.A., Fumega, J., Soriano, J.A., Grueiro, G., Muniatugui, S.,
Lpez-Maha, P., Prada, D., Bayona, J.M., Alzaga, R., Albaigs, J., 2006. Spatial and
temporal distribution of dissolved/dispersed aromatic hydrocarbons in seawater in the
area affected by the Prestige oil spill. Marine Pollution Bulletin, 53: 250-259.
Guo, W., He, M., Yang, Z., Lin, C., Quan, X., Wang, H., 2007. Distribution of polycyclic
aromatic hydrocarbons in water, suspended particulate matter and sediment from Daliao
River watershed, China. Chemosphere, 68: 93-104.
Gustafsson, ., Haghseta, F., Chan, C., Macfarlane, J., Gschwend, P.M., 1997. Quantification
of the dilute sedimentary soot phase: implications for PAH speciation and bioavailability.
Environmental Science and Technology 31, 203-209.
Heemken, O.P., satchel, B., Theobald, N., Wenclawiak, B.W., 2000. Temporal variability of
organic micropollutants in suspended particulate matter of the River Elbe at Hamburg and
the River Mulde at Dessau, Germany. Archieve Environmental Contamination
Toxicology, 38: 11-31.
Hussain, M., Rae, J., Gilman, A., Kauss, P., 1998. Lifetime health rissk assessment from
exposure of recreastional users to polycyclic aromatic hydrocarbons. Archives of
Environmental Contamination and Toxicolology, 35: 527-531.
ICES, 1997. Determination of polycyclic aromatic hydrocarbons (PAHs) in sediments:
Analytical methods. In Report of the ICES Advisory Committee on the Marine
Environment, 1997. ICES Cooperative Research Report 222, 118-124.
Jenkins, B.M., Jones, A.D., Turn, S.Q., Williams, R.B., 1996. Emission factors for polycyclic
aromatic hydrocarbons from biomass burning. Environmental Science and Technology,
30, 2462-2469.
Jonker, M.T.O., Koelmans, A.A., 2002. Sorption of polycyclic aromatic hydrocarbons and
polychlorinated biphenyls to soot and soot-like materials in the aqueous environment :
mechanistic considerations. Environmental Science and Technology, 36, 3725-3734.
Kakareka, S.V., Kukharchyk, T.I., 2003. PAH emission from the open burning of agricultural
debris. The Science of the Total Environmental, 308: 257-261.
Law, R.J., Dawes, V.J., Woodhead, R.J., Matthiessen, P., 1997. Polycyclic aromatic
hydrocarbons (PAH) in seawater around England and Wales. Marine Pollution Bulletin,
34(5): 306 322.

101

Liu, H., Amy, G., 1993. Modeling partitioning and transport interactions between natural
organic matter and polynuclear aromatic hydrocarbons in groundwater. Environmental
Science and Technology, 27, 1553-1582.
Liu, W.X., Dou, H., Wei, Z.C., Chang, B., Qiu, W.X., Liu, Y., Tao, S., 2009. Emission
characteristics of polycyclic aromatic hydrocarbons from combustion of different
residential coals in North China. Science of the Total Environment 407, 1436-1446.
Luo, X-J., Chen, S-J., Mai, B-X., Yang, Q-S., Sheng G-Y., Fu, J-M., 2006. Polycyclic aromatic
hydrocarbons in suspended matter and sedimens from the pearl River Estuary and
adjacent coastal areas, China. Environmental Pollution, 139: 9 20.
McGroddy, S.E., Farrington, J.W., 1995. Sediment porewater partitioning of polycyclic
aromatic hydrocarbons in three cores from Boston Harbor, Massachussetts.
Environmental Science and Technology, 29, 1542-1550.
Neff, J.M., 1979. Polycyclic aromatic hydrocarbon in the aquatic environment: sources, fates
and biological effects. Applied Science Publisher Ltd, Essex, UK, 262 pp.
Oanh, N.T.K., Albina, D.O., Ping, L., Wang, X., 2005. Emission of particulate and polycyclic
aromatic hydrocarbons from selected cookstove fuel systems in Asia. Biomass and
Bioenergy, 28, 579-590.
Olivella, M.A., Ribalta, T.G., de Febrer, A.R., Mollet, J.M., de las Heras, F.X.C., 2006.
Distribution of polycyclic aromatic hydrocarbons in riverine waters after Mediterranean
forest fires. Science of the Total Environment, 355: 156-166.
Page, S.E., Siegert, F., Rieley, J.O., Boehm, H-D. V., Jaya, A., Limin, S., 2002. The amount of
carbon released from peat and forest fires in Indonesia during 1997. Nature 420, 61-65.
Reddy, C.M., Quinn, J.G., 2001. The North Cape oil spill. Hydrocarbons in Rhode Island
coastal waters and Point Judith Pond. Marine Environmental Research 52, 445-461.
Requejo, A.G., Sassen, R., McDonald, T., Denoux, G., Kennicutt II, M.C., Brooks, J.M., 1996.
Polynuclear aromatic hydrocarbons (PAH) as indicators of the source and maturity of
marine crude oils. Organic Geochemistry, 24(10/11): 1017-1033.
Riddle, S.G., Jakober, C.A., Robert, M.A., Cahill, T.M., Charles, M.J., Kleeman, M.J., 2007.
Large PAHs detected in fine particulate matter emitted from light-duty gasoline vehicles.
Atmospheric Envrionment 41, 8658-8668.
Rybicki, B.A., Nock, N.L., Savera, A.T., Tang, D., Rundle, A., 2006. Polycyclic aromatic
hydrocarbon-DNA adduct formation in prostate carcinogenesis. Cancer Letter, 239: 157167.
Sargenti, S.R., McNair, H.M., 1998. Comparison of solid-phase extraction and supercritical
fluid extraction for extraction of polycyclic aromatic hydrocarbons from drinking water.
Journal of Microcolumn Separation 10(1), 125-131.
Schlautman, M.A., Morgan, J.J., 1993. Effects of aqueous chemistry on the binding of
polycyclic aromatic hydrocarbons by dissolved humic materials. Environmental Science
and Technology, 27, 961-969.
See, S.W., Balasubramanian, R., Rianawati, E., Karthikeyan, S., Streets, D.G., 2007.
Characterization and source apportionment of particulate matter  2.5 m in Sumatra,
Indonesia during a recent peat fire episode. Environmental Science and Technology, 41,
3488-3494.
Shi, Z., Tao, S., Pan, B., Fan, W., He, H.C., Zuo, Q., Wu, S.P., Li, B.G., Cao, J., Liu, W.X., Xu,
F.L., Wang, X.J., Shen, W.R., Wong, P.K., 2005. Contamination of rivers in Tianjin,
China by polycyclic aromatic hydrocarbons. Environmental Pollution, 134: 97-111.
Shou, M., Ktausz, K.W., Gonzalez, F.J., Gelboin, H.V., 1996. Metabolic activation of the potent
carcinogen dibenzo(a,h)anthracene by cDNA-expressed human cytochromes P450.
Archives of biocgemistry and biophysics, 38(1): 201-207.
Siegel, H., Stottmeister, I., Reimann, J., Gerth, M., Jose, C., Samiaji, J., 2009. Siak river
system-east Sumatra characterisation of sources, estuarine processes, and discharges into
the Malacca Strait. Journal of Marine Systems 77, 148-159.
Singh, R., Sram, R.J., Binkova, B., Kalina, I., Popov, T.A., Georgieva, T., Garte, S., Taioli, E.,
Farmer, P.B., 2007. The relationship between biomarkers of oxidative DNA damage,
polycyclic aromatic hydrocarbon DNA adducts, antioxidant status and genetic

102

susceptibility following exposure to environmental air pollution in humans. Mutation


Research, 620: 83-92.
Soclo, H.H., Garrigues, PH., Ewald, M., 2000. Origin of polycyclic aromatic hydrocarbons
(PAHs) in coastal marine sediments: case studies in Cotonou (Benin) and Aquitaine
(France) areas. Marine Pollution Bulletin, 40(5): 387-396.
Tremblay, L., Kohl, S.D., Rice, J.A., Gagn, J-P., 2005. Effects of temperature, salinity, and
dissolved humic substances on the sorption of polycyclic aromatic hydrocarbons to
estuarine particles. Marine Chemistry, 96: 21-34.
Turner, A., Rawling, M.C., 2001. The influence of salting out on the sorption of neutral organic
compounds in estuaries. Water Research, 35(18), 4379-4389.
Vila-Escal, M., Vegas-Villarubia, T., Prat, N., 2007. Release of polycyclic aromatic
compounds into a Mediterranean creek (Catalonia, NE Spain) after a forest fire. Water
Research, 41: 2171-2179.
Weber Jr, W.J., McGinley, P.M., Katz, L.E., 1991. Review Paper: Sorption phenomena in
subsurface systems: concepts, models and effects on contaminant fate and transport.
Water Research 25(5), 499-528.
Williamson, K.S., Petty, J.D., Huckins, J.N., Lebo, J.A., Kaiser, E.M., 2002. HPLC-PFD
determination of priority pollutant PAHs in water, sediment, and semipermeable
membrane devices. Chemosphere, 49: 703-715.
Witt, G., 2002. Occurance and transport of polycyclic aromatic hydrocarbons in the water
bodies of the Baltic Sea. Marine Chemistry 79: 49-66.
Yunker, M.B., Macdonald, R.W., Vingarzan, R., Mitchell, H., Goyette, D., Sylvestre, S., 2002.
PAHs in the Fraser River basin: a critical appraisal of PAH ratio as indicators of PAH
source and composition. Organic Geochemistry 33, 489-515.
Zhang, Z.L., Hong, H.S., Zhou, J.L., Yu, G., 2004. Phase association of polycyclic aromatic
hydrocarbons in the Minjiang River Estuary, China. The Science of the Total
Environment, 323: 71-86.
Zhou, J., Wang, T., Huang, Y., Mao, T., Zhong, N., 2005. Size distribution of polycyclic
aromatic hydrocarbons in urban and suburban sites of Beijing, China. Chemosphere 61,
792-799.
Zhou, J.L., Maskaoui, K., 2003. Distribution of polycyclic aromatic hydrocarbons in water and
surface sediments from Daya Bay, China. Environmental Pollution, 121: 269 281.
Zhu L., Chen, Y., Zhou, R., 2008. Distribution of polycyclic aromatic hydrocarbons in water,
sediment and soil in drinking water resource of Zhejian Province, China. Journal of
Hazardous Materials 150, 308-316.
Zielinska, B., Sagebiel, J., Arnott, W.P., Rogers, C.F., Kelly, K.E., Wagner, D.A., Lighty, J.S.,
Sarofim, A.F., Palmer, G., 2004. Phase and size distribution of polycyclic aromatic
hydrocarbons in diesel and gasoline vehicle emissions. Environmental Science and
Technology, 38: 2557-2567.

103

VI.

A COMPARISON OF POLYCYCLIC AROMATIC HYDROCARBONS


(PAHs) IN PEATLAND AND NON-PEATLAND AQUATIC SYSTEM
SURFACE SEDIMENTS: A STUDY OF THE SIAK ESTUARY,
SUMATRA, INDONESIA AND OF THE WENCHANG AND
WANQUAN ESTUARIES, HAINAN ISLAND, CHINA
Muhammad Lukman*, Wolfgang Balzer*, Xiaoliang Tang*,
Christine Jose**, Jing Zhang***

*Marine Chemistry Working Group, Department of Biology/Chemistry, University of Bremen,


Leobener Strasse, 28359 Bremen, Germany
**Department of Chemistry, University of Riau, Jl. Simpang Baru Panam, 28293 Pekanbaru, Riau,
Indonesia
***Institute of Estuarine and Coastal Research, East China Normal University, 3663 Zhongshan Road
North, Shanghai 200062, China

Abstract
The distribution of polycyclic aromatic hydrocarbons (PAHs) in two grain-size fractions
(coarse: 2 mm - 63m) and fine: < 63m) of sediments from peatland aquatic systems of the
Siak Estuary and Coast, Sumatra, Indonesia, was compared with those from non-peatland
systems of Wenchang and Wanquan (WW/WQ) coastal estuaries, Hainan, China. The
sediments were sampled in September 2004 and July 2005 for the Siak Estuary and Coast and in
December 2006 and July 2007 for the WW/WQ estuary, and analyzed for the PAHs of the US
EPA priority list using HPLC with UV and fluorimetric detectors. The PAHs in the peatland
system are characterized by high contents of PAHs and organic matter in the coarse fraction.
The level of the US EPA priority pollutants (PAH) ranged from 125 ng/g d.w. to 1828 ng/g
d.w. (median m = 689 ng/g d.w.) in the coarse fraction, and from 88 ng/g d.w. to 426 ng/g d.w.
(m = 202 ng/g d.w.) in the fine fraction of the Siak sediments, while it ranged from 94.3 ng/g to
386 ng/g d.w. (m = 138 ng/g d.w.), and from 93.8 to 575 ng/g d.w. (m = 430 ng/g d.w.) in the
coarse and the fine fractions of the WW/WQ sediments, respectively. The concentration of
dissolved PAHs in WW/WQ coastal estuary was relatively low; the PAHs ranged from 7.36 to
16.2 ng/l. In contrast, the level of the PAHs dissolved in the Siak estuary and the coast ranged
from 121 to 619 ng/l. The sediment-water distribution coefficients (KD) values of the individual
PAH in the Siak estuary ranged in logarithmic value from 2.22 to 5.58 (median m = 3.37) for
the coarse fraction, and from 1.53 to 5.03 (m = 3.01) for the fine fraction. On the other hand, in
the WW/WQ estuary and coast it ranged greatly from 1.12 to 5.89 (m = 3.93) in the coarse
fraction, and from 2.58 to 5.85 (m = 4.40) in the fine fraction. It suggests that the high DOC
contents in the Siak water may sustain the PAHs in the water column impeding their sorption
onto suspended particles with subsequent sedimentation and incorporation into the sediments.

Keywords: PAHs; Sediment; Siak Sumatra, Wenchang/Wanquan Hainan, Estuary

MANUSCRIPT PREPARED TO BE SUBMITTED


(TABLES MENTIONED IN THIS MANUSCRIPT ARE PROVIDED IN APPENDIX 6: p.139-140)

104

6.1.

Introduction
Studies of land-water interactions with respect to organic pollutants such as polycyclic

aromatic hydrocarbons (PAHs) are important in providing researchers with a better


understanding of how pollutants are transferred and distributed in aquatic compartments. This is
particularly true for very dynamic, transitional environments such as rivers, estuaries and coasts.
One of the most effective ways of understanding such systems is to evaluate sedimentary
pollution as the capacity of sediments to concentrate and retain heavy metals and organic
compounds results from complex geochemical factors such as grain size and organic matter
content (Horowitz & Elrick, 1987; Luthy et al. 1997).
PAHs are one class of organic contaminants that are relevant worldwide owing to their
toxic/carcinogenic properties, their widespread and common anthropogenic sources such as
biomass burning and fossil fuel combustion (Neff, 1979; IARC, 1987). This is particularly
important for Asia, which annually contributes more than 50% of the global PAH emissions
(520 Gigagrams in 2004), with PAHs stemming mainly from biomass burnings (Zhang & Tao,
2009). Since PAHs constitute up to 13% (by mass) of oils (Requejo et al., 1996), we can expect
that oil discharges into water bodies add significant amounts of PAHs into these systems (e.g.
Saeed & Al-Mutairi, 2000; Wang et al., 2004). Signs of petrogenic PAH contamination due to
petroleum product spills have already been recognized for the Malacca Strait, where the Siak
estuary ends (Zakaria et al., 2000).
However, little attention has been paid to PAHs in the context of estuaries draining
peatland which enhances its degradation such as that found in the Siak River estuary. Most
land-water interaction studies of these particular peatland-aquatic systems have focused on the
increasing levels of dissolved organic matter exported into the sea due to land-use changes (e.g.
Baum et al., 2007; Siegel et al., 2009). Simultaneously, however, Sumatra's peatlands are
considered to be potential PAH sources due to widespread fires (e.g. Fang et al., 1999; See et
al., 2007). The degradation of Sumatran peatland (e.g. Page et al., 2002) has been of great
concern, not only for climate changes associated with the emission of up to 2.57 Gt of carbon
recorded during fires in the El Nino event of 1997/1998 (Page et al., 2002), but also due to its
importance for the distribution of pollutant. See et al., (2007) revealed that Sumatra's peatland
burning events observed in the haze around Dumai and Pekanbaru, created ca. 100 to 600 ng m-3
of 16 PAHs from the list of US EPA priority compounds. This was even higher than levels
observed (ca. 116 ng m-3) for the polluted urban areas around Beijing, China (Zhou et al., 2005).
Baum et al (2007) estimated that peat-draining along the Siak River may cause levels of
dissolved organic carbon (DOC) export reaching as high as 0.3 Tg C yr-1. High DOC flux into
aquatic systems can facilitate the water-borne transport of PAHs (e.g. Liu & Amy, 1993;
McGroddy & Farrington, 1995) and can hypothetically determine their fate in these aquatic

105

systems, including their distribution in sediments. The characterization of PAHs in sediment


taken from humic-rich aquatic systems containing degraded peatland is also relevant for a better
environmental management.
This study aims to compare the PAH distributions in surface sediments for two
Wentworth size fractions: coarse (2mm 63m) and fine (<63m). Samples were taken from
both a typical peatland- and a non-peatland aquatic system. This case study links two
international research projects analyzing land-borne material transfer. The first is the Science
for the Protection of Indonesian Coastal Marine Ecosystem (SPICE) program, a cooperative
German-Indonesian research effort. The second is the Land-Sea Interactions along Coastal
Ecosystems of Tropical China: Hainan (LANCET) program, a German-Chinese cooperative
research project. The sediments collected were examined for their total PAH content, their PAH
compositions, and their sources. The two sampling areas were the Siak River estuary in
Sumatra, Indonesia and the Wenchang/Wanquan Estuaries in Hainan, China. The main
difference between the two aquatic systems is comparably very high level of dissolved (humic)
organic carbon (DOC) in the Siak estuary. Thus, one aim of the present study is to investigate
whether a high dissolved humic content stabilize the PAHs in the water column and prevent it
from transfer to the sediment.

6.2.

Materials and Methods

6.2.1.

Study areas, Sample Collection and Fractionation


The Siak estuary is one of the most pertinent peat-draining rivers in Sumatra. Comprising

the Siak River and its tributaries, the Siak basin stretches over 300 km and drains a large area of
lowland and peat swamps with various landscapes, including huge palm-oil plantations,
rainforests and oil refinery near the Siak mouth (Fig. 1a). In terms of PAH relevance, the areas
studied have decades-long records showing frequent dense smog events caused both by the
common practice of agricultural/biomass burning and by naturally occurring forest and swamp
fires. In addition, oil discharges from routine river-boat transports as well as from oil industry
related activities (production, transportation and disposals of residues) are readily observable.
The Wenchang and Wanquan coastal estuaries are typical, non-peatland water bodies on Hainan
Island, China (Fig. 1b). These two aquatic systems are both surrounded by large agricultural
landscapes and human settlements. Although having relatively high PAH content in the
sediment, the two Hainan estuaries are not characterized by intense biomass burning events.
As part of the SPICE program, sediment samples were taken during 2004-2005 from a
total of eight stations along the Siak estuary and the coast (Fig. 1a). A full report on the
distribution and sources of PAHs in the Siak River, its estuary and the coastline was presented
in a previous study (Lukman et al., manuscript being prepared for journal publication). Samples
from the Wenchang and Wanquan coastal estuaries were collected during the December 2006

106

and July 2007 LANCET sampling campaigns. The samples were collected directly from a boat
using a sediment grab sampler. They were immediately homogenized before being placed in
closed aluminum jars and brown glass containers. During homogenization, foreign objects such
as large sticks, stones or any other synthetic waste were removed. The samples were kept cool
(ca. 4oC) during transport to the laboratory, where they were frozen (-20oC) and stored until
further treatment. Fractionation was performed by wet sieving the material to separate the
coarse fraction (2 mm - 63 m). The fine fraction (< 63m) was collected by centrifuging the
remaining sample.

Sumatra

S253
S230

INDONESIA

S 250
S134
S138

S231
S232

S143

107

CHINA

Wenchang

Hainan
Wanquan

BOAO CITY
CC02-1/07

BB4/07

K/06
WRB/06

Wanquan
Coastal Estuary

Aquaculture

WW8/07
WENCHANG
CITY

H/06
WW10/07

Wenchang
Coastal Estuary

Fig. 1. The maps of the sampling stations: (A) the Siak estuary (crossed cycles), Sumatra, Indonesia; (B)
the Wenchang and Wanquan Estuaries (black cycles), Hainan, China.

6.2.2.

Determination of PAHs

6.2.2.1. Extraction and Working-Up


Up to 10g of the sediment fractions were spiked with three surrogate perdeuterated PAH
standards: d10-phenanthrene (PHEN D10), d10-fluoranthene (FLA D10) and d12-perylene
(PERY D12), then extracted in a Soxhlet-modified extractors (SoxTec) with solvents suggested
by the ICES Method (ICES, 1997) for 6 8 hours. The first cycle was extracted using acetone,
followed by a mixture of acetone/hexane (1/9 ; v/v) in the second cycle. The extracts were then
combined and reduced to ca. 1 mL by a rotary evaporator. A 1:2 (w/w) deactivated
alumina/silica (10% / 3%) column, including dried anhydrous Na2SO4 to remove the coextracted water, was used to clean up the extract. During the cleanup process the extracts were
eluted with 40 mL of degassed 3/7 (v/v) dichloromethane/hexane. Excess solvent in the clean
extracts was removed by the evaporator, resulting in sample volumes of ca. 1 mL. Then, the
extracts were subjected to solvent exchange into acetonitrile (ACN) to allow high performance
liquid chromatography (HPLC) analysis.
Dissolved PAHs were obtained by filtering the water samples with GF/F, and the filtrate
was extracted using a SPE cartridge (Chromabond NH2/C18) obtained from Macherey-Nagel
GmbH (Dueren, Germany) at flow rate of ca. 5 mL/min controlled by adjusting the pressure.
Methanol (2% of the filtrates volume) was added to 0.5 to 1 L of the filtrate in order to enhance

108

the recovery. Deuterated surrogate standards were spiked into the sample to compensate the
procedural efficiency and reproducibility from any losses during extraction and work-up
procedure. The SPE cartridges were sequentially conditioned with 5 mL hexane and 10 mL
methanol, then washed with ca. 100 mL Milli-Q water before using. After all of the sample was
aspirated, the cartridge was again washed with Milli-Q water. Then, the column was firmly
wrapped with foil, sealed with a paraffin film, and stored at -4 oC until analysis. Prior to eluting
the retained PAHs, the cartridges were gently dried by a stream of N2. Then, the PAHs were
eluted with 4 mL of dichloromethane (DCM) (repeated three times) at flow rate of 5 mL/min.
The use of DCM for elution of PAH for SPE system has been suggested (e.g. Sargenti &
McNair, 1998). The elutes were reduced to a dry state by gently dried by N2 blow down, then
matrix was changed into ACN for HPLC analysis.
6.2.2.2. HPLC Analysis
The samples were analyzed for 15 out of the 16 PAHs priority pollutants listed by the US
EPA, including naphthalene (NAPH), acenaphthene (ACEN), fluorene (FLU), phenanthrene
(PHEN), anthracene (ANTH), fluoranthene (FLA), pyrene (PYR), benzo(a)anthracene (BaA),
chrysene

(CHRY),

benzo(b)fluoranthene

(BbFLA),

benzo(k)fluoranthene

(BkFLA),

benzo(a)pyrene (BaP), dibenzo(a,h)anthracene (DANTH), benzo(g,h,i)perylene (BPERY) and


indeno(1,2,3-c,d)pyrene (IPYR); the EPA priority PAH acenaphthylene was not analyzed.
Baseline separation and quantification of those compounds was performed using a reverse phase
RP-C18 column (250 x 4.6 mm, 5m) and high performance liquid chromatograph (HPLC LKB
2249 Broma) with ultraviolet and programmable fluorescence detectors. Elution was 55%
acetonitrile (ACN) : 45% water for 10 min at the beginning of the elution program (isocratic),
55% to 100% ACN (a linear gradient at 2.25% ACN/min.) for 20 min, and 100% ACN for 10
min at the last program before the composition of the elution was set back to 55% ACN for 5
min. The elution was programmed at a constant flow rate of 1 mL min-1 for in total of 45
minutes.
HPLC grade acetonitrile and analytical-grade solvents (acetone, dichloromethane and
hexane) were used throughout the analytical procedures. A certified mixture of 16 PAHs
standards (purchased from Dr. Ehrenstorfer GmbH, Germany) were used to identify and
quantify the analytes. The limit of detection ranged from 0.4 ng/mL (BaP) to 12 ng/mL (NAPH)
for fluorescence detection. The three perdeuterated surrogate PAHs were used to achieve
procedural efficiency, reproducibility and data correction from any losses during the extraction
and work-up procedures. The recovery ranged respectively from 64.5% to 114% (mean =
87.8%), 67.3% to 119% (m = 93.2%), and 61.6 to 121% (m = 92.8%) for PHEN D10, FLA D10
and PERY D12 with their relative standard deviations (RSD) of 16.1%, 13.5% and 14.3%. The

109

mean recovery standard deviation of the deuterated surrogates for the SPE was 100% 16.1%
(PHEN D10), 90.6% 4.55% (FLA D10) and 93.4% 10.7% (PERY D12).
6.2.3.

Determination of Sedimentary Organic Matter


Sedimentary organic carbon and nitrogen contents and were analyzed at the Leibniz

Center for Tropical Marine Ecology (ZMT), Bremen, Germany. The content of the elements
was measured using a C/N elemental analyzer (Carlo Erba NA 2100, Milan, Italy) operating at
1100oC (during combustion under supplied oxygen). 30 mg of each homogenized sediment
fraction was placed in a silver cup, treated with 200 L of 1 N HCl to remove inorganic carbon.
This process was repeated twice (when necessary) to ensure all carbonates were transformed
into carbon dioxide, then subsequently dried at 40oC overnight. The method reproducibility was
measured by use of a sediment standard (LECO 1012, Leco, St. Joseph, MI) with 1.30 0.04
wt.-% C (mean standard deviation) and 0.13 0.04 wt.-% N after every fifth sample. The
reproducibility (by relative standard deviation) was 2.71% for C and 4.0% for N for the LECO
standards (n=5), and was 0.43% for C and 6.52% for N applied for the fine fraction sample of
WR-B (n=3).

6.3.

Results

6.3.1.

PAHs and Organic Matter in Sedimentary Size Fractions


The results are summarized in Table 1 (p. 139) and Table 2. (p. 140). The total PAH

content (PAH15) in the sediment samples from the Siak estuary varied from 125 ng/g d.w. to
1828 ng/g d.w. (median, m = 689 ng/g d.w.) in the coarse fraction, and from 88 ng/g d.w. to
426 ng/g d.w. (m = 202 ng/g d.w.) in the fine fraction. For the Wenchang/Wanquan estuaries,
the recorded values varied from 94.3 ng/g to 386 ng/g d.w. (m = 138 ng/g d.w.), and from 93.8
to 575 ng/g d.w. (m = 430 ng/g d.w.) in the coarse and in the fine fraction, respectively. The
OC content of the coarse sediment ranged from 0.05% to 9.07% (m = 4.41%) and from 1.29%
to 3.31% (m = 1.66%) in the fine fraction of the Siak estuary. For the WW/WQ estuary
samples, the values ranged from 0.13% to 3.87% (m = 0.46%) in the coarse fraction and from
1.21% to 4.11% (m = 1.98%) for the fine fraction. The content of the individual PAH in the
Siak sediments ranged from below instrumental detection limits to 1061 ng/g d.w. for DANTH
in the coarse fraction and to 206 ng/g d.w. for DANTH in the fine fraction. Likewise, the
individual PAH in the WW/WQ sediment varied from below detection limits to 97 ng/g d.w. for
DANTH in the coarse fraction and to 178 ng/g d.w. for pyrene in the fine fractions. The
distribution of individual PAHs between the fractions was generally very distinct when
comparing the Siak sediments to the WW/WQ sediments. In the sediments taken from the Siak,
PAHs were enriched in the coarse fractions in most cases, whereas the WW/WQ sediments
mainly showed PAH enrichment mainly in the fine fraction. However, individual PAHs which

110

did not exactly follow the pattern emerged, showing variation between the stations. BPERY
was the only compound which was mainly enriched in the fine fraction from the Siak samples.
In the WW/WQ sediments, only the sample taken from Wanquan station B/06 showed a
similar distribution pattern as those for most of the Siak samples.
With respect to total PAH content, Fig. 2 compares the distributions for the Siak
sediments and the WW/WQ sediments. In general, there was a large difference in PAH
distributions between the two distinct environmental settings. The total content of PAHs tended
to be higher in the coarse fraction (m =

689 ng/g d.w.) for Siak sediments than it was in

WW/WQ sediments. On the other hand, the fine fraction of the Siak sediments contained less
PAHs (m = 202 ng/g d.w.) than that of similar WW/WQ sediments (m = 430 ng/g d.w.). This
pattern of PAH distribution correlates well with the pattern of OC contents. As a proxy, the C/N
ratios suggest dissimilar sources for the organic matter isolated from the coarse and fine
fractions. The C/N ratios for the coarse sediments generally tended to be higher than those of
the fine fractions for both the Siak and the WW/WQ sediments. They ranged from 12.4 to 47.9
with a mean ratio of 23.3 and from 10.1 to 88.2 (mean = 38.8), respectively. On the other hand,
the fine fraction C/N ratios ranged from 8.85 to 20.4 (mean = 16.2) for the Siak and from 11.6
to 52.3 (mean = 19.0) for the WW/WQ sediments. As a point of reference, C/N end-member
values of ca. 7 and higher than 20 are commonly used to define marine and terrigenous origins,
respectively (Meyer, 1994; Holtvoeth et al., 2005). Values between 6-12 suggest a mixture
dominated by marine-originated organic matter (Ruttenberg & Go i, 1997).
2000

Siak - SUMATRA

PAH15 (ng/g d.w.)

1800
1600

Coarse

1400

Fine

1200
1000
800
600
400
200
0

S 143 S 138 S 134 S 250 S 253 S 230 S 231 S 232

PAH15 (ng/g d.w.)

700

Wenchang

600

HAINAN

Wanquan

500
400
300
200
100
0

Fig. 2. Distribution patterns of PAHs (as PAH15) in the coarse and fine fractions of the Siak sediments
(A) and the Wenchang and Wanquan sediments (B).

111

6.3.2.

Relative Composition of PAHs


The relative composition of PAH compounds was largely identical between the fractions

in both studied areas, as can be seen by both the individual and the ring-group compositions
(Fig. 4 and Fig. 5). However, there was a significant difference in the proportions between the
two areas. In the Siak estuary and coast, the relative composition of the PAH compounds was
shared almost evenly by the compounds, except for NAPH and DANTH (m = > 15%). Even,
the relative composition of DANTH reached up to 30% in the coarse fraction. For the WW
estuary and coast, the PAHs were dominated by PHEN, FLA, PYR, and BbFLA with median
values of the relative composition of those compounds > 12%, except for FLA (m = 5.93%) in
the coarse fraction.
The ring-group compositions (Fig. 5) classified as 2 ring (NAPH), 3 rings (ACEN, FLU,
PHEN, ANTH), 4 rings (FLA, PYR, BaA, CHRY), 5 rings (BbFLA, BkFLA, BaP, DANTH),
and 6 rings (BPERY, IPYR) also showed a distinct pattern. The significant difference between
the studied areas was that the presence of 5-ring PAHs was high for the Siak sediments with a
median relative composition > 30%, whereas the WW/WQ sediments were generally dominated
by 3- and 4-ring structures (m = > 30%).

Relative Composition (%)

40
35

Siak - Sumatra
(median values)

30
25

Fine

20

Coarse

15
10
5
0

Relative Composition (%)

40
35

Wenchang and Wanquan Hainan


(median values)

30
25
20
15
10
5
0

Fig. 4. Comparisons of the relative composition of the individual PAH to the to the PAH15 between the
coarse and the fine fractions for the Sumatra and Hainan estuary and coasts.

112

Relative Composition (%)

60

Siak - Sumatra
(median values)

50
40

Fine

30

Coarse

20
10
0

2 rings

Relative Composition (%)

60

3 rings

4 rings

5 rings

6 rings

Wenchang and Wanquan Hainan


(median values)

50
40
30
20
10
0

2 rings

3 rings

4 rings

5 rings

6 rings

Fig. 5. Comparison of the contribution of the PAH ring group to the PAH15 between the coarse and the
fine fractions of (A) Siak sediments and (B) Wenchang and Wanquan sediments.

6.3.3.

Source Apportionment
Ratio of isomers for PAHs with the molecular weights 178 (ANTH, PHEN), 202 (FLA,

PYR) and 228 (BaA, CHRY) have been widely employed to apportion sources to either
combustion or petroleum origins (Neff 1979; Budzinski et al., 1997; Yunker et al., 2002; De
Luca et al., 2005; Soclo et al., 2000). These isomers have molecular stabilities with increasing
temperature during pyrolysis. For instance, phenanthrene is thermodynamically more stable
than the kinetically-stable isomer anthracene (Budzinski et al., 1997). The proportion of
anthracene increases as burning processes involve higher temperatures. In this study, the
molecular ratios of ANTH/(ANTH + PHEN), FLA/(FLA + PYR), BaA/(BaA+CHRY) were
applied. PAHs from combustion sources typically have the following values: ANTH / (ANTH
+ PHEN) > 0.1; FLA/(FLA-PYR) > 0.5, and BaA/(BaA+CHRY) > 0.35 (Yunker et al., 2002)..
In contrast to this, PAHs associated with petroleum, e.g. crude oil, have typical values for the
same isomeric ratios of < 0.1, < 0.4, and < 0.2 respectively (Yunker et al., 2002). For instance,
Alaskan Crude Oil has values of 0.03, 0.26 and 0.10 for these three ratios (Requeojo et al.,
1996). However, the boundaries between the assigned values are not set in stone. Any value
falling between those determining values is usually considered to evidence a mixture of

113

petroleum and combustion sources. Those values are then cross-plotted to discover the tendency
of the data (Fig. 6). Additionally, isomeric ratio data from a range of literature were reviewed
and also plotted in the Fig. 6 to get an understanding on to what extent these determining values
can properly distinguish pyrogenic and petrogenic sources of PAHs.
Our results show that the ratio of ANTH/(ANTH+PHEN) in the Siak sediments was 0.21
0.18 (mean standard deviation, n = 8) and 0.14 0.07 in the coarse and fine fractions,
respectively. Likewise, the coarse and fine ratios of FLA/(FLA-PYR) were 0.59 0.13 and 0.59
0.22, and that of BaA/(BaA + CHRY) was calculated as 0.48 0.15 and 0.36 0.08. These
ratios suggest that the PAH contamination found in Siak River sediments generally stems from
pyrogenic sources. On the other hand, the WW/WQ coarse sediment samples yielded ratios of
ANTH/(ANTH+PHEN), FLA/(FLA-PYR) and BaA/(BaA + CHRY) with values of 0.01 0.01,
0.47 0.24, and 0.39 0.24, respectively. The corresponding ratios in the fine sediments were
0.02 0.01, 0.55 0.21, 0.54 0.26. These values strongly suggest that the PAHs in the
WW/WQ sediments are generated by a mixture of both petrogenic and pyrogenic sources.
6.3.4.

Sediment-Water Distribution Coefficient


Sediment-water distribution coefficient (KD, ml/g) was evaluated to get insight into

underlying processes that control the transport and fates of the PAHs in those different aquatic
systems, in particular the role of the dissolved humic substances which may stabilize the PAHs
in the water column and inhibit it from association with the sediment. The KD value is defined
here as the ratio of PAH concentration in sediment to that in water solution, and calculated for
individual compound for both sediment fractions at three stations (S143, S138, S134) from the
Siak estuary, and four stations (BB4, CC02, WW8, WW10) from the WW/WQ estuary and
coast. The concentration of dissolved PAHs in WW/WQ coastal estuary was relatively low. The
individual PAH extended from undetected to 4.15 ng/l. The PAHs ranged from 7.36 to 16.2
ng/l. This range of concentration was considered low in comparison to aquatic systems around
industrial and intense urban centres in China such as rivers in Tianjin China (45.8 1272 ng/l)
(Shi et al., 2005) and Qiantang River China (70.3 1844 ng/l) (Chen et al., 2007). On the other
hand, the level of dissolved PAHs in the Siak estuary and the coast was reported in our previous
work (Lukman et al., manuscript being prepared for journal publication), which the PAHs
ranged from 121 to 619 ng/l.
The KD values are then plotted in the Fig. 7, which show that in general the KD values of
individual PAH in the Siak estuary are lower than those of the WW/WQ estuary and coast. In
the Siak estuary, it ranged in logarithmic value from 2.22 to 5.58 with a median value of 3.37
for the coarse fraction, and from 1.53 to 5.03 (median m = 3.01) for the fine fraction. On the
other hand, in the WW/WQ estuary and coast it ranged greatly from 1.12 to 5.89 (m = 3.93) in
the coarse fraction, and from 2.58 to 5.85 (m = 4.40) in the fine fraction.

114

6.4.

Discussion
It is obvious from the distribution of PAHs in the size-fractions that the Siak and the

WW/WQ sediments are fundamentally different (Fig. 2). Quantitatively, Siak sediments were
characterized by higher levels of PAHs in the coarse fractions as compared to the finer ones.
The situation for the WW/WQ sediments was the exact opposite, with higher PAH contents
found in the fine fractions. Qualitatively, the relative compositions of individual and ring-group
PAHs was also distinct between the two aquatic systems, however, they were largely similar
between the fractions located within a given system (Fig. 4 and Fig. 5). These deviations
suggest that the PAH sources located in these two sampling areas were different. DANTH and
NAPH significantly contributed to the relative composition of the total PAHs analyzed in the
Siak sediments. During Sumatran peatland burnings, the resulting smoke contained DANTH
levels commensurate with the values observed for Pekanbaru, the capital city of Riau province.
NAPH levels for this area, however, remained low (See et al., 2007). Therefore, high DANTH
in the sediments may be a strong indicator stemming from burnings. However, further
investigation is needed to determine whether DANTH is a main product of Sumatra peatland
burnings. Additionally, high NAPH levels in the Siak sediments were assumed to stem from
other sources than biomass burnings, specifically crude oil. High NAPH contents in the crude
oil obtained from this region have already been confirmed. For example, Requejo et al. (1996)
pointed out that NAPH contributes more than half of the total non-alkylated polyaromatic
compounds in crude oil. Likewise, Wang et al. (1999) observed that naphthalene and its alkylhomologues (C1-C4) comprised up to 86% of the total PAHs in Diesel No.2 and up to 99% in
Jet B fuel. If these findings are general, then Siak sediments have been affected by crude oil
pollution. On the other hand, PAHs in the WW/WQ sediments are characterized by high
contents of PHEN, FLA, PYR, and BbFLA. The isomeric ratio for ANTH/(ANTH+PHEN) was
relatively low with a mean value of 0.01. This value was largely similar to that for bituminous
coal emissions (e.g. Yang et

al., 1998). The ratios for FLA/(FLA+PYR) and

BaA/(BaA+CHRY) are in agreement with those calculated for coal and petroleum combustion
(e.g. Marr et al., 1999; Li et al., 1999).
Sedimentary organic matter is a pivotal factor affecting PAH distribution in the coarser,
grain-size fractions. In this study, higher PAH contents generally correspond to higher OC
contents for both aquatic systems. The distribution of the OC content between the grain-size
fractions showed similar patterns to their corresponding PAH contents. The higher OC contents
(median = 4% d.w. ) and the C/N ratios (mean value = 23) found in the coarse fraction of Siak
sediments confirmed that highly condensed organic matter such as that stemming from burning
residues (often termed black carbon, BC), plant debris, and peat existed. C/N values of > 20 are
commonly associated with terrestrial plants, thermal degradation of biomasses, or peat

115

(Holtvoeth, 2004, Pillon et al., 1985). Therefore, we can conclude that the high PAH levels in
the coarse fractions of Siak sediments are most probably associated with the presence of
burning events, their residues, peat, and fragmented plant materials. These materials show a
strong affinity for PAHs (e.g. Rockne et al., 2002; Ghosh et al., 2003; Cornelissen &
Gustafsson, 2004; Gustafsson et al, 1997; Grathwohl, 1990). Black carbon is also co-produced
with PAHs during combustion processes (Jonker & Koelmans, 2002). The presence of such
carbonaceous materials in Siak River sediments strongly indicates that the Siak, its estuary and
the coastline are affected by peatland fires and biomass burnings. Such large PAH contents can
conceivably be produced during peatland burning, including occlusion of PAHs in matrixes of
the burning residues. These substances are then free to enter the aquatic system via both land
drainage and atmospheric deposition. High erosion rates have been recognized in much of the
Siak estuary and the along the coast as evidenced by increases in the suspended particulate
matter load (see Siegel et al., 2009). In comparison, the coarse fraction of the WW/WQ
sediments were typically characterized by organic-poor matter, especially silica particles as
commonly found along the Hainan coast (e.g. Ghosh et al., 2000). Quantitatively, the organic
matter contents in the coarse fraction are generally small for the Chinese samples, but had high
C/N ratios (mean value = 39). This suggests that the organic matter in these sediments was
limited to that of carbonaceous materials already existing in the fraction.
On the other hand, the PAH content in the WW/WQ fine fractions was higher than in the
coarse fraction of the sediments. As we might expect from the large values of surface area per
gram of sediment, the particles of the fine fractions provided an excellent environment for
sorption from solution. The higher PAH levels in the fine fraction of these sediments can also be
attributed to the larger OC content as compared to that of the corresponding coarse fraction.
This suggests that the PAHs of the sediment particles were controlled by an equilibrium
sorption process of PAHs onto the sedimentary organic matter. The OC contents calculated for
the fine fractions of both the Chinese and Sumatran aquatic systems were largely the same, with
median values of ca. 2% d.w.. The relatively low C/N ratio (median = 13) determined for the
WW/WQ fine sediments suggests a mixture of both marine and terrestrial organic matter. On
the other hand, the relatively high C/N ratio (m = 18) representative of the Siak River's fine
fraction can most likely be attributed to the presence of humic substances.
With regard to PAH-organic matter sorption rates, the lower PAH levels associated with
the fine fractions in Siak sediments (as compared to those for the coarse fraction) can be
hypothetically driven by two causal mechanisms. First, the presence of combustion residues in
the coarse fraction might increase sorptive competition for PAHs, meaning that carbonaceous
materials rather than bulk organic matter would be favoured (see Luthy et al., 2002; Cornelissen
& Gustafsson, 2004; Accardi-Dey & Gschwend, 2002). Gustafsson et al. (1997) observed
elevated sediment-porewater PAH partitioning rates, which occurred whenever soot-partitioning

116

was included in calculating the hydrophobic partition model. Cornelissen & Gustafsson (2004)
concluded that black carbon (BC) emerges as the most important overall geosorbent constitute
in the case of low aqueous PAH (concentrations below ng/l). That would be in agreement with
our studies since the PAH concentration existing in solution was in the range of 121 to 619
ng/lin the Siak estuary, and 7.36 to 16.2 ng/lin the WW/WQ estuary and coast. . In addition,
Accardi-Dey & Gschwend (2002) infer that absorption by natural organic matter and adsorption
onto BC act in parallel to bind PAHs in Boston Harbor sediments. This means that the presence
of BC in the coarse sediments may outcompete the fine grain sediments for the sorption of
PAHs. Second, PAH sequestration in sediments can be affected by humic-enriched or highlyorganic porewater and overlying water. Siak estuary and the coast contained significant amount
of DOC compared to the Wenchang and Wanquan coastal estuaries (Balzer, unpublished data).
The calculation of the sediment-water distribution coefficient showed that the KD values in the
Siak estuary are the generally lower than those of the non-peatland water of Hainan suggesting
that high DOC in the Siak water may sustain the PAHs in the water column impeding their
association onto the sediment organic matter. Such role of colloid or DOC in sustaining the
PAHs in the water has been recognized. For instance, Chin & Gschwend (1992) observed
increased PAH sorption levels onto porewater colloids in heavily contaminated Boston Harbor
sediments. Yu et al. (2009) found that PAHs tend to be released from sediments into porewater
as the dissolved organic matter content of the porewater increases. Therefore, we can infer that
increasing the amount of DOC colloids in the porewater and the water column will affect the
sorption/desorption of PAHs by the fine fraction. However, these insights also need further
investigation to show how accurate these suggestions really are.

117

Petroleum
Combustion

Petroleum

Grass/Wood/Coal
Combustion

1.0

BaA / (BaA + CHRY)

0.9

Peat Burning
Dumai

0.8

Agricultural
debris

0.7
SPM after forest
fires

0.6
0.5

Weat grasses
Rice grasses

Oil spills
"Tanker Enrika"

0.4
0.3

light-duty diesel
Combustion
vehicles

Mixed
Sources

Bituminous coal

Charcoal

0.2
0.1

Petroleum

Alaskan crude oil

0.0

ANTH / (ANTH + PHEN)

0.4

Bituminous coal

0.4
0.3
0.3

Charcoal
Oil spills "Tanker
Enrika"
Peat Burning
Dumai

0.2
0.2
0.1
0.1
0.0

Combustion

Agricultural
debris

WW8
WW10

0.0

0.1

Weat grasses
Rice grasses
light-duty diesel
vehicles

Alaskan crude
oil

0.2

0.3

Petroleum

0.4

0.5

0.6

0.7

0.8

0.9

1.0

FLA / (FLA + PYR)


Fig. 6. Cross plots of PAH isomeric ratios for the source apportionment of PAHs in the Siak coarse (solid
squares) and fine sediment fractions (unfilled squares), and in the Wenchang Wanquan coarse (solid
triangles) and fine (unfilled triangles) sediment fractions. The discrimination values dashed lines were
adopted from Yunker et al. (2002). For comparison, various literature data of PAHs emissions and crude
oils (light grey cycles) are additionally presented to illustrate how the determining values properly
address data variation. References: petrogenic sources: Alaskan Crude Oil (Requejo et al. 1996); Oil-spill
Tanker Enrika (Baars et al., 2002); pyrogenic sources: Bituminous coal (Liu et al 2009); Charcoal (Oanh
et al., 1999); Agricultural debris (Kakareka & Kukharchyk, 2003); Rice & Weat grasses (Jenkins et
al.,1996); light-duty diesel vehicle (de Abrantes et al 2004); Sumatra peatland burning-Dumai (See et al.,
2007); SPM after forest fire (grey cycle with black outline: Vila-Escale et al 2007).

118

Siak
SUMATRA

6.00

Wenchang and Wanquan


HAINAN
NAPH

5.50

ACEN
FLU

5.00

PHEN
ANTH
FLA

4.50

PYR

log KD

BaA

4.00

CHRY
BbFLA
BkFLA

3.50

BaP
DANTH

3.00

BPERY
IPYR
SIAK, MEDIAN

2.50

SIAK, MEDIAN
HAINAN, MEDIAN

2.00

1.50

HAINAN, MEDIAN

Coarse

Fine

Coarse

Fine

Fig. 7. Sediment-water distribution coefficient (log KD) of PAHs presented as a mean value of each
individual compound from the Siak and the Wenchang and Wanquan estuaries. The median values (black
and unfilled marks) were to represent the general differences between the two systems.

6.5.

Conclusion
The Siak aquatic system, which is composed largely of degraded peatland, was found to

have enriched PAH levels associated with the coarse sediment fraction. This was particularly
true in the presence of carbonaceous materials such as biomass burning residues, peat and
fragmented plant debris. These materials have been shown to act as favorable adsorption
matrices for PAHs, especially in coarse sediment fractions. It is suggested that unfavorable
sorption of PAHs onto fine sediments associated with the organic matter in the Siak River
systems could be due to the existence of high levels of DOC found in the water column, which
can increase the partitioning effect on PAHs. However, in non-peatland areas of Hainan,
enriched PAH contents are mainly found in the fine fraction as expected. The relative
composition of PAHs between the grain-size fractions is similar either in the Sumatran or
Hainan areas suggesting that PAH contamination stemmed from similar sources. Furthermore,
the isomeric ratios used for source apportionment indicated that the PAHs found in the Siak
basin had mostly been generated through biomass burnings, whereas PAHs analyzed in the
WW/WQ sediments from Hainan Island stemmed from a mixture of coal and petroleum
combustion. For the total content of PAHs not only the sources of PAHs are important
(pyrogenic or petrogenic), but also other factors such as sediment fraction size, PAH ring size,

119

aromaticity, type of sorbent material and the levels of OM, DOC, BC in both the sediments and
the porewater. Therefore, we can conclude that the high PAH levels in the coarse fractions of
Siak sediments are most probably associated with the presence of burning events, their residues,
peat, and fragmented plant materials.

Acknowledgments
This study is part of the German-Indonesian SPICE Project Cluster 3.1., and of GermanChinese LANCET research project, funded by the Federal German Ministry of Education,
Science, Research and Technology (BMBF, Bonn), and also supported by the German
Academic Exchange Service (DAAD). We acknowledge all their supports. Furthermore, we
would like to extent our grateful to scientists and students from University of Riau, Indonesia,
the Institute of Estuarine and Coastal Research, East China Normal University, China, and to
SPICE German colleagues for their contribution during sampling campaigns and discussion.
Furthermore, we appreciate our colleagues at Leibniz Centre for Tropical Marine Ecology ZMT

Dipl. Ing. Matthias Birkicht, Ms. Dorothee Dasbach, for the organic measurement and
Ms. Sonia Tambou (Marine Chemistry Working Group, Uni-Bremen) for laboratory assistance.
Finally, we thank all reviewers for their constructive critics and comments.

References
Accardi-Dey, A., Gschwend, P.M., 2002. Assessing the combined roles of natural organic
matter and black carbon as sorbents in sediments. Environmental Science and
Technology, 36, 21-29.
Baars, B-J., 2002. The wreckage of the oil tanker Erika human health risk assessment of
beach cleaning, sunbathing and swimming. Toxicology Letters, 128, 55-68.
Baum, A., Rixen, T., Samiaji, J., 2007. Relevance of peat draining rivers in central Sumatra for
the riverine input of dissolved organic carbon into the ocean. Estuarine, Coastal and Shelf
Science 73, 563 570.
Budzinski, H., Jones, I., Bellocq, J., Pierard, C., Garriques, P., 1997. Evaluation of sediment
contaminant by polycyclic aromatic hydrocarbons in the Gironde estuary. Marine
Chemistry 58, 85-97.
Chen, Y., Zhu, L., Zhou, R., 2007. Characterization and distribution of polycyclic aromatic
hydrocarbon in surface water and sediment from Qiantang River, China. Journal of
Hazardous Materials, 141: 148-155.
Chin, Y-P., Gschwend, P.M., 1992. Partitioning of polycyclic aromatic hydrocarbons to marine
porewater organic colloids. Environmental Science and Technology, 26, 1621-1626.
Cornelissen, G., Gustafsson, ., 2004. Sorption of phenanthrene to environmental black carbon
in sediment with and without organic matter and native sorbates. Environmental Science
and Technology, 38, 148-155.
de Abrantes, R., de Assuno, J.V., Pesquero, C.R., 2004. Emission of polycyclic aromatic
hydrocarbons from light-duty diesel vehicles exhaust. Atmospheric Environment 38,
1631-1640.
De Luca, G.D., Furesi, A., Micera, G., Panzanelli, A., Piu, P.C., Pilo, M.I., Spano, N., Sanna.,
2005. Nature, distribution and origin of polycyclic aromatic hydrocarbons (PAHs) in the
sediments of Olbia harbor (Northern Sardinia, Otaly). Marine Pollution Bulletin 50,
1223-1232.

120

Fang, M., Zheng, M., Wang, F., To, K.L., Jaafar, A.B., Tong, S.L., 1999. The solventextractable organic conpounds in the Indonesia biomass burning aerosols
characterization studies. Atmospheric Environment 33, 783-795.
Fernandes, M.B., Sicre, M.-A., Boireau, A., Tronczynski, J., 1997. Polyaromatic hydrocarbon
(PAH) distributions in the Seine river and its estuary. Marine Pollution Bulletin, 34(11):
857-867.
Gauthier, T.D., Seltz, W.R., Grant, C.L., 1987. Effects of structural and compositional
variations of dissolved humic materials on pyrene Koc values. Environmental Science and
Technology, 21(3), 243-248.
Ghosh, U., Gillette, J.S., Luthy, R.G., Zare, R.N., 2000. Microscale location, characterization,
and association of polycyclic aromatic hydrocarbons on harbor sediment particles.
Environmental Science & Technology 34, 1729-1736.
Ghosh, U., Zimmerman, J.R., Luthy, R.G., 2003. PCB and PAH speciation among particle types
in contaminated harbor sediments and effects on PAH bioavailability. Environmental
Science & Technology 37, 2209-2217.
Grathwohl, P., 1990. Influence of organic matter from soils and sediments from various origins
on the sorption of some chlorinated aliphatic hydrocarbons: iplication on Koc correlations.
Environmental Science and Technology 24, 1687-1693.
Gustafsson, ., Haghseta, F., Chan, C., Macfarlane, J., Gschwend, P.M., 1997. Quantification
of the dilute sedimentary soot phase: implications for PAH speciation and bioavailability.
Environmental Science and Technology 31, 203-209.
Gustafsson, ., Haghseta, F., Chan, C., Macfarlane, J., Gschwend, P.M., 1997. Quantification
of the dilute sedimentary soot phase: implications for PAH speciation and bioavailability.
Environmental Science and Technology 31, 203-209.
Holtvoeth, J., 2004. Terrigeneous organic matter in sediments of the eastern equatorial Atlantic
distribution, reactivity, and relation to Late Quaternary climate -. PhD Dissertation,
Fachberiech Geowissenschaften, der Universitt Bremen, Germany.
Holtvoeth, J., Kolonic, S., Wagner, T., 2005. Soil organic matter as an important contributor to
late quaternary sediments of the tropical West African continental margin. Geochimica et
Cosmochimica Acta 69(8), 2031-2041.
Horowitz, A.J., Elrick, K.A., 1987. The relation of stream sediment surface area, grain size and
composition to trace element chemistry. Applied Geochemistry, 2, 437-451.
IARC, 1987. IARC Monographs on the evaluation of carcinogenic risks to humans, Suppl. 7,
overall evaluations of carcinogenicity: an updating of IARC Monographs volumes 1 to
42, Lyon, IARC Press.
ICES, 1997. Determination of polycyclic aromatic hydrocarbons (PAHs) in sediments:
Analytical methods. In Report of the ICES Advisory Committee on the Marine
Environment, 1997. ICES Cooperative Research Report 222, 118-124.
Jenkins, B.M., Jones, A.D., Turn, S.Q., Williams, R.B., 1996. Emission factors for polycyclic
aromatic hydrocarbons from biomass burning. Environmental Science and Technology,
30, 2462-2469.
Jonker, M.T.O., Koelmans, A.A., 2002. Sorption of polycyclic aromatic hydrocarbons and
polychlorinated biphenyls to soot and soot-like materials in the aqueous environment :
mechanistic considerations. Environmental Science and Technology, 36, 3725-3734.
Kakareka, S.V., Kukharchyk, T.I., 2003. PAH emission from the open burning of agricultural
debris. The Science of the Total Environmental, 308, 257-261.
Li, C-T., Mi, H-H., Lee, W-J., You, W-C., Wang, Y-F., 1999. PAH emission from the industrial
boilers. Journal of Hazardous Materials, A69, 1-11.
Liu, H., Amy, G., 1993. Modeling partitioning and transport interactions between natural
organic matter and polynuclear aromatic hydrocarbons in groundwater. Environmental
Science and Technology, 27, 1553-1582.
Liu, W.X., Dou, H., Wei, Z.C., Chang, B., Qiu, W.X., Liu, Y., Tao, S., 2009. Emission
characteristics of polycyclic aromatic hydrocarbons from combustion of different
residential coals in North China. Science of the Total Environment 407, 1436-1446.

121

Luthy, R. G., Aiken, G.R., Brusseau, M.L., Cunningham, S.D., Gschwend, P.M., Pignatello,
J.J., Reinhard, M., Traina, S.J., 1997. Sequestration of hydrophobic organic contaminants
by geosorbents. Environmental Science and Technology, 31, 3341-3347.
Marr, LC., Kirchstetter, T.W., Harley, R.A., Miguel, A.H., Hering, S.V., Hammond, S.K., 1999.
Characterization of polycyclic aromatic hydrocarbons in motor vehicle fuels and exhaust
emissions. Environmental Science and Technology, 33, 3091-3099.
McGroddy, S.E., Farrington, J.W., 1995. Sediment porewater partitioning of polycyclic
aromatic hydrocarbons in three cores from Boston Harbor, Massachussetts.
Environmental Science and Technology, 29, 1542-1550.
Meyer, P.A., 1994. Preservation of elemental and isotopic source identification of sedimentary
organic matter. Chemical Geology 114, 289-302.
Neff, J.M., 1979. Polycyclic aromatic hydrocarbon in the aquatic environment: sources, fates
and biological effects. Applied Science Publisher Ltd, Essex, UK, 262 pp.
Oanh, N.T.K., Albina, D.O., Ping, L., Wang, X., 2005. Emission of particulate and polycyclic
aromatic hydrocarbons from selected cookstove fuel systems in Asia. Biomass and
Bioenergy, 28, 579-590.
Page, S.E., Siegert, F., Rieley, J.O., Boehm, H-D. V., Jaya, A., Limin, S., 2002. The amount of
carbon released from peat and forest fires in Indonesia during 1997. Nature 420, 61-65.
Pillon, P., Jocteur-Monrozier, L., Gonzalez, C., Saliot, A., 1986. Organic geochemistry of recent
equatorial deltaic sediments. Organic Geochemistry, 10, 711-716.
Prahl, F.G., Carpenter, R., 1983. Polycyclic aromatic hydrocarbons (PAH) phase associations in
Washington Coastal sediments. Geochimica et Cosmochimica Acta 47, 1013-1023.
Requejo, A.G., Sassen, R., McDonald, T., Denoux, G., Kennicutt II, M.C., Brooks, J.M., 1996.
Polynuclear aromatic hydrocarbons (PAH) as indicators of the source and maturity of
marine crude oils. Organic Geochemistry 24(10/11), 1017-1033.
Rockne, K.J., Shor, L.M., Young, L.Y., Taghon, G.L., Kosson, D.S., 2002. Distribution
sequestration and release of PAHs in weathered sediment: the role of sediment structure
and organic carbon properties. Environmental Science and Technology 36, 2636-2644.
Ruttenberg, K.C., Go i, M.A., 1997. Phosphorus distribution, C:N:P ratios, and 13Coc in
arctic, temperate, and tropical coastal sediments: tools for characterizing bulk
sedimentary organic matter. Marine Geology 139, 123-145.
Saeed, T., Al-Mutairi, M., 2000. Comparative composition of polycyclic aromatic hydrocarbons
(PAHs) in the sea water-soluble fractions of different Kuwaiti crude oils. Advances in
Environmental Research 4, 141-145.
Sargenti, S.R., McNair, H.M., 1998. Comparison of solid-phase extraction and supercritical
fluid extraction for extraction of polycyclic aromatic hydrocarbons from drinking water.
Journal of Microcolumn Separation 10(1), 125-131.
See, S.W., Balasubramanian, R., Rianawati, E., Karthikeyan, S., Streets, D.G., 2007.
Characterization and source apportionment of particulate matter  2.5 m in Sumatra,
Indonesia during a recent peat fire episode. Environmental Science and Technology, 41,
3488-3494.
Shi, Z., Tao, S., Pan, B., Fan, W., He, H.C., Zuo, Q., Wu, S.P., Li, B.G., Cao, J., Liu, W.X., Xu,
F.L., Wang, X.J., Shen, W.R., Wong, P.K., 2005. Contamination of rivers in Tianjin,
China by polycyclic aromatic hydrocarbons. Environmental Pollution, 134: 97-111.
Siegel, H., Stottmeister, I., Reimann, J., Gerth, M., Jose, C., Samiaji, J., 2009. Siak river
system-east Sumatra characterisation of sources, estuarine processes, and discharges into
the Malacca Strait. Journal of Marine Systems 77, 148-159.
Soclo, H.H., Garrigues, PH., Ewald, M., 2000. Origin of polycyclic aromatic hydrocarbons
(PAHs) in coastal marine sediments: case studies in Cotonou (Benin) and Aquitaine
(France) areas. Marine Pollution Bulletin 40(5), 387-396.
Vila-Escal, M., Vegas-Villarubia, T., Prat, N., 2007. Release of polycyclic aromatic
compounds into a Mediterranean creek (Catalonia, NE Spain) after a forest fire. Water
Research, 41: 2171-2179.

122

Wang, Z., Fingas, M., Lambert, P., Zeng, G., Yang, C., Hollebone, B., 2004. Characterization
and identification of the Detroit River mystery oil spill (2002). Journal of
Chromatography A 1038, 201-214.
Wang, Z., Fingas, M., Page, D.S., 1999. Oil spill identification. Journal of Chromatography A
843, 369-411.
Yang, H-H., Lee, W-J., Chen, S-J., Lai, S-O., 1998. PAH emission from various industrial
stacks. Journal of Hazardous Materials 60, 159-174.
Yu, Y., Xu, J., Wang, P., Sun, H., Dai, S., 2009. Sediment-porewater partition of polycyclic
aromatic hydrocarbons (PAHs) from Lanzhou Reach of Yellow River China. Journal of
Hazardous Materials 165, 494-500.
Yunker, M.B., Macdonald, R.W., Vingarzan, R., Mitchell, H., Goyette, D., Sylvestre, S., 2002.
PAHs in the Fraser River basin: a critical appraisal of PAH ratio as indicators of PAH
source and composition. Organic Geochemistry 33, 489-515.
Zakaria, M.P., Horinouchi, A., Tsutsumi, S., Takada, H., Tanabe, S., Ismail, A., 2000. Oil
pollution in the Straits of Malacca, Malaysia: Application of molecular markers for
source identification. Environmental Science and Technology, 34, 1189-1196.
Zhang, Y., Tao, S., 2009. Global atmospheric emission inventory of polycyclic aromatic
hydrocarbons (PAHs) for 2004. Atmospheric Environment, 43, 812-819.
Zhou, J., Wang, T., Huang, Y., Mao, T., Zhong, N., 2005. Size distribution of polycyclic
aromatic hydrocarbons in urban and suburban sites of Beijing, China. Chemosphere 61:
792-799.

123

Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Chrysene
Benzo(a)anthracene
Benzo(a)pyrene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(g,h,i)perylene
Indeno(1.2.3-cd)pyrene
Dibenzo(a,h)anthracene

Compounds

C10H8
C12H8
C12H10
C13H10
C14H10
C14H10
C16H10
C16H10
C18H12
C18H12
C20H12
C20H12
C20H12
C22H12
C22H12
C22H14

Molecular
Formula

128,17
152,20
154,21
166,22
178,23
178,23
202,26
202,26
228,29
228,29
252,32
252,32
252,32
276,34
276,34
278,35

Mol.
Weighta,b
(g/mol)

91-20-3
208-96-8
83-82-9
86-73-7
85-01-8
120-12-7
206-44-00
129-00-0
218-01-9
56-55-3
50-32-8
205-99-2
207-08-9
191-24-2
193-39-5
53-70-3

CAS
number a,b

218
256-275
279
293-295
340
340
375
393
448
435
495*
n.a.
480
542*
260*
n.a.

Boiling
Point a
(BP oC)

80.5
92-93
96.2
116-117
101
214-216
151*
156
255-256
162
176.4*
162-165*
217
277-279*
160*
269-270

Melting
Point a
(MP oC)

30.2
16.1a
3.93
1.90
1.18
0.076
0.260
0.135
0.0015
0.011
0.0038
0.0140
0.0080
0.0003
0.0005
0.0005

Water
Solubilityd
mg/L

8.89E-02
2.90E-02
3.75E-03
3.24E-03
6.80E-04
2.55E-05
8.13E-06
4.25E-06
7.80E-09
1.54E-07
4.89E-09
8.06E-08
9.59E-11
1.00E-10
1.40E-10
2.10E-11

Vapor
Pressurea
(mm Hg)

4.50E-03
n.a.
2.40E-04
7.40E-05
2.70E-04
1.80E-06
1.95E-03
1.30E-05
n.a.
1.20E-06
7.40E-05
n.a.
2.70E-07
2.00E-07
n.a.
2.00E-09

3.45
3.98e
4.22
4.38
4.46
4.54
5.20
5.30
5.61
5.91
6.35
5.78
6.20
6.90
6.51
6.75

Log Kow

(atm m3/mol)

KH

Octanol-Water
Partition
d
Coefficient

Henrys Law
e
Constant

n.a.
= data not available; a = Weast & Astle, 1985; b = Bojes & Pope, 2007; c = Schwarzenbach et al. 1993; d = Williamson et al., 2002; e = Martinez et al., 2004;
*
= Beilstein Database from various references as follows:
Fluoranthene (MP): Ref. 2, 1718469; Journal; Hershberg bei Fieser; Novello; JACSAT; Journal of the American Chemical Society; 62; 1940; 1855,1859 Anm. 23; DOI:
10.1021/ja01864a061; ISSN: 0002-7863.
Benzo(a)pyrene (BP): Ref. 1, 4467311; Journal; Lewis; Edstrom; JOCEAH; Journal of Organic Chemistry; 28; 1963; 2050,2053; ISSN: 0022-3263.; (MP): Ref. 9, 4258666; Journal;
Murray et al.; CJCHAG; Canadian Journal of Chemistry; 52; 1974; 557; DOI: 10.1139/v74-087; ISSN: 0008-4042.
Benzo(b)fluoranthene (MP): Ref. 2, 159677; Journal; Badger,G.M.; Spotswood,T.M.; JCSOA9; Journal of the Chemical Society; English; 1960; 4420 - 4427; DOI:
10.1039/jr9600004420; ISSN: 0368-1769.
Benzo(g,h,i)perylene (BP): Ref. 3, 6482028; Journal; Gonzalez, Maykel Perez; Toropov, Andrey A.; Duchowicz, Pablo R.; Castro, Eduardo A.; MOLEFW; Molecules; English; 9; 12;
2004; 1019 - 1033; ISSN: 1420-3049.; (MP): Ref. 9, 4276855; Journal; Aronson; Katlowitz; JINCAO; Journal of Inorganic and Nuclear Chemistry; 41; 1979; 1579; DOI: 10.1016/00221902(79)80180-3; ISSN: 0022-1902.
Indeno(1,2,3-cd)pyrene (BP): Ref. 1 4623412; Journal; Studt; JLACBF; Justus Liebigs Annalen der Chemie; 1978; 530; ISSN: 0075-4617.; (MP): Ref. 3 4467311; Journal; Lewis;
Edstrom; JOCEAH; Journal of Organic Chemistry; 28; 1963; 2050,2053; ISSN: 0022-3263.

1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.

No.

Appendix 1. Physical & chemical properties of 16 PAHs of the US EPA priority pollutants.

124

Appendix 2. Emission factors and isomeric ratios of PAHs from various pyrogenic
and petrogenic sources (literature data)
Appendix 2.1. Emission factors of PAHs (mg/kg fuel) from various coal, biomass, and
petroleum combustions
Sources

Emission
Factor
(mg/kg fuel)

Remarks

References

Cu addition at 800oC
Cupric oxide, CuO, addition at
1000oC
Cookstove in Southeast Asia
Domestic heating
Residential heating coal source:
the Jinxi coal mine in Beijing
Coal source: the Datong coal
mine in Shanxi province
coal source: the Jinxi coal mine in
Beijing
Coal source: the Datong coal
mine in Shanxi province
Coal source: the Datong coal
mine in Shanxi province
Samples from Elmsworth
gasfield, Canada
Elmsworth gasfield, Canada
Ruhr basin, Osterfeld Germany
Ruhr basin, Hugo Germany
Ruhr basin, Westerholt Germany
Ruhr basin, Blumenthal Germany

Yan et al 2004

Coal
Bituminous Coal

121

Bituminous Coal

248

Coal Briquette
Bituminous Coal

101
109

Bituminous Coal

1435

Bituminous Coal

1096

Anthracite Coal

77,8

Anthracite Coal

151

Honeycomb Briquette

812

Bituminous Coal

152

Bituminous Coal
Bituminous Coal
Bituminous Coal
Bituminous Coal
Bituminous Coal

137
153
124
164
155

Yan et al 2004
Oanh et al., 1999
Lee et al., 2005
Liu et al 2009

Willsch & Radke, 1995 cited in


Achten & Hofmann, 2009

Biomasses
Almond

14,1

Walnut
Pine
Fir
Chinese clay woodstoves
Wood
Eucalyptus wood 1
Oak Wood
Lao traditional wood chips
Thai bucket wood chips
Cambodian traditional wood
chips
Vietnam traditional wood
chips
Eucalyptus wood
Paper

23,1
26,7
28,0
23,3
36,8
33,5
38,0
41,2
50,2

Woods
Wind tunnel simulation of open
burning

Domestic heating
Domestic burning
Residential fireplace
Cookstove fuel system in Asia

Jenkins et al.,1996
Jenkins et al.,1996
Jenkins et al.,1996
Jenkins et al.,1996
Oanh et al., 2005
Lee et al., 2005
Schauer et al., 2001
Schauer et al., 2001
Oanh et al., 2005
Oanh et al., 2005

53,4

Oanh et al., 2005

56,4

Oanh et al., 2005

110
68,8

Open burning in Southeast Asia


Joss paper furnace

Oanh et al., 1999


Yang et al., 2005

Grasses / debris
Wind tunnel simulation of open
burning

Rice grasses

15,1

Jenkins et al.,1996

Corn grasses
Agricultural debris
Petroleum
Natural Gas
Diesel
Heavy oil
Fuel Oil
Diesel
Diesel*
*mg/km

45,3
410

Open burning

Jenkins et al.,1996
Kakareka & Kukharchyk, 2003

2,82
2,86
13,6
3,51
4,66
5,84

Industrial boiler
Industrial boiler
Industrial boiler
Industrial stacks, heavy oil plant
Heavy-duty Vehicles
Light-duty Diesel Vehicles

Li et al., 1999
Li et al., 1999
Li et al., 1999
Yang et al., 1998
Marr et al., 1999
de Abrantes et al., 2004

125


Appendix 2.2. The isomeric ratios for source apportionment from various organic matter
combustion/burnings

Sources

LMW/HMW

MW 178

MW 202

MW 228

MW 276

References

Coal and Petroleum combustions


Coal- Cement
Industry
Bituminous Coal
Charcoal
Coal Briquette
Bituminous Coal
Diesel, Industrial
Boiler
Fuel Oil, Heavy-oil
Plant
Diesel Heavy-duty
Vehicles
Heavy oil, Industrial
Boiler
Light-duty Diesel
vehicles

18

0,10

0,24

0,58

0,65

Yang et al., 1998

10
7,3
15
5,1

0,12
0,21
0,55
0,40

0,21
0,33
0,32
0,93

0,35
0,37
0,15
0,49

0,31

0,36

Yang et al., 1998


Oanh et al., 1999
Oanh et al., 1999
Lee et al., 2005

1,1

0,28

0,91

0,61

0,28

Li et al., 1999

6,0

0,14

0,24

0,35

0,35

Yang et al., 1998

0,43

0,56

1,0

Marr et al., 1999

6,5

0,25

0,52

0,53

0,70

Li et al., 1999

1,7

0,04

0,93

0,87

de Abrantes et al.,
2004

Jenkins et al.,1996
Jenkins et al.,1996

Biomass Burnings
Almond
Walnut
Chinese clay
woodstoves
Pine-domestic
heating
Fir
Eucalyptus wood
Wood, Domestic
Heating
Oak Wood
Lao traditional wood
chips
Thai bucket wood
chips
Cambodian
traditional wood
chips
Vietnam traditional
wood chips
Eucalyptus wood
Joss Paper
Bamboo

8,3
8,6

Woods/Bamboo
0,14
0,54
0,16
0,57

0,51
0,43

4,3

0,25

0,79

0,56

8,8

0,14

0,56

0,53

6,1
2,4

0,16
0,18

0,55
0,56

0,54
0,48

0,00
0,49

Jenkins et al.,1996
Schauer et al., 2001

2,6

0,20

0,52

0,53

0,57

Yang et al., 2005

2,5

0,19

0,57

0,45

2,4

0,26

0,77

0,51

0,64

Oanh et al., 2005

4,0

0,28

0,79

0,39

0,82

Oanh et al., 2005

2,9

0,24

0,81

0,53

0,77

Oanh et al., 2005

4,2

0,21

0,80

0,40

0,88

Oanh et al., 2005

7,4
14
0,2

0,48
0,11
0,50

0,69
0,56
1,00

Oanh et al., 1999


Yang et al., 2005
Oros et al., 2006

0,42
0,52
0,68
0,47
0,40

1,00

0,50
0,94

Oros et al., 2006


Oros et al., 2006
Oros et al., 2006
Jenkins et al.,1996
Jenkins et al.,1996
Kakareka &
Kukharchyk, 2003

0,59

Oanh et al., 2005


Jenkins et al.,1996

Schauer et al., 2001

Sugarcane
Pampas grass
Mixed ryegrass
Rice grasses
Corn grasses

0,4
0,2
0,7
10
0,3

0,29
0,67
0,04
0,65
0,30
0,53
Grasses/debris/peats
0,23
0,53
0,16
0,43
0,21
0,44
0,15
0,59
0,04
0,50

Agricultural Debris

3,4

0,23

0,71

0,69

0,26

0,7

0,16

0,38

0,73

0,62

See et al 2007

0,7

0,79

0,49

0,90

0,68

See et al 2007

Peat Burning Dumai


(ng/m3)
Peat Smoke
Pekanbaru (ng/m3)

126


Appendix 2.2.
Sources

LMW/HMW

MW 178 MW 202 MW 228


Crude oil / Oil Spills

MW 276

References

Alaskan North Slope


10
0,03
0,26
0,10
0,18
Requojo et al. 1996
Crude (ppm)
Oil spills "Oil Tanker
1,0
0,15
0,15
0,37
0,21
Baars et al., 2002
Enrika" (mg/kg)
Oil spills "Enrika"
0,7
0,13
0,11
0,31
0,37
Beach Sample1
Oil spills "Enrika"
0,4
0,14
0,09
0,31
0,47
Beach Sample2
Detroil Oil Spill_1
1,3
0,10
0,57
0,51
0,47
Wang et al 2004
(g/g TSEM*)
Detroil Oil Spill_2
1,5
0,11
0,57
0,50
0,47
Detroil Oil Spill_3
1,0
0,05
0,61
0,51
0,49
LMW/HMW = sigma low molecular weight compounds (2-3 rings) / sigma high molecular weight compounds (4-6
rings)

MW 178= the ratio of anthracene /( anthracene + phenanthrene)


MW 202 = the ratio of fluoranthene /( fluoranthene + pyrene)
MW 228 = the ratio of benzo(a)anthracene/( benzo(a)anthracene + chrysene)
MW 276 = the ratio of indeno(1,2,3-cd)pyrene /( indeno(1,2,3-cd)pyrene + benzo(g,h,i)perylene)
*TSEM = total solvent-extractable materials

127


Dilution Factor
Volume standard (L)
Acetonitrile (L)

1000

1000
2000
1000
200
100
100
200
100
100
100
200
100
100
200
200
100

Dilution Fractor (DF)


Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene

Making a dilution

Stock

400
3600

10
100
200
100
20
10
10
20
10
10
10
20
10
10
20
20
10

Sub-Stock

10
400
3600

1
100
10
20
10
2
1
1
2
1
1
1
2
1
1
2
2
1

3
400
2,5
5
2,5
0,5
0,25
0,25
0,5
0,25
0,25
0,25
0,5
0,25
0,25
0,5
0,5
0,25

Prepared from Sub-Stock


20
40
200
100
3800
3900

2
200
5
10
5
1
0,5
0,5
1
0,5
0,5
0,5
1
0,5
0,5
1
1
0,5

80
50
3950

4
800
1,25
2,5
1,25
0,25
0,125
0,125
0,25
0,125
0,125
0,125
0,25
0,125
0,125
0,25
0,25
0,125

Prepared from CAL-2


5
10
20
800
400
200
3200
3600
3800

Calibration standard (CAL-#)


5
6
7
1000
2000
4000
1
0,5
0,25
2
1
0,5
1
0,5
0,25
0,2
0,1
0,05
0,1
0,05
0,025
0,1
0,05
0,025
0,2
0,1
0,05
0,1
0,05
0,025
0,1
0,05
0,025
0,1
0,05
0,025
0,2
0,1
0,05
0,1
0,05
0,025
0,1
0,05
0,025
0,2
0,1
0,05
0,2
0,1
0,05
0,1
0,05
0,025

9
20000
0,05
0,1
0,05
0,01
0,005
0,005
0,01
0,005
0,005
0,005
0,01
0,005
0,005
0,01
0,01
0,005

10
40000
0,025
0,05
0,025
0,005
0,0025
0,0025
0,005
0,0025
0,0025
0,0025
0,005
0,0025
0,0025
0,005
0,005
0,0025

Prepared from CAL-7


2.5
5
10
1600
800
400
2400
3200
3600

8
10000
0,1
0,2
0,1
0,02
0,01
0,01
0,02
0,01
0,01
0,01
0,02
0,01
0,01
0,02
0,02
0,01

= 16 PAH Mix 61 (US EPA 16) in 1 mL acetone/methanol (1:1) obtained from Dr. Ehrenstorfer GmbH, Augsburg, Germany
= acetonitrile (HPLC grade)
= ng/l

Compound

Standard Stock
Dilution solvent
Concentration unit

Appendix 3.1. Dilution system of 16 PAH reference standards for calibration

128

Appendix 3.2.1. Percent recovery of surrogate standards used for compensating the
procedural efficiency and reproducibility in surface sediment fractions.

Sample Station Sediment


d10-phenanthrene d10-fluoranthene d12-perylene
Fraction (m)
River

S 24
S 101
S 20
S 104
S 105
S 35
S 116
S 42
S 145

> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63

74,1
79,5
83,2
74,1
99,2
86,9
81,1

109
98,5
117
94,0
113
103
105

100
90,8
90,5
95,0
94,9
101
88,5

40,6

117

85,0

72,5

82,3

88,5

78,1
69,7
94,1
109
76,3
81,2

105
99,4
84,2
105
102
108

87,3
119
73,9
99,7
92,4
83,9

> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63

64,5
77,0
86,3
75,7
66,5
92,2
83,2
93,4
82,8
82,8
69,5
82,4
113
99,1
81,2
82,8

93,9
105
119
93,0
72,9
93,6
108
76,7
97,5
97,5
77,8
103
109
87
115
109

82,5
94,3
96,7
94,5
75,8
86,2
104
106
95,1
109
89,6
82,7
121
119
87,5
107

> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63
> 63
< 63

73,1
68,0
82,8
82,8
82,8
82,8
77,2
79,4
94,2
70,3
97,3
84,0
82,8
82,8
114
96,8
95,2
82,6
83,6
78,4

70,2
69,4
92,0
141
88,4
99,9
94,9
87,3
104
86,4
90,5
117
114
89,0
87,0
97,1
81,1
97,9
91,3
74,7

102
103
75,6
120
81,5
104
99,7
86,7
88,8
87,3
107
118
91,6
90,4
79,8
83,2
83,2
99,4
95,5
84,9

Estuary
S 143
S 142
S 138
S 134
S 252
S 125
S 250
S 251
Coast
S 269
S 226
S 227
S 228
S 267
S 266
S 253
S 230
S 231
S 232

129


Appendix 3.2.2. Percent recovery of surrogate standards used for compensating the
procedural efficiency and reproducibility in SPM and SPE.

Sample Station d10-phenanthrene d10-fluoranthene d12-perylene


SPM
River

Estuary

Coast

SPE

S102
S105
S115
S205
S272
S218
S216
S 291
S 301
S 139
S 140
S 137
S 134
S 133
S 132
S 124
125
126
128
130
S 225
S 226
S 227
S 228
S 267
S 266
S 233
S 230
S 231
S 232
S 229
S 251
S 291
S 370
S 301
S 305
S 307
S 312
S 314
S 316
S 317
S 318
S 324

75,6
118
80,2
84,2
101
103
109
85,5
103
103
108
102
93,9
89,2
79,0
98,5
102
91,0
101
87,2
106
88,2
107
89,7
87,2
104
101
96,1
88,5
85,2
86,2
93,3

93,3
91,1
85,1
114
102
96,1
103
103
103
73,5
88,8
85,2
118
86,8
98,1
84,8
112
92,0
86,3
105
94,7
105
96,5
95,7
90,9
102
107
108
88,5
90,0
95,9
104

79,4
99,5
79,4
87,7
95,4
92,2
92,2
88,4
96,6
92,3
90,3
82,4
113
101
97,4
93,4
115
93,7
95,5
104
106
96,1
105
98,4
105
96,1
101
101
76,0
94,8
101
93,2

85,4
87,8

85,3
97,1

98,2
96,7
88,0
95,9
88,0
94,6
103
79,9

88,4
77,2
101
97,9
81,2
91,5
96,8
71,5

130


Appendix 4. Tables for the Manuscript in the Chapter IV


Table 1. Sampling stations and properties of the sediment fractions
Station

Sand

Mud

Percentage Organic
Total
C/N
Percentage Organic
Total
C/N
of total
Carbon Nitrogen (mol/mol)
of total
Carbon Nitrogen (mol/mol)
sediment
(%)
(%)
sediment
(%)
(%)
River
S 24
17.2
2.34
0.30
9.12
S 101
45.5
1.71
0.25
8.11
S 20
76.6
0.29
0.01
41.7
S 104
99.0
0.05
0.003
18.5
S 105
98.4
0.14
0.008
19.4
S 35
87.3
0.22
0.01
24.3
S 116
34.1
3.52
0.34
11.9
S 42
6.06
14.0
0.62
26.2
S 145
69.0
2.08
0.11
22.5
Estuary
S 143
33.0
4.04
0.36
12.9
S 142
24.2
24.0
1.14
24.6
S 138
12.3
9.07
0.64
16.5
S 134
16.3
4.78
0.45
12.4
S 252
59.1
1.25
0.11
12.7
S 125
33.5
1.96
0.21
10.7
S 250
20.8
7.29
0.55
15.5
S 251
91.8
0.22
0.03
9.08
Coast
S 269
6.78
0.11
0.01
24.0
S 226
18.8
0.23
0.01
28.1
S 227
28.5
2.66
0.07
42.2
S 228
18.0
2.37
0.06
43.8
S 267
64.7
0.01
0.001
11.1
S 266
55.0
0.30
0.01
49.2
S 253
24.1
0.28
0.01
32.9
S 230
3.75
0.05
0.004
17.4
S 231
7.25
1.13
0.04
30.6
S 232
4.58
6.72
0.16
47.9
n.a. = data not available due to insufficient sediment

77.8
45.3
16.8
0.40
1.11
2.22
62.6
63.3
28.4

3.08
2.39
2.98
n.a.
n.a.
n.a.
3.70
3.24
1.41

0.75
0.40
0.29
n.a.
n.a.
n.a.
0.26
0.38
0.20

4.82
6.91
11.8
n.a.
n.a.
n.a.
16.4
9.89
8.32

60.4
70.3
70.3
72.3
37.4
59.7
55.4
4.14

2.25
1.27
1.46
1.61
2.04
1.13
3.31
1.82

0.13
0.16
0.19
0.11
0.27
0.09
0.35
0.08

20.4
9.53
8.85
17.6
8.67
14.7
10.9
27.2

84.3
72.8
65.9
75.8
29.0
26.9
69.5
65.5
69.2
83.0

1.05
0.41
0.34
0.46
1.19
1.15
1.44
1.29
2.18
1.71

0.09
0.04
0.04
0.05
0.10
0.09
0.09
0.09
0.14
0.11

13.4
12.2
9.80
11.7
13.8
14.8
18.7
16.2
18.8
18.6

131

Table 2. Median and range of PAH contents (ng g-1 d.w.) in the global sediment fractions from
the Siak River, estuary and the coastal areas. The PAHs are ordered by following their elution
time. PAHs Bulk was calculated from the fraction contents
Compound

River
Median Range

Estuary
Median Range

Coast
Median Range

SAND
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene
PAHs

48,8
108
23,5
7,96
20,2
1,65
31,2
45,2
7,86
10,1
14,9
4,97
2,88
129
4,52
7,23
556

14.2 310
9.68 4235
1.93 155
1.92 - 89.0
9.05 321
0.58 - 4.49
9.13 138
7.32 335
1.98 - 31.4
3.98 - 47.0
2.06 278
1.35 - 34.3
1.01 - 32.5
4.09 580
0.48 - 40.1
0.39 - 18.1
164 5474

45,8
104
23,3
7,19
18,8
3,58
45,9
20,7
10,3
18,1
25,3
10,4
15,9
70,4
5,46
17,7
425

13.0 221
80.7 2085
7.24 - 83.5
2.83 - 26.6
10.3 - 78.0
1.52 - 24.5
12.2 119
4.54 114
0.93 - 42.6
1.26 - 49.8
6.75 137
1.47 - 76.7
1.31 210
8.63 885
0.37 - 43.3
0.77 - 71.8
208 3913

42,0
805
14,5
1,87
7,73
1,10
59,3
13,7
5,79
9,59
17,9
2,22
8,30
26,7
10,5
3,90
1442

3.24 - 84.0
426 - 1780
5.50 - 133
0.68 - 4.99
3.51 - 31.6
0.07 - 15.4
8.99 - 378
1.20 - 94.7
0.31 - 81.6
0.15 - 114
0.84 - 179
0.07 - 50.5
0.87 - 249
1.42 - 1061
0.47 - 194
0.30 - 37.1
594 - 2495

PAHs

63,0
144
18,4
3,84
16,4
1,40
35,2
30,6
5,06
8,60
17,8
5,86
4,31
198
7,87
7,86
521

23.3 - 67.4
38.6 836
3.82 - 40.9
1.90 - 15.1
5.27 - 25.2
0.69 - 2.17
18.7 - 97.9
11.5 - 93.6
3.38 - 18.6
3.91 - 52.0
2.59 - 79.7
1.90 - 17.3
0.32 - 11.8
0.98 405
4.03 - 16.0
1.81 - 11.7
319 1143

34,8
162
24,9
3,54
13,4
1,77
26,0
14,3
3,88
8,29
14,9
4,57
5,52
38,7
5,21
3,77
468

12.5 - 61.1
9.42 314
2.23 - 65.1
1.57 - 4.82
6.44 - 34.3
0.53 - 6.31
8.58 106
2.19 - 38.0
1.08 - 7.68
2.37 - 12.8
5.56 - 38.9
0.05 - 10.9
2.45 - 21.1
6.10 206
1.82 - 16.7
2.09 - 13.8
126 584

28,6
403
10,7
2,83
6,20
0,60
34,4
10,9
2,55
5,51
8,86
1,79
4,12
15,9
9,94
4,24
633

15.3 - 41.9
257 - 762
1.60 - 70.7
1.08 - 3.93
0.94 - 23.8
0.09 - 3.35
3.73 - 228
2.29 - 40.4
1.24 - 40.0
0.67 - 59.1
2.89 - 51.8
0.88 - 7.97
0.45 - 31.7
4.18 - 188
0.70 - 48.9
0.09 - 16.8
443 1314

PAHs Bulk

484

161 1055

443

145 880

577

454 1234

MUD
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene

132

Table 3. Summary of PAHs (ng g-1 d.w.) from different river, estuarine and coastal sediments
in several Asian and European countries. The contents were mainly derived from bulk sediment.
The sediment fraction contents from this study are included as comparison
Locations
Asia
Indonesia, Riau Province
Siak river, estuary and the
coastal areas
Siak river
Estuary
Riau Coast
Malaysia
Rivers
Estuaries
Coastal areas
Malacca Straits
Thailand, Chao Praya River and the
estuary
River
Estuary
Whole Thai Coast
China
Minjiang River estuary,
China
Tonghui River, Beijing,
China
Pearl River and Estuary,
China
Middle and lower Yellow
River, China
Daya Bay, China
Jiulong River Estuary, China
Korea
Mouth of Han river,
Kyeonngi Bay, Korea
Masan Bay, Korea
India, Yamuna River, Delhi, (river
bank sediment)
Europe
Italy, Porto Torres harbor, Northern
Sardinia
Olbia harbor, Northern
Sardinia, Italy
Lagoon of Venice, Italy

Median*

Content
range

16

837

164 - 5474

16

493

126 - 1314

16
16
16
15

484
443
577

161 - 1055
145 - 880
454 - 1234

75
257
24
8.5

20 - 924
19 - 431
9 - 39
4 - 73

Sampling
year

PAH

2004 - 2006

1998, 1999

2003, 2004

Reference

Sand fraction, this


study
Mud fraction, this
study
Bulk, this study
Bulk, this study
Bulk, this study
Zakaria et al., 2002

Boonyatumanond et al.,
2006

17
251
88
36

33 - 594
30 - 724
6 - 228

1999

16

433**

112 877

Zhang et al, 2004

2002

16

540**

127 928

Zhang et al, 2004

2002

18

279

189 - 637

Luo et al., 2006

13

91

31 - 133

Li et al., 2006

1999
1999

16
16

481**
238

115 1134
59 - 1177

Zhou & Maskaoui, 2003


Maskaoui et al. 2002

1995

18

83*

29 230

Kim et al, 1999

1997

18

680**

Yim, et al, 2005

2003

16

9490

207 2670
4500 23530

1999

16

740

70 1210

De Luca et al, 2004

n.a

16

315

160 - 770

De Luca et al., 2005

1987
1993
1998
1999
1996

13
13
13
16
15

136
177
131
631

20 502
23 570
23 - 532
24 4710

Secco et al, 2005

2004

The UK, Brighton marina


Coastal areas of the Adriactic Sea
Chioggia
367
24 501
Ancona
192
34 - 307
Coastal areas of the Black Sea
1995
18
61
7 - 638
*) own calculation from the literature data
**) mean value (unavailable data from given literatures for median calculation)

Agarwal et al., 2006

King et al, 2004


Magi et al., 2002

Readman et al., 2002

133

355
123
ND
9.26
66.3
10.5
91.8
25.5
25.9
10.8
45.2
13.5
16.5
12.7
10.4
7.92
824

00 33,531
101 23,712

ND = not detected; <DL = below detection limit

Compound
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene
PAHs

Latitude
Longitude

S 291

57.5
ND
ND
3.91
52.3
4.07
58.7
22.8
11.3
5.43
15.1
3.85
4.71
5.54
2.73
1.68
250

00 35.857
101 35.410

S 370

89.1
189
ND
23.0
385
11.4
275
41.6
5.33
4.43
4.28
1.32
2.66
2.16
1.95
ND
1035

00 42, 684
101 40, 122

River
S 305

212
313
92.9
72.4
1262
110
2505
402
58.5
48.9
29.0
0.26
10.8
22.5
1.19
ND
5140

00 45, 760
101 47, 685

S 301

8.86
ND
ND
<DL
64.9
2.6
28.9
16.2
1.03
2.15
<DL
<DL
0.93
3.71
ND
ND
129

00 49, 771
102 03, 566

S 307

49.9
76.8
34.4
28.8
120
18.7
ND
28.1
7.86
7.86
6.84
1.60
1.47
ND
2.65
ND
385

01 07, 608
102 08, 508

S 312

31.5
140.1
ND
13.3
138
3.99
156
33.2
5.82
11.8
6.65
3.94
7.21
39.2
24.5
3.55
619

01 11, 576
102 09, 866

Estuary
S 314

57.6
76.7
ND
9.37
55.7
8.76
80.3
17.4
4.39
5.05
2.89
0.72
0.73
ND
ND
ND
320

01 14, 506
102 10, 245

S 316

S 324

25.2
ND
ND
<DL
31.0
1.73
54.9
7.18
0.64
0.47
<DL
0.26
<DL
ND
<DL
ND
121

01 37, 549
101 53, 988

Table 1. Concentration of PAHs (ng L-1) in the water from the Siak River, the estuary and the coastal areas of Riau Province

Appendix 5. Tables for the Manuscript in the Chapter V

18.7
ND
ND
3.68
42.9
6.97
39.8
11.5
1.80
1.67
1.59
0.32
0.22
ND
<DL
0.53
130

01 53, 576
102 00, 410

Coast
S 317

S 318

29.8
ND
ND
2.38
25.2
3.57
49.4
11.2
1.76
1.73
2.60
0.19
0.13
0.91
1.08
0.37
130

01 53, 906
102 00, 560

134

1127
1293
78.2
221
8060
79.2
8556
11076
1851
2409
2353
521
246
1309
528
35.6
39743

17.5
694

10477
14770
96.7
284
6350
85.9
654
15017
2138
2826
2309
395
182
3085
348
32.1
59050

12.1
715

00 42,698

23.0
362

1718
5136
ND
69.1
1531
31.7
1904
2915
737
1077
370
17.5
29.0
125
43.6
ND
15704

101 40,105

00 32,516

101 26,118

00 36' 40,7''

101 19' 05,5''

28.8
529

835
724
ND
47.3
1335
22.4
8845
3897
768
1062
137
84.3
48.3
512
45.8
10.5
18373

101 18 34,5

00 35 46,5

Sep-04
Upstream
S102

ND = not detected; <DL = below detection limit; PKU = Pekanbaru; PRW = Perawang

Compound
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene
PAHs
SPM (mg/L)
PAHs (ng/L)

Station
Latitude
Longitude

PRW
S9

PKU
S10

Mar-04
Upstream
S17

Table 2.a. PAHs content (ng g-1) in the SPM from Siak River waters.

40.5
101

90.7
1018
ND
56.0
535
7.65
317
257
17.0
35.1
8.58
6.37
1.76
111
39.2
4.68
2504

101 22 59,7

00 33 32,9

PKU
S105

79.2
185

119
1100
ND
25.2
683
3.95
149
50.8
49.11
5.69
7.54
<DL
1.46
121
13.4
1.46
2330

101 35,319

00 35,846

PRW
S115

152
647

313
269
ND
12.9
199
5.51
1998
1053
69.6
140
80.9
6.12
5.16
91.0
6.78
11.5
4262

101 18, 990

00 36, 796

Jul-05
Upstream
S205

38.9
81

29.7
125
ND
23.2
333
5.00
252
501
30.1
50.5
49.1
18.0
19.0
356
278
16.5
2086

101 28, 121

00 32, 753

S272

S218

42.3
62

15.2
66.9
ND
31.5
369
5.61
261
217
26.8
43.1
16.1
6.89
21.3
271
87.9
36.0
1475

101 28, 161

00 32, 770

PKU

59.8
144

2413

157
282
ND
44.2
584
11.6
471
319
21.3
37.3
59.5
6.95
21.2
257
124
17.1

101 46, 364

00 46, 675

PRW
S216

28.3
182

2146
ND
124
65.8
1012
9.50
732
927
122
117
167
10.3
12.2
545
460
ND
6449

101 23,712

00 33,531

Mar-06
PKU
S291

36.8
195

788
ND
ND
30.0
527
12.6
1498
1715
186
352
75.2
6.24
21.7
ND
74.1
ND
5286

101 47, 685

00 45, 760

PRW
S301

135

10.1
33.6
ND
5.95
40.8
1.32
40.1
38.6
2.55
9.12
12.7
11.7
5.24
254
13.0
3.18
482
262
127

129
233

3.0

1.3

169
ND
ND
22.8
213
5.71
214
841
7.73
76.3
18.5
10.2
9.72
199
17.5
3.72
1808

01 07,724
102 07,793

01 04,616
102 07,958

280
128

25.3
ND
ND
5.79
47.3
1.84
11.3
35.2
1.42
11.4
15.5
11.6
5.19
265
16.9
2.72
457

5.0

01 07,845
102 09,150

Low (0 10 psu)
S 140
S 137

= estimated value; ND = not detected; <DL = below detection limit

Compound
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene
PAHs
SPM (mg/L)
PAHs (ng/L)

Latitude
Longitude
Salinity (psu)

S 139

38.2
293

1203
2532
ND
31.3
814
18.3
813
1394
300
412
74.8
14.0
11.0
52.3
ND
ND
7669

10.9

69.2
25.4

71.7
ND
<DL
5.78
77.0
1.70
71.4
50.1
2.50
8.57
47.3
0.45
5.97
16.5
6.49
1.95
367

20.8

01 15,591
102 10,212

625
97.3

42.5
ND
2.63
3.59
26.6
0.73
22.1
20.4
4.20
5.69
17.4
0.79
5.06
1.82
2.21
ND
156

25.3

01 15,578
102 10,256

Medium (10 25 psu)


S 134
S 130
S 126
01 07,915
102 09,633

Table 2.b. PAHs content (ng g-1) in the SPM from estuarine waters

74.2
95.2

271
463
ND
18.2
164
3.27
231
61.3
9.04
1.82
11.7
0.79
11.7
30.1
4.49
1.63
1282

26.2

01 14,474
102 10,166

S 132

135
69.7

51.3
179

218
2760
ND
18.9
202
3.65
152
59.2
5.18
8.65
11.7
<DL
20.1
17.4
6.40
ND
3483

26.3

26a
15.3
124
ND
6.08
75.9
2.03
58.0
76.6
7.20
7.01
31.7
2.93
7.89
84.9
6.42
10.0
516

01 13,354
102 10,072

48.3
51.2

337
ND
ND
ND
144
2.74
141
198
23.6
22.4
32.6
17.7
3.73
137
ND
ND
1059

27.4

01 14,033
102 11,981

High ( > 25 psu)


S 133
S 125

01 15,539
102 10,022

S 124

141
107

75.1
28.9
<DL
15.3
179
3.12
156
107
12.5
14.4
29.6
3.71
5.48
86.2
31.3
10.0
758

27.6

01 13,999
102 11,992

S 128

136

318
462
<DL
<DL
168
5.32
255
318
48.5
45.0
83.1
4.68
21.0
167
185
ND

2082

12.5
26.1

1687

16.6
28.0

101 34, 150

101 27, 859

658
ND
ND
15.8
250
2.28
122
343
40.5
46.7
76.5
5.19
14.1
109
4.35
ND

S 226

01 39, 970

S 225

01 41, 745

S 227

22.9
33.1

1449

342
337
<DL
ND
70.9
3.55
138
321
24.5
28.5
82.3
0.96
11.2
83.4
6.33
ND

101 43' 410''

01 41' 020''

S 228

13.4
15.8

1182

241
87.5
22.5
7.08
98.6
4.42
206
209
50.2
38.5
23.3
4.52
5.92
89.4
94.1
ND

101 51' 110''

01 38, 070

= stations adjacent to the Malacca Strait; ND = not detected; <DL = below detection limit

Compound
Naphthalene
Acenaphthylene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene
PAHs
SPM (mg/L)
PAHs (ng/L)

Station
Latitude
Longitude

9.25
11.7

1262

271
283
ND
<DL
78.9
2.45
166
185
23.7
21.3
59.2
3.62
8.04
70.0
81.4
7.25

102 00, 395

01 53, 497

S 229a

14.4
147

10234

2611
ND
137
ND
855
39.2
1870
2054
401
207
842
36.0
239
764
179
ND

101 53, 990

01 37, 380

S 267

12.5
82.3

6559

804
3105
86.7
ND
342
7.54
335
897
76.1
79.4
183.3
4.25
32.7
578
25.9
ND

101 59, 930

01 31, 240

S 266

15.6
35.4

2263

638
ND
25.7
15.1
943
1.89
106
263
20.4
38.9
40.3
2.78
9.06
144
14.4
ND

102 06, 140

01 26, 590

S 233

58.3
84.9

1457

215
530
ND
6.74
98.1
2.23
134
163
19.5
17.0
77.3
4.19
11.9
77.6
94.9
5.09

102 09, 118

01 16, 350

S 251

58.5
31.5

539

105
65.0
ND
4.20
49.7
1.40
77.0
77.4
17.6
13.3
25.1
5.55
2.88
51.8
43.0
ND

102 11, 405

01 20, 675

S 230

22.7
48.7

2146

279
ND
<DL
13.0
172
7.50
279
724
24.9
86.0
156
3.53
11.3
359
31.2
ND

102 13, 930

01 12, 650

S 231

125.5
40.8

326

38.7
ND
ND
1.61
26.1
0.87
37.2
45.0
6.46
8.17
19.7
3.79
2.84
116
17.2
1.90

102 13, 748

01 03, 835

S 232

Table 2.c. PAHs content (ng g-1) in SPM from coastal waters. The stations are labelled from the north part of the Riau coastline starting at the industrial city of
Dumai and continuing down to the south path of Selat Panjang

137

138
Table 3. Comparison of dissolved and particulate PAHs (ng L-1; ng g-1 d.w. respectively) from
various rivers, estuaries and coasts in several Asian and European countries.
Locations
Indonesia, Riau Province
Siak river
Siak estuary
Riau coastal areas

Sampling
PAHs
Median*
year
Dissolved PAHs (ng L-1)

Content
range

Reference

129 5140
320 619
121 130

This study
This study
This study

Shi et al., 2005


Zhou & Maskaoui,
2003
Zhang et al., 2004
Chen et al., 2007
Guo et al., 2007

2006

16
16
16

16

119**

45.8 1272

Daya Bay

1999

16

10984***

4228 29325

Minjiang River estuary


Qiantang River
Daliao River

1999
2005
2005

16
15
16

72400***
283***
5648

9900 474000
70.3 1844
946 13145

1993

10

4 36

1992 - 1998

15

5.3**

< 20

Fernandes et al.,
1997
Witt, 2002

1993 - 1995

15

139

n.d. 10724

Law et al., 1997

2002

16

270***

60 2090

4 days after

n.a

13700 497000

32 days after

n.a

~ 300

China
Rivers in Tianjin

824
385
130

Europe
Seine River, France
Baltic Sea
Coastal & Estuarine waters around
England & Wales
Oil Spill
The Prestige, NW & northern
Spanish coast,
The North Cape, Point Judith Pond,
NY, *

Forest Fire
The Llobregat river, Catalonia,
northern Spain
Mediterranean creek, Catalonia,
Spain

Gonzles et al.,
2006
Reddy & Quinn,
2001

1994

12

70

2 336

Olivella et al., 2006

2003

16

2.29

0.34 4.29

Vila-Escal et al.,
2007

Particulate PAHs (ng g-1 d.w.)


Indonesia, Riau Province
Siak river
Siak estuary
Riau coastal areas
China
Rivers in Tianjin
Pearl River and Estuary
Daliao River

2004-2006
16
16
16

5286
758
1572

1475 59050
156 7669
326 10234

This study
This study
This study

16
15
16

1325**
536
1759

938 64200
298 1337
305 237050

Shi et al., 2005


Luo et al., 2006
Guo et al., 2007

1993

10 + 1
alkyl

5000

1000 14000

Fernandes et al.,
1997

1994 - 1995

18 + 4
alkyl

4150

2540 8980

11850

6300 16570

2002
2005

Europe
Seine River, France
Elbe River, Germany
At Hamburg
At Dessau

*) own calculation from the literature data


**) median of the mean values of the sum PAHs
***) mean value
n.a = unavailable data from the given literature for median calculation

Heemken et al.,
2000
Heemken et al.,
2000

Appendix 6. Tables for the Manuscript in the Chapter VI


Table 1. Sampling stations and properties of the sediment fractions
Coarse

Station

Fine

Organic
Carbon
(%)

Total
Nitrogen
(%)

S 143
S 138
S 134
S 250
S 253
S 230
S 231
S 232

33.0
12.3
16.3
20.8
24.1
3.75
7.25
4.58

4.04
9.07
4.78
7.29
0.28
0.05
1.13
6.72

0.36
0.64
0.45
0.55
0.01
0.004
0.04
0.11

WR-B/06
K/06
H/06
BB4/07
WW8/07
WW10/07
CC021/07

41.2
60.4
68.8
19.3
32.6
40.6

0.52
0.46
1.24
0.13
0.39
0.37

0.04
0.02
0.02
0.01
0.01
0.02

17.2
22.9
81.3
10.1
33.8
18.4

85.7

3.87

0.05

88.2

Percentage*
of total
sediment

C/N
(mol/mol)

Organic
Carbon
(%)

Total
Nitrogen
(%)

C/N
(mol/mol)

60.4
70.3
72.3
55.4
69.5
65.5
69.2
83.0

2.25
1.46
1.61
3.31
1.44
1.29
2.18
1.71

0.13
0.19
0.11
0.35
0.09
0.09
0.14
0.11

20.4
8.85
17.6
10.9
18.7
16.2
18.8
18.6

58.8
39.6
31.2
80.7
67.4
59.4

1.21
2.84
1.91
1.98
2.14
1.80

0.12
0.24
0.19
0.18
0.15
0.16

12.1
14.0
11.6
12.6
16.7
13.5

14.3

4.11

0.09

52.3

Percentage*
of total
sediment

Siak Sumatra
12.9
16.5
12.4
15.5
32.9
17.4
30.6
47.9

Wenchang and Wanquan Hainan

*) The sum of the percentage of total sediment was not 100% due to the existence of particles bigger than 2 mm.

139

Table 2. Median and range of PAH contents (ng g-1 d.w.) in the sediment fractions from the Siak
estuary and coast (Sumatra), and the Wenchang and Wanquan estuary (Hainan, China). The PAHs are
ordered by following their elution time.

Siak

Compound
Median

Range

Wenchang and
Wanquan
Median

Range

Coarse

Naphthalene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene
PAHs

47.1
15.0
4.99
12.4
2.57
33.6
15.4
9.13
14.6
18.9
7.76
17.0
522
10.5
9.30
689

36.0 - 221
5.50 40.0
1.48 26.6
3.51 78.0
0.46 24.5
8.99 119
6.97 114
3.23 42.6
2.51 49.8
7.67 - 137
0.73 76.7
4.85 - 249
11.2 - 1061
0.63 43.3
0.66 71.8
125 1828

4.21
0.37
1.34
42.8
0.27
14.1
27.7
4.10
4.61
25.8
1.77
1.21
4.61
5.28
4.06
192

0.79 75.7
ND 0.42
ND 8.00
16.0 83.7
ND 2.11
4.02 84.3
2.95 87.1
1.05 7.57
2.62 32.1
9.17 63.0
0.01 7.90
0.17 11.9
1.15 97.3
0.72 19.7
1.64 30.2
112 386

34.5
8.70
2.55
9.28
1.26
17.4
11.7
4.34
6.32
10.8
3.38
4.89
38.7
9.41
4.04
202

23.3 61.1
2.23 32.8
1.08 4.82
1.71 34.3
0.30 3.27
3.73 86.0
2.19 38.0
1.24 7.68
3.73 12.8
4.20 30.1
0.05 8.28
3.54 28.6
10.96 - 206
0.70 17.2
1.08 13.8
88 426

12.8
2.37
10.6
64.0
1.12
92.3
85.2
13.2
13.7
24.2
6.90
2.62
1.03
2.53
20.5
473

1.73 15.8
0.21 4.39
0.34 43.0
26.6 193
0.49 2.28
6.15 119
0.54 194
2.56 42.1
0.90 48.5
6.33 155
1.03 18.9
0.68 14.5
ND 49.2
0.56 12.7
2.74 - 32.8
93.8 676

Fine
Naphthalene
Acenaphthene
Fluorene
Phenanthrene
Anthracene
Fluoranthene
Pyrene
Benzo(a)anthracene
Chrysene
Benzo(b)fluoranthene
Benzo(k)fluoranthene
Benzo(a)pyrene
Dibenzo(a,h)anthracene
Benzo(g,h,i)perylene
Indeno(1,2,3-c,d)pyrene
PAHs

ND = not detected

140

Das könnte Ihnen auch gefallen