Sei sulla pagina 1di 30

Marine Pollution Bulletin 47 (2003) 423–452

www.elsevier.com/locate/marpolbul

Development of oil hydrocarbon fingerprinting and


identification techniques
1
Zhendi Wang *, Merv F. Fingas
Emergencies Science and Technology Division, ETC, Environment Canada, 3439 River Road, Ottawa, Ont., Canada K1A 0H3

Abstract

Oil, refined product, and pyrogenic hydrocarbons are the most frequently discovered contaminants in the environment. To ef-
fectively determine the fate of spilled oil in the environment and to successfully identify source(s) of spilled oil and petroleum
products is, therefore, extremely important in many oil-related environmental studies and liability cases. This article briefly reviews
the recent development of chemical analysis methodologies which are most frequently used in oil spill characterization and iden-
tification studies and environmental forensic investigations. The fingerprinting and data interpretation techniques discussed include
oil spill identification protocol, tiered analytical approach, generic features and chemical composition of oils, effects of weathering
on hydrocarbon fingerprinting, recognition of distribution patterns of petroleum hydrocarbons, oil type screening and differenti-
ation, analysis of ‘‘source-specific marker’’ compounds, determination of diagnostic ratios of specific oil constituents, stable isotopic
analysis, application of various statistical and numerical analysis tools, and application of other analytical techniques. The issue of
how biogenic and pyrogenic hydrocarbons are distinguished from petrogenic hydrocarbons is also addressed.
 2003 Published by Elsevier Ltd.

Keywords: Fingerprinting; Oil fingerprinting; Biomarkers; Diagnostic ratios; Oil analysis; Spill identification; Oil forensic analysis; Oil composition

1. Introduction gallons (10,000 tonnes) of heavy fuel oil spilled, and an


equal amount remained aboard the sunken stern. A
Oil spill occurs everyday worldwide. A historical re- deadly storm after the spill hurled the sticky, heavily
view and analysis of reported oil spills over 10,000 gal- emulsified oil from the Erika ashore, churning tar into
lons in the International Oil Spill Database (Etkin, sandy beaches, splattering cliffs, roads, and car parks.
1998) shows that since the early 1960s, nearly 300 mil- This incident became FranceÕs most damaging oil spill in
lion gallons of oil have spilled into US marine waters 20 years, causing an environment of devastation on the
which occurred in 826 incidents involving tankers, bar- French Coast.
ges, and other vessels, and about 200 million gallons of Oil spills cause extensive damage to marine life, ter-
oil onto US soil from land pipeline spills (on average, 99 restrial life, human health, and natural resources.
land pipeline spills per year). An estimated 500 million Therefore, to unambiguously characterize spilled oils
and over 200 million gallons of oil have spilled from and to link them to the known sources are extremely
tankers in Europe and Pacific Asia since 1965, respec- important for environmental damage assessment, un-
tively. derstanding the fate and behaviour and predicting the
The most recent example of a large-scale spill is the potential long-term impact of spilled oils on the envi-
‘‘Erika’’ spill. The Maltese tanker Erika broke into two ronment, selecting appropriate spill response and taking
during a fierce storm on December 12, 1999 about 110 effective clean-up measures. In addition, successful fo-
km south of Brest, France. An estimated 2.8 million rensic investigation and analysis of oil and refined
product hydrocarbons in contaminated sites and recep-
*
tors yield a wealth of chemical ‘‘fingerprinting’’ data.
Corresponding author. Tel.: +1-613-990-1597; fax: +1-613-991- These data, in combination with historic, geological,
9485.
E-mail addresses: zhendi.wang@ec.gc.ca (Z. Wang), merv.fingas@
environmental, and any other related information on the
ec.gc.ca (M.F. Fingas). contaminated site can, in many cases, help to settle legal
1
Tel.: +613-998-9622; fax: +613-991-9485. liability and to support litigation against the spillers.
0025-326X/$ - see front matter  2003 Published by Elsevier Ltd.
doi:10.1016/S0025-326X(03)00215-7
424 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

This paper briefly reviews the recent development of methods for detailed component analysis. In the non-
chemical analysis methodologies which are most fre- specific methods such as the US EPA methods 1614 and
quently used in oil spill characterization and identifica- 418.1, only groups or fractions of chemical hydrocarbons
tion studies and environmental forensic investigations. (for example, measurement of total petroleum hydrocar-
The fingerprinting and data interpretation techniques bons (TPH) and EPA priority PAHs) are determined.
discussed include oil spill identification protocol, a tiered In response to the oil spill identification need and
analytical approach, generic features and chemical specific site investigation needs, attention has focussed
composition of oils, effects of weathering on hydrocar- on the development of flexible, tiered analytical ap-
bon fingerprinting, recognition of distribution patterns proaches which facilitate the detailed compositional
of petroleum hydrocarbons, oil type screening and dif- analysis by GC–MS, GC-FID, and other analytical
ferentiation, analysis of ‘‘source-specific marker’’ com- techniques that determine a broad range of individual
pounds, determination of diagnostic ratios of specific oil petroleum hydrocarbons. A variety of diagnostic ratios,
constituents, stable isotopic analysis, application of especially ratios of PAH and biomarker compounds, for
various statistical and numerical analysis tools, and interpreting chemical data from oil spills have been
application of other techniques in the oil forensic in- proposed for oil source identification and monitoring of
vestigations. The issue of how biogenic and pyrogenic weathering and biological degradation processes.
hydrocarbons are distinguished from petrogenic hydro-
carbons is also addressed. 2.1. Selected US EPA methods and the limitations of
these methods for oil analysis

2. Development of oil hydrocarbon fingerprinting tech- Table 1 summarizes selected EPA methods, major
niques applications and limitations of these methods for oil
analysis (EPA Methods, 1983, 1986, 1997, 1999). These
Petroleum contains thousands of different organic com- EPA methods have been used as routine procedures for
pounds. Successful oil fingerprinting involves appropriate determination of volatile and semivolatile aromatic hy-
sampling, analytical approaches and data interpreta- drocarbons presented in spilled oil and petroleum
tion strategies. A wide variety of instrumental and non- product samples. However, these methods were origi-
instrumental techniques are currently used in the analysis nally designed for waste water and industrial waste. The
of oil hydrocarbons, which include gas chromatography fundamental shortcoming with these methods is that
(GC), gas chromatography–mass spectrometry (GC–MS), they can not provide information on detailed chemical
high-performance liquid chromatography (HPLC), size- components which comprise the complex spill oil sam-
exclusion HPLC, infrared spectroscopy (IR), supercritical ples. The data generated from these methods are gen-
fluid chromatography (SFC), thin layer chromatography erally insufficient to answer the fundamental questions
(TLC), ultraviolet (UV) and fluorescence spectroscopy, (such as type and source, weathering status of spilled oil,
isotope ratio mass spectrometry, and gravimetric methods potential spillers, and so on) raised in an oil spill in-
(Wang et al., 1999a). Of all these techniques, GC tech- vestigation. Of the more than 160 priority-pollutant
niques are the most widely used. Compared to the mole- compounds determined by these methods, only 20 are
cular measurements two decades ago, GC methods have petroleum-related. Further, only half of these 20 com-
now been enhanced by more sophisticated analytical pounds are found in significant quantities in oils and
techniques, such as capillary GC–MS, which is capable of petroleum products. Also, the PAH compounds in oils
analyzing the oil-specific biomarker compounds and are dominated almost exclusively by the C1 to C4 al-
polycyclic aromatic hydrocarbons. The accuracy and pre- kylated homologues of the parent PAH, in particular,
cision of analytical data has been improved and optimized naphthalene, phenanthrene, dibenzothiophene, fluorene
by a series of quality assurance/quality control measures, and chrysene, none of which are measured by the stan-
and the laboratory data handling capability has been dard EPA methods. Other important classes of petro-
greatly increased through advances in computer technol- leum hydrocarbons (e.g., aliphatics and biomarkers) are
ogy. Several reviews (Sauer and Boehm, 1991, 1995; Krahn not measured by these methods at all. Another example
et al., 1993; Whittaker et al., 1995; Lundanes and Grei- is the use of the EPA 418.1 method to determine TPH
brokk, 1994; Kaplan et al., 1996, 1997; Wang et al., 1999a; content. The EPA 418.1 method, based on measuring
Stout et al., 2002) have been published on the analytical the absorption of C–H bond in the 3200–2700 wave-
methodologies for characterization and identification of oil number range, was originally intended for use only with
hydrocarbons using various analytical techniques. liquid waste but had been one of the most widely used
Depending on chemical/physical information needs, the methods for the determination of TPH in soils before its
point of application and the level of analytical detail, the demise because of the use of a chloro-fluoro carbon
methods used for oil spill study can be, in general, divided extractant. For some site assessments, Method 418.1
into two categories: non-specific methods and specific was the sole criterion for verification of site cleanup.
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 425

Table 1
Major applications and limitations of standard EPA methods for oil analysis
EPA standard method Target compounds and application Limitation for oil work
EPA 418.1 TPH by IR spectroscopy ––Inherent in accuracy of the method (positive
or negative bioses)
––Subject to various interference
––Lack of effective reference standards
EPA 1664 n-Hexane extractable materials and silica ––Only measures total extractable materials in
gel treated n-hexane extractable material by aqueous matrices
extraction and gravimetry ––Heavy interference
––Low molecular weight hydrocarbons could be
lost during distillation

EPA 600 series (method standards for


waste water)
602 Purgeable aromatics, by GC/FID ––600 and 8000 series cannot provide detailed
composition information of spilled oil
610 16 polynuclear aromatics, by HPLC/GC
624 Purgeable volatiles, by GC/MS ––Only BTEX measured, do not measure over 100
important oil hydrocarbons
625 Semivolatiles and pesticides, by GC/MS ––Do not measure dominated alkylated PAH
homologues, aliphatics and biomarker

8000 series (method standards for solid


waste, SW 846)
8015B Non-halogenated volatiles by GC/FID
8020A Aromatic volatiles by GC
8100 Volatiles by capillary GC/MS
8260A 24 PAHs by GC/FID
8270 B Semi

TPH: Total petroleum hydrocarbons; PAH: polycyclic aromatic hydrocarbons.

However, there were some problems associated with this spill identification requires further elaboration of oil
method such as inherent inaccuracy in the method (i.e., target analytes to include identification of the individual
positive or negative biases caused by various factors) specific target compounds and isomeric groups. The
and the lack of effective reference standards when selection of appropriate target oil analytes is dependent
working with an unknown. mainly on the type of oil spilled, the particular envi-
In recent several years, many EPA and ASTM ronmental compartments being assessed, and on ex-
methods have been modified to improve specificity and pected needs for current and future data comparison. In
sensitivity for measuring spilled oil and petroleum general, the major petroleum-specific target analytes
products in soils and waters (EPA Methods, 1997; that may be needed to be chemically characterized for
ASTM Method, 1997a,b,c), such as the EPA Method oil source identification and environmental assessment
8270 has been modified to increase analytical sensitivity include the following:
and to expand the analyte list to include petroleum-
specific compounds such as the alkylated PAHs and bio- • Individual saturated hydrocarbons including n-alk-
marker compounds. The principal modification to EPA anes (C8 through C40 ) and selected isoprenoids pris-
Method 8270 is the use of the GC–MS selected-ion-mode tane and phytane (in some cases, another three
(SIM) analysis that offers increased sensitivity relative to highly abundant isoprenoid compounds farnesane,
the full-scan mode. The modified EPA Method 8270, trimethyl-C13 , and norpristane are also included);
combined with column-cleanup and rigorous QA mea- • Volatile hydrocarbons including BTEX (benzene, tol-
sures, has been used to identify and quantify low levels of uene, ethylbenzene, and three xylene isomers) and
hydrocarbons by many environmental laboratories. alkylated benzenes (C3 - to C5 -benzenes), volatile par-
affins and isoparaffins, and naphthenes (mainly cyclo-
2.2. Selection of source-specific target analytes pentane and cyclo-hexane compounds);
• EPA priority parent PAHs and, in particular, the
In addition of groups or fractions of oil hydrocarbons petroleum-specific alkylated (C1 –C4 ) homologues of
(such as contents of asphaltenes and resins, TPH, total selected PAHs (that is, alkylated naphthalene, phe-
saturates, total aromatics, and EPA priority PAHs) are nanthrene, dibenzothiophene, fluorene, and chrysene
determined (Jokuty et al., 1999; Wang et al., 1998c) oil series). These alkylated PAH homologues (Table 2)
426 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

are the backbone of chemical characterization and generates information of great importance in deter-
identification of oil spill assessments; mining source(s), weathered state and potential treat-
• Biomarker terpane and sterane compounds (Table 3). ability; and
Analysis of selected ion peaks produced by these • Measurements of total petroleum hydrocarbons, the
characteristic, environmentally persistent compounds unresolved complex mixtures (UCM), and stable car-

Table 2
Source-specific target PAHs and alkylated homologous PAHs for oil spill studies
Compound Code Ring numbers Target ions
Oil-characteristic alkylated PAHs
Naphthalenes
C0 -naphthalene C0 N 2 128
C1 -naphthalenes C1 N 2 142
C2 -naphthalenes C2 N 2 156
C3 -naphthalenes C3 N 2 170
C4 -naphthalenes C4 N 2 184
Phenanthrenes
C0 -phenanthrene C0 P 3 178
C1 -phenanthrenes C1 P 3 192
C2 -phenanthrenes C2 P 3 206
C3 -phenanthrenes C3 P 3 220
C4 -phenanthrenes C4 P 3 234
Dibenzothiophenes
C0 -dibenzothiophene C0 D 3 184
C1 -dibenzothiophenes C1 D 3 198
C2 -dibenzothiophenes C2 D 3 212
C3 -dibenzothiophenes C3 D 3 226
Fluorenes
C0 -fluorene C0 F 3 166
C1 -fluorenes C1 F 3 180
C2 -fluorenes C2 F 3 194
C3 -fluorenes C3 F 3 208
Chrysenes
C0 -chrysene C0 C 4 228
C1 -chrysenes C1 C 4 242
C2 -chrysenes C2 C 4 256
C3 -chrysenes C3 C 4 270

Other EPA priority PAH pollutants


Biphenyl Bph 2 154
Acenaphthylene Acl 3 152
Acenaphthene Ace 3 153
Anthracene An 3 178
Fluoranthene Fl 4 202
Pyrene Py 4 202
Benz[a]anthracene BaA 4 228
Benzo[b]fluoranthene BbF 5 252
Benzo[k]fluoranthene BkF 5 252
Benzo[e]pyrene BeP 5 252
Benzo[a]pyrene BaP 5 252
Perylene Pe 5 252
Indeno[1,2,3-cd]pyrene IP 6 276
Dibenz[a,h]anthracene DA 5 278
Benzo[ghi]perylene BP 6 276
Surrogates and internal standard
[2 H10 ]Acenaphthene 164
[2 H10 ]Phenanthrene 188
[2 H12 ]Benz[a]anthracene 240
[2 H12 ]Perylene 264
[2 H14 ]Terphenyl 244
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 427

Table 3
Source-specific target biomarker terpane and sterane compounds for oil spill studies
Peak Compound Empirical formula Target ions
Terpanes
1 C19 tricyclic terpane C19 H34 191
2 C20 tricyclic terpane C20 H36 191
3 C21 tricyclic terpane C21 H38 191
4 C22 tricyclic terpane C22 H40 191
5 C23 tricyclic terpane C23 H42 191
6 C24 tricyclicterpane C24 H44 191
7 C25 tricyclic terpane C25 H46 191
8 C24 tetracyclic terpane + C26 tricyclic terpanes C24 H42 + C26 H48 191
9 C28 tricyclic terpane C28 H52 191
10 C28 tricyclic terpane C28 H52 191
11 C29 tricyclic terpane C29 H54 191
12 C29 tricyclic terpane C29 H54 191
13 Ts 18a(H),21b(H)22,29,30-trisnorhopane C27 H46 191
14 17a(H),18a(H),21b(H)-25,28,30-trisnorhopane C27 H46 191
15 Tm: 17a(H),21b(H)22,29,30-trisnorhopane C27 H46 191
16 17a(H),18a(H),21b(H)-28,30-bisnorhopane C28 H48 191
17 17a(H),21b(H)-30-norhopane C29 H50 191
18 18a(H),21b(H)-30-norneohopane C29 H50 191
19 18a(H) and 18b(H)-oleanane C30 H52 191
20 17a(H),21b(H)-hopane C30 H52 191
21 17b(H),21a(H)-hopane C30 H52 191
22 22S-17a(H),21b(H)-30-homohopane C31 H54 191
23 22R-17a(H),21b(H)-30-homohopane C31 H54 191
24 Gammacerane C30 H52 191
25 17b(H),21b(H)-hopane (Internal standard) 191
26 22S-17a(H),21b(H)-30,31-bishomohopane C32 H56 191
27 22R-17a(H),21b(H)-30,31-bishomohopane C32 H56 191
28 22S-17a i(H),21b(H)-30,31,32-trishomohopane C33 H58 191
29 22R-17a(H),21b(H)-30,31,32-trishomohopane C33 H58 191
30 22S-17a(H),21b(H)-30,31,32,33-tetrakishomohopane C34 H60 191
31 22R-17a(H),21b(H)-30,31,32,33-tetrakishomohopane C34 H60 191
32 22S-17a(H),21b(H)-30,31,32,33,34-pentakishomohopane C35 H62 191
33 22R-17a(H),21b(H)30,31,32,33,34-pentakishomohopane C35 H62 191

Steranes
34 C20 5a(H),14a(H),17a(H)-sterane C20 H34 217 & 218
35 C21 5a(H),14b(H),17b(H)-sterane C21 H36 217 & 218
36 C22 5a(H),14b(H),17b(H)-sterane C22 H38 217 & 218
37 C27 20S-13b(H),17a(H)-diasterane C27 H48 217 & 218
38 C27 20R-13b(H),17a(H)-diasterane C27 H48 217 & 218
39 C27 20S-13a(H),17b(H)-diasterane C27 H48 217 & 218
40 C27 20R-13a(H),17b(H)-diasterane C27 H48 217 & 218
41 C28 20S-13b(H),17a(H)-diasterane C28 H50 217 & 218
42 C29 20S-13b(H),17a(H)-diasterane C29 H52 217 & 218
43 C29 20R-13a(H),17b(H)-diasterane C29 H52 217 & 218
44 C27 20S-5a(H),14a(H),17(H)-cholestane C27 H48 217 & 218
45 C27 20R-5a(H),14b(H),17b(H)-cholestane C27 H48 217 & 218
46 C27 20S-5a(H),14b(H),17b(H)-cholestane C27 H48 217 & 218
47 C27 20R-5a(H),14a(H),17a(H)-cholestane C27 H48 217 & 218
48 C28 20S-5a(H),14a(H),17a(H)-ergostane C28 H50 217 & 218
49 C28 20R-5a(H),14b(H),17b(H)-ergostane C28 H50 217 & 218
50 C28 20S-5a(H),14b(H),17b(H)-ergostane C28 H50 217 & 218
51 C28 20R-5a(H),14a(H),17a(H)-ergostane C28 H50 217 & 218
52 C29 20S-5a(H),14a(H),17a(H)-stigmastane C29 H52 217 & 218
53 C29 20R-5a(H),14b(H),17b(H)-stigmastane C29 H52 217 & 218
54 C29 20S-5a(H),14b(H),17b(H)-stigmastane C29 H52 217 & 218
55 C29 20R-5a(H),14a(H),17a(H)-stigmastane C29 H52 217 & 218
Monoaromatic steranes 253
Triaromatic steranes 231
428 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

bon isotope ratio (d13 C) are also included in many effects on chemical composition changes, or source dif-
cases. ferentiation, the tiered analytical approaches may vary.
In the comprehensive study of the fates and effects of the
Another potentially valuable hydrocarbon group for Exxon Valdez oil spill in Prince William Sound (PWS)
oil spill identification is nitrogen and oxygen hetero- of Alaska, Boehm et al. (1998) and Page et al., 1995
cyclic hydrocarbons. These heterocyclic hydrocarbons have applied various tiered analytical approaches to
are generally only present in oils at quite relatively low obtain hydrocarbon fingerprints in thousands of sedi-
concentrations compared to PAHs. However, they be- ment and biological samples and to allocate complex
come enhanced with weathering because they are bio- hydrocarbons to multiple sources. Wang et al. have re-
refractory and persistent in the environment. Most ported application of a tiered analytical approach
organic nitrogen hydrocarbons in crude oils are present for identification of the source(s) of various unknown
as alkylated aromatic heterocycles with a predominance spilled oil samples (Wang et al., 1997b, 1998b, 1999c,
of neutral pyrrolic structures over basic pyridine forms. 2002). The tiered approach they used includes the fol-
They are chiefly associated with high boiling fractions, lowing:
much of the nitrogen in petroleum being in asphaltenes.
Individual and alkyl homologues of carbazole, quino- • Tier 1, determination of hydrocarbon groups in oil
line, and pyridine have been identified in many crude residues;
oils. These compounds may provide important clues for • Tier 2, product screening and determination of n-alk-
potential sources of hydrocarbons in the environment anes and TPH;
and for tracing petroleum molecules back to their • Tier 3, distribution pattern recognition of target
biological precursors. Compared to the PAHs and PAHs and biomarker components (sometimes the
biomarkers, the application of nitrogen and oxygen- volatile hydrocarbons are monitored);
containing heterocyclic hydrocarbons in source iden- • Tier 4, determination and comparison of diagnostic
tification is still in its infancy, and more research is ratios of the ‘‘source-specific marker’’ compounds
needed. with the potential source oil and with the correspond-
ing data from database; and
• Tier 5, determination of weathered percentages of the
2.3. Oil spill identification protocol
residual oil.
The oil spill identification system currently used is
Stout et al. (2002) recently proposed a similar tiered
largely based on GC-FID and GC–MS techniques. Data
analytical approach to hydrocarbon forensic analysis,
produced from these two methods are used to compare
which is summarized in Fig. 2. This tiered approach
spill samples with samples taken from suspected sources
gives environmental forensic investigator the flexibility
(Wang et al., 1999a). Very recently, SINTEF Applied
to gather information necessary to address site- or in-
Chemistry of Norway and Battelle of the USA pub-
cident-specific questions about the nature and extent,
lished the ‘‘Improved and standardized methodology for
and ultimately source(s), of hydrocarbon contamina-
oil spill fingerprinting’’ (Daling et al., 2002), which in-
tion.
cludes four tiered ‘‘levels’’ of analyses and data treat-
ment. The recommended methodology approach is a
2.5. Quality assurance
result of documented and analytical improvements and
a more quantitative treatment of analytical data from
In order to support spilled oil forensic investigations,
GC-FID and GC–MS and the operational experiences
well-defined quality management (including quality as-
over past few years among the participating forensic
surance and quality control procedures) must be a fun-
laboratories. Fig. 1 presents the modified ‘‘Protocol/
damental element of any analytical lab program. The
decision chart for the oil spill identification methodol-
chemical measurements must be conducted within the
ogy’’. The final assessment are concluded by the four
framework of highly stringent, defensible and reliable
operational and technical defensible identification terms:
QA and QC program.
positive match, probable match, inconclusive or non-
Besides the routine quality control measures required
match.
by standard EPA and ASTM methods such as instru-
ment calibration, surrogate spiking, procedural blanks,
2.4. Tiered analytical approach matrix spike recoveries, replicate analyses, some refine-
ments have been further implemented by many oil spill
Tiered analytical approaches have been increasingly analysis laboratories in order to achieve improved ana-
applied for oil spill identification in recent years. De- lytical precision and accuracy (Page et al., 1995; Short
pending on the needs of spilled oil characterization, et al., 1996; Wang et al., 1994a; Stout et al., 2002). The
support for biological studies, monitoring weathering key refinements include the following:
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 429

Fig. 1. Protocol/decision chart for the oil spill identification (from Daling et al., 2002).

• To establish 5-point calibration curves that demon- in Appendix B, 40CFR (Code of Federal Regula-
strate the linear range of the analysis. tions) Part 136.
• To analyze quality control standards prepared from • To increase sample size and to reduce the sample ex-
the National Institute of Standards and Technology tract pre-injection volume for those sediment samples
(NIST) certified standard reference materials (SRMs) with very low concentrations of hydrocarbons.
with the sample batches for accuracy assessment.
• To apply more rigorous calibration check standards
of ±15%. 3. Oil and petroleum product type screening and differ-
• To determine relative response factors (RRFs) of tar- entiation
get analytes of interest from authentic standards.
• To manually set the baselines for alkylated PAHs at Crude oil is a complex mixture of thousands of dif-
various alkylation levels. ferent organic compounds formed from a variety of
• To estimate the method detection limits (MDLs) for organic materials that are chemically converted under
each target analytes using the procedure described differing geological conditions over long periods of time.
430 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

Fig. 2. Tiered analytical approach to hydrocarbon fingerprinting (from Stout et al., 2002).

Crude oils contain primarily carbon and hydrogen Generally, the type and identity of fresh to mildly
(which form a wide range of hydrocarbons from light weathered oils and petroleum products can be readily
gases to heavy residues), but also contains smaller revealed from their GC-FID traces, especially where the
amounts of sulfur, oxygen and nitrogen as well as metals spilled oil or petroleum product is heavy and back-
such as nickel, vanadium and iron. The infinitely vari- ground hydrocarbon levels are low in an impacted en-
able nature of these factors results in distinct chemical vironment. In addition to measuring TPH in samples,
differences between oils. Refined petroleum products are GC-FID chromatograms provide a distribution pattern
fractions usually derived by distillation of crude oil. of petroleum hydrocarbons (e.g., carbon range and
Because of dissimilarities in characteristics of crude oil profile of UCM), fingerprints of the major oil compo-
feed stocks and variations in refinery processes, refined nents (e.g., individual resolved n-alkanes and major
oil products differ in their chemical compositions. Thus, isoprenoids), and information on the weathering extent
all crude oils and petroleum products, to some extent, of the spilled oil. Comparing biodegradation indicators
have chemical compositions that differ from each other. (such as n-C17 /pristane and n-C18 /phytane) for the spil-
This variability in chemical compositions results in un- led oil with the source oil can be also used to monitor
ique chemical ‘‘fingerprints’’ for each oil and provides a the effect of microbial degradation on the loss of hy-
basis for identifying the source(s) of the spilled oil. drocarbons at the spill site for short time periods.
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 431

3.1. General chemical composition features of crude oils reforming, isomerization, alkylation, and blending pro-
cesses:
Crude oil compositions vary widely. Depending on
the sources of carbon from which the oils are generated (1) Light distillates are typically products in the C3 –C12
and the geologic environment in which they migrated carbon range. They include aviation gas, naphtha,
and from which reservoir, they can have dramatically and automotive gasoline. The GC trace of fresh light
varied compositions in the C5 –C40 carbon range such as distillates are featured with dominance of light-end,
relative amounts of paraffinic, aromatic and asphaltenic resolved hydrocarbons and a minimal UCM. The
compounds, large differences in the n-alkane distribu- most common light distillate, gasoline, has consider-
tions and UCM, and significantly different relative ratios able BTEX and alkylated benzene compounds.
of isoprenoids to normal alkanes. Fig. 3 shows GC-FID Other compound classes found in abundance in gas-
chromatograms for six different oils. Clearly, these six oline includes branched-chain alkanes and olefin
oils are very different, as not only are there large dif- hydrocarbons with double bonds.
ferences in the n-alkane distributions and UCMs, but (2) Mid-range distillates are typically products in a rel-
also in relative ratios of isoprenoids to normal alkanes. ative broad carbon range (C6 –C26 ) and include kero-
Note that the Orimulsion sample, has nearly no n-alk- sene, jet fuel, and diesel products. Jet fuels are
anes in its GC-FID chromatogram. similar in gross composition, with many of the dif-
ferences in them attributable to additives designed
3.2. General chemical composition features of refined to control some fuel parameters such as freeze and
products pour point characteristics. For example (Fig. 4),
Jet B fuel from Alaska is dominated by GC-resolved
Refined petroleum products are obtained from crude n-alkanes in a narrow range of n-C8 to n-C18 with
oil through a variety of distillation, cracking, catalytic maximum being around n-C12 . The UCM is well

Fig. 3. GC-FID chromatograms for six oils. These six oils are different, as not only are there large differences in the n-alkane distributions and UCMs,
but also in relative ratios of isoprenoids to normal alkanes. Note that the Orimulsion sample, has nearly no n-alkanes on its GC-FID chromatogram.
432 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

Fig. 4. GC-MS (m=z 85) chromatograms of four petroleum products, illustrating distinguishing features of n-alkane distribution patterns between
these oil products.

defined. As for Diesel No. 2, characteristic wider-re- (4) Lubricating oils are specialty products that have a
solved hydrocarbons (in the carbon-range of C6 –C23 broad profile in the C18 –C40 with nearly no resolved
with maxima being around C12 –C14 ) and a central alkanes being present. Common type of lubricating
UCM ‘‘hump’’ are obvious. The properties of a oils include crackcase oil, transmission fluid, hy-
given diesel are largely a function of the crude oil draulic fluid, and cutting oil. In some hydraulic flu-
feedstock. ids, for example, the PAH concentrations can be
(3) Classic heavy fuel types include fuel Nos. 5 and 6 very low, while the biomarker concentrations could
(also known as Bunker C) or residual. The heavy be very high, in comparison with most other refined
residual fuels are largely used in marine diesel and products. These features in chemical compositions
industrial power generation. The chemical compo- can be used to distinguish refined products.
sition of Bunker C can vary widely and remarkably,
depending on production oilfields, production 3.3. Source identification by comprehensive two-dimen-
years, and processes it has undergone. Currently, sional GC
many Bunker type fuels are produced by blending
residual oils with diesel fuels or other lighter Gains et al. (1999) recently described application of
fuels in various ratios for marine or power plant two dimensional GC–GC with FID detection for spill
use. A case study in which biomarkers and PAHs source identification. In this method, each analyte in oil
were used to correlate a spilled Bunker C to its sus- is subject to two different separations achieved using two
pected sources was recently reported (Stout et al., GC columns connected serially by a thermal modulator.
2001). Compounds of similar chemical structure were grouped
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 433

together in ordered two-dimensional chromatograms. alkanes > low-molecular-weight aromatics > high-mole-
The GC–GC analysis resulted in a match between the cular-weight aromatics and cyclic alkanes.
spill samples and one of the source samples. This result Photooxidation. Photooxidation is considered to be
was consistent with the results obtained by GC–MS. another most important factor involved in the trans-
formation of crude oil or its products released into the
marine environment (Garrett et al., 1998). The photo-
4. Effects of weathering on oil hydrocarbon fingerprinting chemical degradation yields a great variety of oxidized
compounds which are highly soluble in water.
When crude oil or petroleum products are acciden- Aggregation. Following the Exxon Valdez oil spill
tally released to the environment, they are immediately in Prince William Sound, Alaska in March 1989, the
subject to a wide variety of weathering process (Jordan process of oil–mineral aggregates (OMA, also called
and Payne, 1980). These weathering processes can in- clay–oil flocs or clay–oil aggregates in the past) was dis-
clude: (1) evaporation, (2) dissolution, (3) microbial covered as a mechanism affecting the rate of natural
degradation, (4) other processes such as dispersion and cleansing of oil residues from shorelines (Bragg and
water–oil emulsification, photooxidation, adsorption Owens, 1994). Oil-mineral aggregates was found to re-
onto suspended particulate materials, and oil–mineral sult from interactions among the oil residues, fine min-
aggregation. eral particle, and seawater. Since then, a number of lab
Evaporation. In the short term after a spill (hours to studies, field trials, and spill site investigations have
days), evaporation is the single most important and been conducted. OMA formation has now been identi-
dominant weathering process, in particular for the light fied as an important process that facilitates the natural
petroleum products. In the first few days following a removal of oil stranded in coastal sediments. OMA
spill, the loss can be up to 70% and 40% of the volume of formation is enhanced by physical processes such as
light crudes and petroleum products. For heavy or re- wave, energy, tides or currents. It has recently been
sidual oils the losses are only about 5–10% of volume noted that oil biodegradation may be enhanced by
(Fingas, 1995). OMA formation.
Dissolution. The amount of the oil hydrocarbons
dissolving in water phase from oil slick largely depend 4.1. Chemical composition changes due to weathering
on the molecular structure and polarity of a given oil
component, and the relative solubility of the oil com- Weathering causes considerable changes in the chem-
ponent in water phase versus its solubility in the oil ical and physical properties of spilled oils. The degree
phase. In general, (1) the aromatic hydrocarbons are (lightly, moderately, or highly weathered) and rate of
more soluble than aliphatic hydrocarbons, (2) solubili- weathering is different for each spill and is controlled by
ties increase as the alkylation degrees of alkylated ben- a number of spill conditions and natural processes such
zene or PAHs decrease, (3) the lower molecular weight as type of the spilled oil, spill site and environmental
hydrocarbons are more soluble than the high molecular conditions, and microbiological activities. Major chem-
weight hydrocarbons in that class. Therefore, it can be ical compositional changes due to weathering can be
readily understood why the BTEX and lighter alkyl- summarized as the following (Page et al., 1995; Wang
benzene compounds and some smaller PAH compounds et al., 1999a):
such as naphthalene are particularly susceptible to dis-
solution or ‘‘water-washing’’. • For lightly weathered oils, significant losses occur
Biodegradation. Biodegradation of hydrocarbons by in the low-molecular-weight n-alkanes, however the
natural populations of microorganisms represents one ratios of n-C17 /pristane and n-C18 /phytane have been
of the primary mechanisms by which petroleum and found to be virtually unaltered from those measured
other hydrocarbon pollutants are eliminated from the for the source oil. Therefore, for fresh or mildly
environment (Prince, 1993; Leahy and Colwell, 1990). weathered oil products, alkane and isoprenoid ana-
The biodegradation of petroleum and other hydrocar- lysis may be used for oil source identification. In
bons in the environment is a complex process, whose heavily weathered oils, the n-alkanes, and even the
quantitative and qualitative aspects depend on the type, isoprenoids in some cases, may be completely lost.
nature and amount of the oil or hydrocarbon present, Under such circumstances, GC-FID analysis is of lit-
the ambient and seasonal environmental conditions tle value for suspect source identification.
(such as temperature, oxygen, nutrients, water activity, • Rapid loss of volatile aromatic compounds including
salinity, and pH), and the composition of the auto- BTEX (benzene, toluene, ethylbenzene, and xylenes)
chthonous microbial community. Hydrocarbons differ and C3 -benzenes. When oils are weathered to a cer-
in their susceptibility to microbial attack. In general, the tain degree (approximately in the range of 20–25%
degradation of hydrocarbons is ranked in the following weathering), the BTEX and C3 -benzenes are com-
order of decreasing susceptibility: n-alkanes > branched pletely lost.
434 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

• Pronounced decrease in the naphthalenes relative to as an internal marker for evaluation of the biodegra-
other alkylated PAH series. dation and weathering of crude oil.
• Development of a profile in each alkylated PAH fam-
ily showing the distribution of C0 < C1 < C2 < C3 . 4.3. Kinetic model to evaluate PAH weathering
• Enhancement of the chrysenes relative to other PAH
series and significant decrease in the relative ratios Short and Heintz (Short and Heintz, 1997) have de-
of the sum of naphthalenes, phenanthrenes, dibenzo- veloped a first-order loss-rate (FOLR) kinetic model of
thiophenes, and fluorenes to chrysenes. PAH weathering to evaluate 7767 environmental samples
• Gradual concentration in the relative abundances of collected for the Exxon Valdez oil spill for the presence of
biomarker compounds (terpanes and steranes) be- spilled oil. The modelled PAHs included the 14 most
cause of their refractory nature and high resistance persistent compounds of 31 analyzed by GC-MS. Pa-
to biodegradation. Also, the relative ratios of paired rameters include loss-rate constants related to energy
terpane compounds including Ts=Tm, C23 /C24 , C29 / required for PAHs to escape from petroleum and a
C30 , C31 22S/(22S + 22R), C32 22S/(22S + 22R), C33 quantitative index of weathering. The model accounts for
22S/(22S + 22R) are almost not altered by weathering. 91% of the temporal variability of modelled PAH con-
centrations. Short (Short, 2002) has applied this model to
4.2. Degradation rate four independent case studies. The approach used in these
case studies evaluates a goodness-of-fit metric between
The early effect of microbial degradation is moni- the measured hydrocarbon composition of an environ-
tored by the ratios of biodegradable to less degradable mental sample and a suspected source after correcting for
compounds such as the n-C17 /pristane and n-C18 /phy- PAH weathering loss based on FOLR kinetics.
tane ratios. It has been demonstrated, however, that
changes in these ratios may substantially underesti- 4.4. Estimation of weathering age of refined products
mate the extent of biodegradation because isoprenoids
also biodegrade to a significant degree. Later, highly As discussed above, oil weathering is a very complex
degradation-resistant components such as C30 17a(H), process, and the chemical composition changes due to
21b(H)-hopane are selected to serve as conserved ‘‘inter- weathering, the weathering degree and weathering rate
nal standards’’ for determining rate and extent of are determined by many factors. Estimation of weath-
weathering for the spilled residual oil (Butler et al., 1991; ering age of a spill is, therefore, a very complex issue and
Prince et al., 1994). For distilled oil products, such as difficult task, and it should be dealt case by case in most
diesel and jet fuel samples, which may not contain sig- situations. For example, we found during the 25-year
nificant quantities of biomarker compounds and chry- old Nipisi spill study (Wang et al., 1998a) that the
senes, less ‘‘conservative’’ PAHs with a high degree of weathering degrees of the residual oil were dramatically
alkylation such as C4 or C3 -phenanthrenes can be se- different from sample to sample: some samples were
lected and used as alternative internal standards. highly weathered with all n-alkanes and isoprenoids
As an example, Fig. 5A and 5B show the GC chro- being completely lost, while some samples were only
matograms for TPH and n-alkane analysis and GC– very lightly weathered (indicated by existence of large
SIM–MS chromatograms for alkylated benzene and quantities of BTEX and alkylbenzene compounds, and
alkylated PAH distribution analysis for 25-year-old the n-alkane distribution being almost not changed in
Nipisi spilled oil samples (Wang et al., 1998a), illus- comparison with the reference oil).
trating the effect of field weathering conditions and Christensenand and Larsen (1993) assembled data
sample depths on chemical composition changes of the from 12 diesel spill sites with known spill time for each
spilled oil during the 25 year period. Fig. 5 clearly spill in northern Europe. The data from these sites ul-
demonstrates that 25 years after the spill, the remaining timately yielded a linear correlation between the ‘‘time
underground oil is still relatively ‘‘fresh’’ in comparison since the release in years’’ and ‘‘average n-C17 /pristane
to the reference oil and the surface residual oil. Sub- ratio’’. Kaplan et al. (1995) extracted the plotted data
surface oil degradation has been demonstrated to be points from Christensen and Larsen and published a
slow and will proceed at a very slow rate because the linear equation used for weathering age determination
peat in this wetland habitat system is acidic, and water of diesel. However, it should be noted that their cali-
saturated, i.e., largely anaerobic. Also, the average an- bration chart was developed for particular spill sites
nual temperature is only 1.7 C, based on 22 years of under a unique set of environmental conditions. Bacte-
weather records. rial activity could be greatly different from site to site,
Sasaki et al. (1998) found similar to 17a, 21b-hopane, resulting in very different rate of biodegradation. Other
vanadium, the most abundant heavy metal in crude oils, environmental factors such as temperature, water level,
was not significantly reduced even after an intensive salinity, and many others will also play key roles in
biodegradation of crude oil. Thus, they used vanadium determining the rate of biodegradation. Stout et al.
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 435

Fig. 5. GC-FID (Fig. 5A, left panel) and GC–SIM–MS (Fig. 5B, right panel) chromatograms of saturated and aromatic fractions for the Nipisi
reference source oil PL-B, and samples N2-1A (0–2 cm), N2-1B (12–16 cm), and N2-1C (30–40 cm), illustrating the effects of field weathering
conditions and sample depths on chemical composition changes of aliphatic (in particular n-alkanes) and alkylbenzenes and alkylated PAHs, re-
spectively. Sur. and IS represent surrogate and internal standard. B and N represent benzene and naphthalene respectively; n, 0, 1, 2, and 3 represent
carbon numbers of alkyl groups in alkylbenzenes and alkylated naphthalene homologues.

(2002) and Philp (2002a) have discussed in detail the suspected candidate sources, and background materials
numerous caveats that must considered in this age-dat- may sufficiently meet the needs of a forensic investiga-
ing method that otherwise might lead to oversimplica- tion. However, when the chemical similarity/difference
tion of the very complex issue of age dating diesel or between spilled oil and the suspected source(s) is not
other oil product contamination. obvious, or a large number of candidate sources are
involved, or spilled oil has undergone heavy weathering
5. Oil correlation and source identification by PAH and significant alteration in its chemical compositions,
fingerprinting analysis the qualitative approach would be difficult to defend,
and therefore the quantitative fingerprinting analysis of
In some instances, qualitative chemical analysis and degradation-resistant PAH and biomarker compounds
visual comparison of chromatograms of spilled oil, must be performed.
436 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

5.1. PAH distribution pattern recognition thus making PAHs one of the most valuable finger-
printing classes of hydrocarbons for oil identification.
Crude oils and refined products from different sources Even differences between the same types of products are
can have very different PAH distributions. Also, many discernible through examination of the PAH distribu-
PAH compounds are more resistant to weathering than tion. Examples of PAH distribution of some oil and
their saturated hydrocarbon counterparts (n-alkanes petroleum product types are illustrated in Fig. 6. The oil
and isoprenoids) and volatile alkylbenzene compounds, products differ significantly in the PAH distribution

Fig. 6. Alkylated homologous PAH and other EPA priority PAH distributions for the Federated crude, Jet B, Diesel, and Orimulsion, illustrating
differences in PAH distribution features between different oil and oil products. Note that for clarity, different scales are used for Y-axis. N, P, D, F,
and C represent naphthalene, phenanthrene, dibenzothiophene, fluorene, and chrysene, respectively; 0–4 represent carbon numbers of alkyl groups in
alkylated PAH homologues. See Table 2 for abbreviations for other EPA priority PAHs.
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 437

from the crude oils and from each other. Jet B fuel has A method using the double ratio plots of alkylated
extremely high content of the naphthalene series (99%) PAH homologues, C2 D/C2 P versus C3 D/C3 P, for iden-
with the other four alkylated PAH series being only 1% tification and differentiation of petroleum product sour-
in total. In addition, no 4- to 6-ring PAHs were detected ces has been developed and extensively used in the studies
of the other 15 EPA priority PAHs. Diesel No. 2 has a of the Gulf War oil spill (Sauer et al., 1993, 1998) and of
high naphthalene content (86%), a low phenanthrene the 1989 Exxon Valdez oil spill to distinguish Alaska
content (5%), and no chrysenes. In the Orimulsion 400, North Slope (ANS) crude, its weathering products, and
the unusually high contents of the alkyl phenanthrenes diesel refined from ANS feed stock from other petrogenic
and dibenzothiophenes were pronounced, accounts for hydrocarbons (Bence and Burns, 1995; Bence et al., 1996;
approximately 38% and 22%, respectively, of the total Page et al., 1995; Brown and Boehm, 1993). Further-
PAHs. In addition, a profile in each alkylated PAH more, Douglas et al. (1996) have defined the C3 D/C3 P
family showing the distribution of C0 < C1 < C2 < C3 is and C3 D/C3 C as ‘‘source ratios (the ratios that be almost
very apparent, similar to the severely weathered oil. constant because the compounds degraded at the same
rate)’’ and ‘‘weathering ratios (the ratios that change
5.2. Diagnostic ratios of PAH compounds substantially with weathering and biodegradation)’’, re-
spectively. They were applied to described oil depletion
A number of diagnostic ratios of target alkylated and to identify sources in subtidal sediment data from the
PAH species have been successfully used as indicators Exxon Valdez spill and a North Sea oil spill.
for oil spill identification. These are briefly summarized Hostettler et al. (1999) reported a method using the
in Table 4. A benefit of comparing diagnostic ratios of PAH refractory index, ratio of two of the most refractory
spilled oil and suspected source oils is that any con- constituents of most oils (triaromatic steranes and methyl-
centration effects are minimized. In addition, the use of chrysenes), as a source discriminant of hydrocarbon input
diagnostic ratios to correlate and differentiate oils tends for differentiation of three different oils (Exxon Valdez oil,
to induce a self-normalizing effects on the data since Katalla oil, and PWS sediment hydrocarbons).
variations due to instrument operating conditions, op- In the studies of characterization of spilled oil resi-
erators, or matrix effects are minimized. dues from the Arctic environments and northern inland,

Table 4
Diagnostic ratios of ‘‘source-specific’’ PAH compounds used for oil spill studies
Diagnostic ratios Application References
Double ratio plots (C2D/C2P vs. ––Distinguishing between sources with similar chemical compositions Bence et al. (1996);
C3D/C3P) ––Useful in establishing statistical models for source allocation Boehm et al. (1997);
––Examples: Exxon Valdez oil spill study; Gulf War oil spill study Page et al. (1995, 1996)
Double ratio plots (C3D/C3P vs. ––Distinguishing among weathered crude oils Douglas et al. (1996);
C3D/C3C) ––Distinguishing spilled oil from other sources Sauer and Boehm (1995)
––Examples; to describe oil depletion and to identify in subtidal data
from the M/C Haven spill in Italy, Exxon Valdez spill, and a North
Sea oil spill
P
4- to 6-ring non-alkylated PAH/ ––Used for multi-source hydrocarbons identification in the study Kennicutt (1998);
PPAH P hydrocarbon contamination on the Antartic Peninsula Kvenvolden et al. (1995);
Naphs/ PAH Sauer and Boehm (1995)
P P
P Phens/P dibens
Phen/ phens

Ratios of 3 m-DBT isomers ––Used for source identification of unknown spilled oils Fayad and Overton (1995);
––Distinguishing between oils with a similar chemical compositions Wang and Fingas (1995b, 1999a)
––Differentiation between oils due to physical weathering and biode-
gradation
––Marker of biodegradation
C0 C; C1 C: C2 C: C3 C ––Used for source identification in the Arrow and BIOS spill studies Boehm et al. (1997);
P P
P chrys/P phens ––Weathering indicator Sauer et al. (1998);
chrys/ dibens ––Differentiation between compositioin changes due to physical Wang et al. (1994b, 1999a)
Relative distribution of PAH in weathering and biodegradation
each homologous family
P
PAH: the sum of total PAH including five target alkylated PAH homologue and the other EPA priority PAH; m-DBT: methyl-dibenzothiophene;
4- to 6-ring non-alkylated PAH include fluoranthene, pyrene, benzofluoranthenes, benzopyrenes, indenopyrene dibenzoanthracene and benzo-
peryene.
438 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

Wang et al. (1994b, 1998a,c), utilized a number of di-


agnostic ratios of selected source-specific alkylated
PAHs in combination with determination of ratios of
selected paired biomarkers for source identification and
differentiation, determination of weathering extent and
degree of surface and subsurface samples, and distin-
guishing between composition changes due to physical
weathering and biodegradation.

5.3. PAH isomer analysis

The use of the sum of the alkylated PAHs as multi-


component analytes in deriving diagnostic ratios for oil
characterization and spill assessment have been made
considerable advances as described above. Recently, Fig. 7. Plot of the relative ratios of 2-/3-methyldibenzothiophene to 4-
research has been further expanded to use individual methyldibenzothiophene versus the relative ratios of 1-methyl-
source-specific isomers within the same alkylation level dibenzothiophene to 4-methyldibenzothiophene for 49 different oils
and to determine the relative isomer-to-isomer distribu- and oil products. The circles 1 and 2 indicate related samples from
origins of North Slope and Terra Nova, respectively.
tion for source identification of spilled oil. For example,
chromatographically well-separated C1 -dibenzothio-
phene isomers (Fayad and Overton, 1995; Wang and shows how scattered the data points representing the
Fingas, 1995b) present in all oils at relatively high con- various oils are. Another pronounced feature observed
centrations and their relative abundance distributions from Fig. 7 is that related oils produce tight clusters on
vary significantly from different sources. As the alkyla- the plot. The use of these ratios complements existing
tion levels increase, more isomers will be detected (for methods of oil characterization, but has its own distinct
example, the C3 -dibenzothiophenes as a group, contain advantages. The ratios of the C1 –DBT compounds, in
more than 20 individual isomers with different relative combination with other fingerprinting data, have been
abundances). The differences between the isomer dis- used to successfully discriminate different oil samples
tributions reflect the differences of the depositional (Wang et al., 1999c) and to identify the source of oil on
environment during oil formation. Compared to PAH contaminated birds (Wang et al., 1997b).
homologous groups at different alkylation levels, higher During January and February 1996, a significant
analytical accuracy and precision may be achieved due number of tarball/patty incidents occurred along the
to the close match of physical/chemical properties of the coast of Vancouver Island (British Columbia), Wash-
isomers. Also, the relative distribution of isomers are ington, Oregon, and California (Wang et al., 1998b). In
subject to little interference from weathering in short- addition of determination of conventional diagnostic
term or lightly weathered oils. Hence this approach can PAH and biomarker ratios, a number of ‘‘source-spe-
be positively used for oil spill identification. On the cific’’ isomeric PAHs within the same alkylation levels
other hand, it has been demonstrated that the position were quantified, and their relative abundance ratios were
of the alkylation on the PAHs can influence the biode- computed in order to definitively identify and differen-
gradation rate of the isomers within an isomer group. tiate the sources of the tarballs. The selected isomers
This information can be used to sort out environmental include three isomers each within C3 -naphthalenes and
factors such as the impact of biodegradation on the C4 -naphthalenes, four isomers each within C1 -phen-
PAH distribution and to differentiate oil compositional athrenes and C2 -phenathrenes, two isomers within
changes due to physical weathering from those due to C4 -phenathrenes, and three isomers each within C1 -flu-
biodegradation. orenes and C1 -dibenzothiophenes. Comprehensive ana-
GC–MS method has been developed for the differen- lysis results revealed that California/Oregon tarball
tiation and source identification of crude, weathered and samples were chemically similar to the British Columbia/
biodegraded oils using the relative abundances of three Washington tarball samples, but they originated from
isomeric methyldibenzothiophene (4-:2-/3-:1-m-DBT) two different Bunker type fuels.
compounds (Wang and Fingas, 1995b). A database of
the ratios of the C1 –DBT isomers for several hundred
crude, weathered and biodegraded oils, and petroleum 6. Oil correlation and source identification by biomarker
products has been established. Fig. 7 plots 2-/3-methyl- fingerprinting analysis
dibenzothiophene versus 1-methyldibenzothiophene
(both isomers are normalized relative to 4-methyldi- Biomarker fingerprinting has historically been used
benzothiophene) using data from the database. Fig. 7 by petroleum geochemists in characterization of marine
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 439

oils in terms of source rock, genetic family, migration


and maturation properties, and in identification of pe-
troleum deposits. Chemical analysis of source-charac-
teristic and environmentally persistent biomarkers
generates information of great importance in determin-
ing the source of spilled oil, differentiating oils, moni-
toring the degradation process and weathering state of
oils under a wide variety of conditions. In the past de-
cade, use of biomarker fingerprinting techniques to
study spilled oils has greatly increased, and biomarker
parameters have been playing a prominent role in al-
most all oil spill work.

6.1. Biomarker distributions

Much of the knowledge of biomarkers and their di-


agnostic ratios comes from the oil geochemistry (Peters
and Moldowan, 1993). A wide variety of biomarkers
have been identified as being of use in characterization
of crude oils and oil fractions, including tricyclic, tetra-
cyclic and pentacyclic terpanes (m=z 191), methyl-
hopanes (m=z 205), steranes (m=z 217/218), methylsteranes Fig. 8. GC–MS chromatograms of biomarker terpanes at m=z 191 for
North Slope, Cook Inlet, Diesel No. 2, and Jet B fuel from Alaska. The
(m=z 217/231), and diasteranes (m=z 217/259). The dis-
Cook inlet oil showed significantly different biomarker distribution
tribution patterns are, in general, different from oil to from the North Slope oil with much lower total concentration of ter-
oil. As for refined products, no biomarker compounds panes and with only C30 ab-hopane being prominent. No biomarker
are detected in jet fuel and only trace C20 –C24 terpanes compounds were detected in Jet B fuel and only trace C20 –C24 terpanes
were present in the diesel, while most lube oils and hy- present in the diesel. Obviously, refinery processes have removed most
high molecular weight biomarkers from the corresponding crude oil
draulic fluids contain very high quantity of biomarkers.
feed stocks.
Obviously, refining processes have removed or concen-
trated most high molecular mass biomarkers from the
corresponding crude oil feed stocks.
an interpretational advantage in fingerprinting sources
Based on analysis of triterpane distribution patterns
of spilled oils and to provide additional diagnostic in-
and determination of two pentacyclic C27 triterpanes,
formation on the types of organic matter that give rise to
Shen (1984) distinguished four Arabian crudes, which in
the crude oil. For example, the geologically rare acyclic
their weathered forms were extremely similar to one
alkane botrycoccane (C34 H70 ) was used to identify a new
another. A study by Bieger et al. (1996) demonstrated
class of Australian non-marine crude oils (McKirdy et al.,
the use of tricyclic and pentacyclic terpanes as indicators
1986). The biomarkers 18a(H)-oleanane and 17a(H),
of the origin of diffuse lubricating oil contamination in
18a(H), 21b(H)-28, 30-bisnorhopane have been of spe-
plankton and sediments around St. JohnÕs, Newfound-
cial interest. The presence of biomarker 18a(H)-olean-
land, Canada. In this study, variable inputs of auto-
ane in benthic sediments in PWS, coupled with its
motive and outboard motor oils in different areas were
absence in Alaska North Slope crude and specifically in
clearly recognized and identified.
Exxon Valdez oil and its residues, confirmed another
As an example, Fig. 8 shows GC–MS chromatograms
petrogenic source (Bence et al., 1996; Page et al., 1996).
at m=z 191 for North Slope, Cook Inlet, Diesel No. 2,
Other ‘‘specific’’ pentacyclic terpanes include gammace-
and Jet B fuel from Alaska. Fig. 8 clearly demonstrates
rane, moretanes, lupanes, and bicadinanes.
that the biomarker distributions are significantly differ-
ent from oil to oil, and from oil to petroleum products.
6.3. Sesquiterpane and diamondoid biomarkers for corre-
lation of oil and refined products
6.2. Unique biomarker compounds
Various bicyclic sesquiterpanes and diamondoid com-
Biomarker terpanes and steranes are common con- pounds boiling within diesel range have been identified
stituents of crude oils. However, a few ‘‘specific’’ bio- in a number of crude oils. Diamondoid hydrocarbons
marker compounds including several geologically rare (such as adamantane, C10 , and diamantane, C14 , and their
acyclic alkanes are found to exist only in certain oils alkyl homologues) are rigid, three-dimensionally fused-
and, therefore, can be used as unique markers to provide cyclohexane-ring alkanes that have a diamond-like
440 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

structure. These biomarker compounds are thermally • BTO: 17a(H), 18a(H), 21b(H)-28,30-bisnorhopane (B),
stable and highly resistant to biodegradation. 17a(H), 18a(H), 21b(H)-25,28,30-trisnorhopane (T)
In a series cracking experiments in the laboratory, and oleanane (O).
Dahl et al. (1999) found that the increase in methyl- • steranes C27 abb/C29 abb, C28 abb=ðabb þ aaaÞ, and
diamantane (C15 ) concentration is directly proportional C29 abb=ðabb þ aaaÞ.
to the extent of cracking, indicating that under the
conditions of their experiments, diamondoids are nei- The triplet ratio, in general, varies in oils from dif-
ther destroyed nor created. Instead, they are conserved ferent sources and is dependent upon sources, deposi-
and concentrated, and hence can be considered to be a tional environment and maturity. The triplet ratio was
naturally occurring ‘‘internal standard’’ by which the first used in a chemistry study of North Slope crude by
extent of oil destruction can be determined. Chen et al. Kvenvolden et al. (1985), in which the ratio is 2. Exxon
(1996) identified 26 diamondoid hydrocarbons in oil Valdez oil (an Alaska North Slope crude) and its resi-
and source rock samples from the Tarim, Yinggehai, dues also have triplet ratios of 2; in contrast, many
Qiongdongnan and other Chinese basins. Two dia- tarballs collected from the shorelines of the Sound have
mondoid hydrocarbon ratios have been developed and triplet ratios of 5.
used as novel high maturity indices to evaluate the During the Arrow oil spill work, the ratio of the most
maturation and evolution of crude oils and condensates abundant C29 –C30 hopane was defined and used as a
in these Chinese basins. Noble et al. (1986) compared reliable source indicator (Wang et al., 1994b). Recently,
the chromatographic distributions of bicyclic sesquiter- Zakaria et al. (2000) studied oil pollution in the Straits
panes (m=z 123) of two weathered diesel fuel #2 samples of Malacca. Various samples including Malaysia oil,
from two adjacent petroleum terminal properties. The Middle East crude oils, South East Asian crude oils,
samples have been highly weathered with n-alkanes tarballs, sediments, and mussels were collected and an-
being completely lost. However, GC/MS (m=z 123) ana- alyzed. The analytical results in
Pthis study demonstrated
lysis results showed very different distribution profiles the utility of C29 /C30 ratio, (C31 –C35 )/C30 , and ho-
of sesquiterpane compounds between two samples, mohopane index as molecular tools to distinguish the
strongly indicating that two sources of diesel existed in source of petroleum in the Straits of Malacca. Barakat
the study area. et al. (1997) studied biomarker properties of five crude
Clearly, the smaller and lower boiling sesquiterpane oils from the Gulf of Suez, Egypt. The results reveal
and diamondoid biomarkers are useful and promising significant difference in biomarker distribution within
for source correlation and differentiation of refined the oils that suggest two oil types and one mixed type.
products, in particular for the weathered refined pro- Type 1 oils show a high relative abundance of gam-
ducts, because of their stability and resistance to bio- macerane indicating a marine saline-source depositional
degradation. It can be anticipated that more work will environment. Another related feature of these oils is the
be published in this fertile area of research (Stout et al., predominance of C35 over C34 17a(H)-homohopanes.
2002; Philp, 2002a). Type 2 oils show oleanane indices over 20%, indicating
they originated from an angiosperm-rich, Tertiary
6.4. Diagnostic ratios of biomarkers source rock. Type 3 oil has geochemical characteristics
intermediate between type 1 and 2 oils. Barakat et al.
A number of ratios of selected terpanes and steranes (2002) recently also reported a study results using mo-
has been defined and increasingly used for tracking lecular ratios of triaromatic sterane compounds (ratios
spilled oils and identifying spill sources. The diagnostic of C28 20R/C28 20S, C27 20R/C28 20R, and C28 20s/[C26
ratios of biomarkers most used include the following: 20R + C27 20S] triaromatic steranes) for oil weathering
monitoring and source correlation.
• Tm=Ts: 17a(H), 21b(H)-22,29,30-trisnorhopane/18a(H),
21b(H)-22,29,30-trisnorhopane. 6.5. Effect of biodegradation on biomarker distribution
• Triplet ratio: [C26 -tricyclic terpane (S?) + C26 -tricyclic
terpane (R?)]/C24 -tetracyclic terpane. The effects of biodegradation on biomarkers have
• C23 /C24 : C23 -tricyclic terpane/C24 -tricyclic terpane. been investigated by many authors. Several in vitro and
• C29 /C30 : 17a(H), 21b(H)-norhopane/17a(H), 21b(H)- in situ studies have shown that hopane and sterane
hopane. compounds are very resistant to biodegradation. In
• C23 /C30 and C24 /C30 . These ratios are used as source laboratory studies of biodegradation of nine Alaska oils
parameters in biodegradation assessments. and oil products (Wang et al., 1997a) and eight Cana-
• C31 S/(S + R), C32 S/(S + R), C33 S/(S + R): These al- dian oils (Wang et al., 1998c) by a defined bacterial
kylhopane epimer ratios is typical terpane-maturity consortium incubated under freshwater and cold/marine
parameter used extensively in petroleum geochemis- conditions, it is found that the fingerprint patterns of
try. triterpanes and steranes showed no changes after incu-
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 441

bation, despite extensive saturate and aromatic losses, compounds with the abundance of 2,6-bis(1,1-dimethyl-
and the ratios of selected paired biomarkers also re- ethyl)-phenol being largely higher than N-phenyl-1-
mained constant. Therefore, biomarkers can, in many naphthalenamine.
cases, be and have been used as conserved internal ref- All these three compounds identified are antioxidant
erences (Butler et al., 1991; Prince et al., 1994) for oil compounds, also called inhibitors. They are added to
weathering and biodegradation studies, in particular for oxidizable organic materials (such as lubricants and
those highly degraded or long-term weathered residual gasoline) to retard autooxidation and, in general, to
oils and complex petroleum-contaminated environmen- prolong the useful life of the substrates. The identifica-
tal samples. tion and differentiation of these additives clearly sup-
However, it should be noted that in severely weath- ports the general conclusion obtained from PAH and
ered or long-term weathered oil, degradation of some biomarker characterization, that is, Samples 2996 and
biomarkers was observed (Williams et al., 1986; Seifert 2997 are identical and come from the same source, and
et al., 1984). Munoz et al. (1997) found that isoprenoids the sample 2998 is different from samples 2996 and 2997
were severely degraded and biomarkers were more or and does not come from the same source as samples
less altered eight years after an oil spill in a peaty 2996 and 2997 do.
mangrove in a tropical ecosystem. They also found that
norhopanes were the most biodegradation-resistant
among the studied terpane and sterane groups and the 8. Using the isotopic composition of oil hydrocarbons for
C30 ab-hopane appeared more sensitive to weathering oil correlation and source identification
than its higher homologues. Because of the differential
resistance of biomarkers to biodegradation, compari- The stable carbon composition of oils and their re-
sons of their relative ratios can be used to rank oils as to fined products is essentially a manifestation of the
the extent of biodegradation. In a very recent study on myriad physical and biological processes that influence
long-term fate and persistence of the spilled Metula oil oil formation and refining (Whittaker et al., 1995; Philp,
in a marine marsh environment, Wang et al. (2001) 2002a; Philp et al., 2002b). The complex isotopic frac-
found that in highly degraded asphalt pavement sam- tionation patterns induced by these processes result in
ples, even the most refractory biomarker compounds characteristic 13 C/12 C ratios that can be used to classify
showed some degree of biodegradation, and that the crude oils, petroleum products and tars and can be in-
biomarkers were generally degraded in the declining terpreted to elucidate oil depositional environment,
order of importance of diasterane > C27steranes > tri- maturity, migration and biodegradation. In many cases,
cyclic terpanes > pentacyclic terpanes > norhopanes  the absence or low concentrations of biomarkers and
C29 abb-steranes. Peters and Moldowan (1993) have high molecular weight PAHs in refined products pre-
created a ‘‘quasi-stepwise’’ sequence for assessing the clude their successful use for making unique correlation.
extent to which an oil has been biodegraded, based on An alternative and, sometimes, complimentary tech-
the relative abundances of the various biomarker hydro- nique for correlation of such products is evolving
carbon classes. through the use of combined GC-isotope ratio MS
(GC–IRMS).
The isotopic analysis technique involves combustion
7. Characterization of additives for source identification of samples to CO2 and H2 O at high temperature fol-
of refined products lowed by analysis of the isotopic composition of purified
CO2 using isotope ratio mass spectrometer (IRMS). The
Refined products often contain additives that can be ratio of 13 C/12 C is reported in the usual delta-notation
used for source identification, such as the alkyl lead (d13 C) in parts per thousand (‰) relative to the PeeDee
added to automotive gasolines in the US in the pre-1996 belemnite (PDB) standard. In a review Whittaker et al.
time (Stout et al., 2002). (1995), placed particular emphasis on a critical evalua-
Recently, Wang et al. (2002) identified major unknown tion of the potential use of bulk and compound specific
compounds with remarkable abundances in three un- carbon isotope ratio mass spectrometry in the elucida-
known oil samples with very similar bulk chemical com- tion of waste source terms and extent of weathering.
positions and nearly identical GC-FID chromatograms.
They are identified as 2,6-bis(1,1-dimethylethyl)-phenol, 8.1. Stable isotope of bulk hydrocarbons
butylated hydroxytoluene or 2,6-di-tert-butyl-4-methyl-
phenol, and N-phenyl-1-naphthalenamine. Samples 2996 The primary advantage of using the carbon-isotopic
and 2997 contained these three compounds with nearly technique lies in the analysis of samples that have un-
equal concentrations. However, butylated hydroxytolu- dergone extensive weathering, since the isotopic com-
ene (BHT) was not detected in the sample 2998. The positions of the whole oil remains largely unaffected and
sample 2998 only contained the first and the last can therefore can be used as a correlation parameter.
442 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

This was demonstrated by Kvenvolden et al. (1985, analyses on oil residues and oiled bird feathers sampled
1993), who used similarity and dissimilarity in isotopic along the Atlantic Coast of France. Bulk stable carbon
composition as further evidence of correlation between isotopic composition has been shown to be a valuable
weathered shoreline Exxon Valdez oils (d13 C ¼ 29:4  screening correlation tool as it confirms the link of
0:1‰) and North Slope oil (d13 C ¼ 29:2 to )29.5‰), samples collected in the north part of Atlantic Coast
and differentiation between Exxon Valdez oils, the pre- with the Erika oil spill. Molecular isotopic composition
spill petroleum hydrocarbon background Katalla seep of saturated hydrocarbons and of phenanthrene com-
oil (d13 C ¼ 25:7‰), and Monterey-type (California- pounds also allows unambiguous differentiation of
source) shoreline tars (d13 C ¼ 23:7  0:2‰), respec- samples related to the Erika oil spill from those due to
tively. tar ball incidents.
Isotopic fingerprinting may be particularly useful for The hydrogen isotope ratio is interesting. In com-
characterization of heavy oil contaminated land since it parison with the carbon isotopes, the hydrocarbon iso-
does not rely upon GC resolution of waste components topes are commonly fractionated to a much greater
and therefore avoids the primary analytical constraint extent and as a result display large variation in d values.
associated with refractory wastes. Whittaker et al. Pond et al. (2002) recently studied the effect of in vitro
(1996) have reported utility of the IRMS technique as a aerobic biodegradation on the hydrogen isotopic com-
screening tool for identifying a predominance of polar position of individual n-alkanes from crude oil. They
and asphaltenic material in heavy oils and for providing found that the n-alkanes that were degraded at the fast
an additional source correlation parameter for unde- rates also showed the largest overall isotopic fractiona-
graded oily contaminants. tion (12–15& deuterium enrichment), suggesting that
While most emphasis has been placed on the use of the lower molecular weight n-alkanes can be used to
bulk carbon isotopes for correlation purposes, there are monitor in situ bioremediation of crude oil contamina-
a number of forensic applications that have used other tion.
stable isotopes. Kaplan et al. (1997) applied deuterium
values against carbon isotope values as a viable tool for
correlation of ground-water extracts and suspected 9. Oil spill identification by statistical and numerical
gasoline. analysis

8.2. Compound-specific stable isotope analysis The quantitative oil analysis data, when used in
conjunction with multivariate statistical analyses, can
One of the most significant advances in geochemistry increase the precision of analytical data and provide an
in the past few years has undoubtedly been the devel- unbiased and defensible means to differentiate among
opment of individual hydrocarbon compound-specific qualitatively similar oils. In modern oil forensic inves-
stable isotope analysis. Several studies that describe the tigations, various statistical and numerical analysis
use of this technique to elucidate the origin of various techniques have been developed and applied to the data
hydrocarbon products found in the aquatic environment interpretation (Burns et al., 1997; Stout et al., 1998).
have been published (Hayes et al., 1990; OÕMalley et al., These powerful tools include principal components
1994, 1996; Dowling et al., 1995). In conjunction with analysis, discriminant analysis and cluster analysis.
the existing tools, Mansuy et al. (1997) utilized GC– Among them, principal components analysis (PCA) is
IRMS as a correlation tool for hydrocarbons spilled in the most widely used and powerful multivariate analysis
aquatic environments with their suspected source(s) technique. It is used to transform original sample com-
based on comparison of the isotopic composition of the position data into new, smaller and uncorrelated vari-
bulk saturates and aromatics and individual n-alkanes. ables called principle components. For data sets with
Differences in the isotope composition of n-alkanes are large number of interrelated variables, PCA is a pow-
immediately apparent and reflect their different origins. erful tool for analyzing the structure of the data and
In this work, a method for the correlation of severely reducing the dimensionality of the pattern vectors. By
biodegraded oils based on the isotopic characterization using the pattern recognition and PCA techniques,
of asphaltene pyrolysates was also discussed. OÕMalley spilled jet fuels which had undergone weathering in
et al. (1994) performed compound-specific carbon iso- subsurface environment were characterized chromato-
tope analysis on individual PAH compounds isolated graphically and then successfully classified (Lavine et al.,
from environmental samples and concluded that the 1995) as different jet fuel types (JP-4, Jet-4, JP-5, JP-8, or
dominant signatures identified in the surface sediments JPTS).
of St. JohnÕs Harbour and Conception Bay were at- Aboul-Kassim and Simoneit (1995a,b) have used a
tributed to crankcase oil. variety of statistical techniques for source oil identifi-
After the Erika spill, Mazeas and Budzinski (2002) cation. In their analysis of the aliphatic and aromatic
performed both bulk and individual compound isotopic compositions in particulate fallout samples (PFS) in
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 443

Alexandria (Aboul-Kassim and Simoneit, 1995a), mul- can be matched to their sources even years after the spill
tivariate statistical analyses, including extended Q-mode occurred by the use of petroporphyrin analysis. Appli-
factor analysis and linear programming, were performed cation of petroporphyrins (Xu et al., 1994) as chemical
in order to reduce the hydrocarbon data set into a indicators of soil contamination by crude oils has been
meaningful number of end members (sources). Their also reported.
analysis indicated that there are two significant end Unresolved Complex Mixture (UCM), or ‘‘hump’’ of
members explaining 90% of the total variation among hydrocarbons, are a common feature of the gas chro-
the samples and confirming petrochemical (79.6%) and matograms of crude oils and certain refined products
thermogenic/pyrolytic (10.4%) sources in the PFS model. such as lubricating oils, and it is especially pronounced
In a study of sediment samples in the Eastern Har- for weathered and biodegraded oils and oil-polluted
bour (EH) of Alexandria (Aboul-Kassim and Simoneit, sediment extracts. It is perhaps surprising that virtually
1995b), a similar multivariate statistical approach, in- little is known about UCM compositions and molecular
cluding both factor analysis and linear programming structures, even though the concentrations of these
techniques, was used to determine the end member com- components in oils are significant. Work by several au-
positions and evaluate sediment partitioning and trans- thors (Gough and Rowland, 1990, 1991; Killops et al.,
port in the EH area. In this study, Aboul-Kassim and 1990) have shown that oxidative degradation of UCM
Simoneit determined that the untreated sewage was yields some gas chromatographically resolved products.
the main source of petroleum hydrocarbons in the EH It has been suggested that these products, which were
area rather than direct inputs from boating activities or identified to be acids, lactones and ketones, may be
urban run-off. useful for ‘‘fingerprinting’’ UCM. The oxidation prod-
In a case study, Stout et al. (2001) reported a strategy ucts of UCM isolated from sediments known to be
and methodology for defensibly correlating spilled oil to contaminated with a tank oil (Howells et al., 1986) have
source candidates. By using principle component anal- been successfully used to identify the source oils. There
ysis method (PCA), only six of the 29 vessels containing is much to be learned about UCM and the subject
oils were positively identified to be correlated to the should provide a fruitful area for further research.
spilled oil.
The application of sophisticated statistical analysis
techniques to oil analysis is a relatively new and dy- 11. Distinguishing biogenic hydrocarbons from petrogenic
namic area of research which enhances the interpretive hydrocarbons
power of petroleum hydrocarbon fingerprinting and
promises to greatly improve the identification of oil spill Characterization and differentiation of hydrocarbons
sources. from different sources is an essential part of any objec-
tive oil spill study. After oil spills, oil hydrocarbons of-
ten mix with other background hydrocarbon sources in
10. Oil spill identification by petroporphyrins analysis and the impacted area. One of the potential sources of hy-
by characterization of UCM drocarbons contributing to the background is biogenic
hydrocarbons. Hydrocarbons from both anthropogenic
Porphyrins are naturally occurring pigments that are and natural sources including biogenic source are very
composed of four pyrrole rings. These near-planar common in the marine and inland environments.
structures are usually found in nature in the form of Biogenic hydrocarbons are generated either by bio-
metal complexes. The metalloporphyrins, also known as logical processes or in the early stages of diagenesis in
petroporphyrins, found in crude oil are the product of recent marine sediments. Biological sources include land
the metabolism of chlorophyll by microorganisms and plants, phytoplankton, animals, bacteria, macroalgae
are usually complexed with metals they have scavenged and microalgae.
from the environment, predominately nickel and vana- It has been recognized (Venkatesan, 1988; Cretney
dium. The chemical identity of petroporphyrins varies et al., 1987; Kolattukudy, 1976; Blumer and Young-
between sources, depending upon the biological condi- blood, 1975; Page et al., 1995; Bence et al., 1996) that
tions inherent to each site. Using HPLC (Xu and Lesage, the biogenic hydrocarbons have the following chemical
1992), these various compounds can be differentiated composition characteristics:
based on the different alkyl substituents attached to each
porphyrin ring. Each oil sample produces a petro- • n-alkanes show a distribution pattern of odd carbon-
porphyrin ‘‘fingerprint’’ which links the sample to its numbered alkanes being much abundant than even
geographic origins. By comparing these fingerprints, oils carbon-numbered alkanes in the range of n-C21 to
from different origins can then be distinguished. Similar n-C33 , resulting in unusually high carbon preference
to biomarker compounds, petroporphyrins break down index (CPI) values, which is defined as the sum of
very slowly in the environment. This means that oil spills the odd carbon-numbered alkanes to the sum of the
444 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

even carbon-numbered alkanes (oils characteristically changes in residual oil and vegetative recovery 25 years
have CPI values around 1.0); after the spills. The comprehensive chemical data (Wang
• notable absence of the ‘‘unresolved complex mixture et al., 1998a) from analysis of the Nipisi samples indi-
(UCM)’’ hump in the chromatograms; cate the following:
• pristane is often more abundant than phytane, sug-
gesting a phytoplankton input and resulting in abnor- • The Nipisi samples can be categorized into three
mally high pristane/phytane ratio values; groups plus the background group, according to the
• presence of ‘‘biogenic cluster’’(identified as olefinic contamination level and degradation degree of the
hydrocarbons of biogenic origin) in the gas chroma- samples.
tograms of the aromatic fractions; and • The background samples showed typical biogenic n-
• wide distribution of the biogenic PAH perylene, an alkane distribution in the range of C21 –C33 with abun-
unsubstituted PAH produced in subtidal sediments dances of odd-carbon-number n-alkanes being much
by a process known as early diagenesis. higher than that of even-carbon-number n-alkanes.
The biogenic cluster was also obvious and no UCM
In a study of hydrocarbon biogeochemical setting of was observed (see Fig. 9). No petrogenic hydrocar-
the Baffin Island oil spill (BIOS) experimental site, bons, in particular no alkylated PAH homologues
Cretney et al. (1987) found that the BIOS subtidal and petroleum-characteristic biomarker compounds
samples showed very high pristane/phytane ratios (5–15) such as pentacyclic hopanes and C27 –C29 steranes
and CPI values (3–11). High concentrations of pristane were detected. In addition, three vegetation biomar-
relative to phytane in most of beach and subtidal sedi- ker compounds with remarkable abundances were de-
ments indicate biological hydrocarbon input from a tected and identified.
marine biological source. In addition, the GC chroma- • Hydrocarbons in the Group 3 samples taken from a
tograms of the aromatic fractions were typified by the depth of 40–100 cm were identified to be mixtures
olefinic hydrocarbon clusters. This cluster is a common of vegetation hydrocarbons and lightly contaminated
feature of coastal marine subtidal sediments and is be- oil hydrocarbons. Fig. 9 presents the GC chromato-
lieved to be of marine biological (planktonic or bacte- grams of two representative background and Group
rial) origin. The possibility of in situ genesis of PAHs is 3 samples, illustrating the differences between bio-
indicated by the presence of perylene as a major PAH in genic hydrocarbons from petrogenic hydrocarbons.
almost all the beach and subtidal sediments, however, it
should be noted that it cannot be used alone as a de-
finitive source identification criterion because perylene is 12. Distinguishing pyrogenic hydrocarbons from petro-
also produced in combustion processes. In a recent genic hydrocarbons
study of chemical characterization of oil residues in the
Baffin Island intertidal sediment samples, Wang et al. PAH distributions are the most useful tool in distin-
(1995a) noted that the BIOS sample S-3 demonstrated guishing pyrogenic hydrocarbons from petrogenic hy-
some distinct characteristics of biogenic hydrocarbons drocarbons. It has been recognized that in general,
including much higher abundance of odd n-alkanes than pyrogenic PAHs are characterized by:
even n-alkanes in the range of n-C21 –n-C33 and high CPI
and pristane/phytane values. However, the presence of • the dominance of the unsubstituted compounds over
petrogenic hydrocarbons were also obvious, indicated their corresponding alkylated homologues (i.e.,
by the distribution of n-alkanes in a wide range from C15 skewed or sloped PAH distribution profile of C0 
to C40 and the notable presence of the chromatographic C1 > C2 > C3 > C4),
UCM. This conclusion was further confirmed by the • the dominance of the high molecular mass 4–6 ring
presence of petrogenic PAH and biomarker compounds, PAHs over the low molecular mass 2–3 ring PAHs,
which showed similar distribution pattern to other oil and
residue samples. Page et al., 1995, 1996 found that the • on the gross level PAH can comprise a much higher
background alkane distribution in the PWS benthic mass percentage in most pyrogenic source materials
sediments is dominated by biogenic components. than in most petrogenic source materials (Bjøeseth,
During the years 1970–1972 the Nipisi, Rainbow and 1985; Benlahcen et al., 1997; Blumer and Young-
Old Peace River pipeline spills occurred in the Lesser blood, 1975). In contrast, petrogenic PAHs exhibit
Slave Lake area of northern Alberta. The Nipisi spill the characteristic bell-shaped distribution profiles,
was by far the largest of the three spills and is also one of which are readily modified to the distribution profile
the largest land spills in Canadian history. The most of C0 < C1 < C2 < C3 by weathering or degradation.
recent field survey was conducted in 1995 in order to
determine which cleanup methods were most successful, As an example, Fig. 10 compares PAH fingerprints
and to provide up-to-date information about any for the 1994 Mobile Burn starting oil, burn residue, and
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 445

Fig. 9. GC chromatograms of two representative Nipisi samples: background sample N3-1 (left, biologic hydrocarbons) and Group 3 sample R7-1
(right, mixture of vegetation hydrocarbons and lightly contaminated oil hydrocarbons), illustrating differences between petrogenic and biogenic
hydrocarbon distributions. Top panel (A) and (D): GC-FID chromatograms for total petroleum hydrocarbon (saturates and aromatics) analysis.
Sample R7-1 showed distribution of mixed petrogenic and biogenic n-alkanes, while Sample N3-1 showed only typical biogenic n-alkane distribution
in the range C21 –C33 . The biogenic cluster was also obvious and no UCM was observed. Middle panel (B) and (E): total ion GC–MS chromatograms
of aromatic fractions. Surrogate: o-terphenyl; internal standard (IS): d14 -terphenyl. Bottom panel (C) and (F): GC–MS chromatograms (m=z 191) of
saturated fractions. Petrogenic alkylated PAH homologues and biomarkers were detected in Sample R7-1. In contrast, No petrogenic PAHs and
biomarkers were detected in sample N3-1. However, three vegetation biomarkers with remarkable abundances were detected.

soot sample, illustrating the distinguishing features of types of hydrocarbons: some of them originate from
pyrogenic PAH distribution from the petrogenic PAH natural sources (plant waxes with predominance of odd-
distribution (Wang et al., 1999b). Fig. 11 compares ex- chain n-alkanes); some originate from petroleum sources
tracted ion chromatograms at m=z 178, 228, 252, and (unresolved complex mixture), and some originate from
276 for the 1994 Mobile diesel, residue sample MB-16 the incomplete combustion of fossil fuel (unsubstituted
and soot sample TSP-B3. The changes in relative dis- PAHs). Gonzalez-Vila et al. (1991) investigated the in-
tribution patterns of selected PAHs clearly demonstrate fluence of different burning conditions upon the PAH
the formation of pyrogenic PAH from 3-ring anthracene patterns generated from combustion of pine biomass
to 6-ring indeno[1,2,3-cd]pyrene and benzo[ghi]perylene and found that when burned at moderate temperature
due to combustion. The concentrations of PAHs with for a long period of time, the most abundant homo-
five or more rings were many times greater in the smoke logues generated were alkylated derivatives of naph-
than in oil. thalene, whereas phenanthrene alkyl derivatives were
In order to obtain an approximation of the hydro- predominant after a forest wildfire, where higher tem-
carbon input by the river Elbe into the North sea, peratures were reached in less time. More condensed
Theobald et al. (1995) performed UV–VIS, GC and PAHs were detected only in minor amounts. Page et al.
GC–MS measurements for over 2.5 years at a fixed (1999) and Bence et al. (1996) identified combustion
station at the freshwater edge of the Elbe to determine products throughout the marine environment in Prince
the composition and concentration of hydrocarbons in William Sound and Gulf of Alaska. Two components
its water. The characterization results indicate three of pyrogenic PAHs occur in the benthic sediments: a
446 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

94-MB Diesel Other 3-6 ring PAHs


3-ring
Naph
7000 100
Concentration (µg/g oil) 6000
80
60 4-ring
40
5- & 6-ring
5000
20
4000 0

Acl

Ace

An

Py

BaA

BbF

BkF

BeP

BaP

Pe

IP

DA

BP
Fl
3000 Dibenz
Fluo
2000 Phen 2-ring 3-6 ring PAHs
Chry
1000
0
C0-N
C1-N
C2-N
C3-N
C4-N
C0-P
C1-P
C2-P
C3-P
C4-P
C0-D
C1-D
C2-D
C3-D
C0-F
C1-F
C2-F
C3-F
C0-C
C1-C
C2-C
C3-C
Bph
Acl
Ace
An

Py
BaA
BbF
BkF
BeP
BaP
Pe

DA
BP
Fl

lP
94-MB Residue Other 3-6 ring PAHs
3-ring
4-ring
7000 100
5- & 6-ring
Concentration (µg/g oil)

80
6000 60
5000 40
20
4000 Dibenz 0
Phen

Fl
Acl

Ace

An

Py

BaA

BbF

BkF

BeP

BaP

Pe

IP

DA

BP
3000 Naph
2000 Fluo 3-6 ring PAHs
Chry
1000
0
C0-N
C1-N
C2-N
C3-N
C4-N
C0-P
C1-P
C2-P
C3-P
C4-P
C0-D
C1-D
C2-D
C3-D
C0-F
C1-F
C2-F
C3-F
C0-C
C1-C
C2-C
C3-C
Bph
Acl
Ace
An
Fl
Py
BaA
BbF
BkF
BeP
BaP
Pe
IP
DA
BP
94-MB Soot (PM10-B1) Other 3-6 ring PAHs
5- & 6-ring
400 160
3-ring
Concentration (µg/g soot)

350 120 4-ring


300 80

40
250
0
200

IP
FL
Acl

Ace

An

Py

BaA

BbF

BkF

BeP

BaP

Pe

BP
DA
Naph
150 3-6 ring PAHs
Phen
Dibenz Fluo Chry
100
50
0
C0-N

C2-N

C4-N

C1-P

C3-P

C0-D

C2-D

C0-F

C2-F

C0-C

C2-C

Bph

Ace

FL

BaA

BkF

BaP

IP

BP

94-MB Soot (TSP-B1) Other 3-6 ring PAHs


5- & 6-ring
400 160
3-ring 4-ring
Concentration (µg/g soot)

350 120
80
300
40
250 0
IP
FL
Acl

Ace

An

Py

BaA

BbF

BkF

BeP

BaP

Pe

BP
DA

200
150 Naph 3-6 ring PAHs
100 Phen Chry
Dibenz Fluo
50
0
C0-N
C1-N
C2-N
C3-N
C4-N
C0-P
C1-P
C2-P
C3-P
C4-P
C0-D
C1-D
C2-D
C3-D
C0-F
C1-F
C2-F
C3-F
C0-C
C1-C
C2-C
C3-C
Bph
Acl
Ace
An
Fl
Py
BaA
BbF
BkF
BeP
BaP
Pe
IP
DA
BP

Fig. 10. PAH fingerprints and distinguishing features of distribution patterns between petrogenic and pyrogenic PAHs for the starting oil, burn
residue, and soot samples from 1994 Mobile burn study. N, P, D, F, and C represent naphthalene, phenanthrene, dibenzothiophene, fluorene, and
chrysene, respectively; 0–4 represent carbon number of alkyl groups in the alkylated PAH homologous series. The abbreviations from Acl to BgP
represent the other EPA priority unsubstituted PAHs (please refer Table 2 for the full names of these PAHs). For comparison, the fingerprints of the
other 3–6 ring PAHs have been enlarged and shown in the left insets. Note that for clarity, different Y -axis scales are used for soot samples.

relatively uniform, low level, regional pre-spill back- local relatively high background, at some nearshore lo-
ground believed to be due to atmospheric input, and a cations, that is a consequence of current or historical
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 447

Fig. 11. Comparison of representative extracted ion chromatograms at m=z 178, 228, 252, and 276 for the starting oil, burn residue MB-16, and soot
sample TSP-B3, illustrating changes in the relative distribution of unsubstituted PAH isomers and demonstrating the formation of pyrogenic PAHs
from 3-ring anthracene (An) to 6-ring indeno[1,2,3-cd]pyrene (IP) and benzo[ghi]perylene (BP) due to combustion. Refer Table 2 for the full names of
the remaining EPA priority unsubstituted PAHs shown in Fig. 2.

human activities. Probable local sources include the pyrogenic PAHs from other hydrocarbon sources
burning of coal, diesel fuel, and wood at communities, (Benlahcen et al., 1997; Blumer and Youngblood, 1975;
fish camps, and campsites, and creosote from marine Lake et al., 1979; Sicre et al., 1987), including phe-
installations. Identification of pyrogenic components in nanthrene/anthracene (Ph/An), phenanthrene/methyl-
subtidal sediments is based entirely upon the PAH phenanthrene (Ph/m-Ph), fluoranthene/pyrene (Fl/Py),
fraction. The dominance of the 4–6-ring unsubstituted benz[a]anthracene/chrysene (BaA/Ch), Ph/(Ph + An),
PAHs was very distinct in the local source samples. benzo[e]pyrene/(benzo[e]pyrene + benzo[a]pyrene), and
In field studies, it is often difficult to identify which indeno[1,2,3-cd]pyrene/(indeno[1,2,3-cd]pyrene + benzo-
PAHs have been introduced from petrogenic or pyro- [ghi]perylene). Benlahcen et al. (1997) reported a method
genic sources. This is because there are many ways in using the ratio phenanthrene/anthracene < 10 and fluo-
which PAHs are introduced into the environment that ranthene/pyrene > 1 to identify contamination sources of
the PAH signature from one source could be obscured combustion processes. By determination of characteris-
by PAHs from another source(s). In addition, under the tic ratios of three pairs of unsubstituted PAHs, Bauza de
comparable combustion conditions, the same amount of Mirabo et al. (1997) identified five pyrogenic hydrocar-
organic materials can yield largely different amount of bon sources which might be potential sources of PAHs
PAHs at different temperatures. Therefore, in addition to in aerosol samples collected from semi-rural stations in
qualitative criteria, quantitative criteria should be de- the island of Majorca (Spain). Zeng and Vista (1997)
fined to recognize sources of PAHs. Numerous quanti- applied several compositional indices of PAHs to iden-
tative diagnostic ratios have been defined to differentiate tify the sources of PAHs in the coastal environment off
448 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

San Diego, California and concluded that PAHs de- quantitative ratio, combined with other qualitative cri-
tected in the microlayer and sediments collected in 1994 teria, can be used to unambiguously differentiate sources
in San Diego Bay were mainly derived from combustion of PAHs.
sources rather than oil spills, despite the heavy shipping
activities in the area.
In the 1994 Mobile burn study (Wang et al., 1999b), 13. Allocation of complex hydrocarbons to multiple
we found that the Pyrogenic Index (which was defined as sources
the ratios of the total of the other EPA priority un-
substituted 3–6 ring PAHs toPthe total of five target 13.1. Quantitative allocation of complex hydrocarbons to
alkylated PAH homologues), (other 3–6 ring PAHs)/
P multiple sources
(5 alkylated PAHs), in the soot samples were signifi-
cantly different from that for crude oils and petroleum Chemical fingerprinting techniques to identify the
products and can be positively used to differentiate the sources of environmental pollutants have been in use for
pyrogenic and petrogenic PAHs. P For comparison, Fig. the last two decades and major advances have been
12
P depicts the relative ratios of (other 3–6 ring PAHs)/ made in recent years to address difficult oil spill issues on
(5 alkylated PAHs) versus the relative ratios of phe- source identification and, in particular, quantitative al-
nanthrene to anthracene determined from 1994 soot location of complex hydrocarbons to multiple sources.
samples and over 60 oils and petroleum products in- The major approach used in source identification and
cluding jet fuel, diesel, lube oil, Bunker C, Pand heavy allocation in complex mixtures include pattern recog-
fuel. Fig. 12 Pdepicts the relative ratios of (other 3–6 nition, source-specific diagnostic ratios of target ana-
ring PAHs)/ (5 alkylated PAHs) versus the relative lytes, and principal component analysis.
ratios of phenanthrene to anthracene using the data After the 1989 Exxon Valdez oil spill, contract re-
obtained from this study. It can be clearly seen from Fig. search organizations and university scientists collected
12 that the
P jet fuel, diesel, and most
P crudes show the and analyzed thousands of sediment samples from the
ratios of (other 3–6 ring PAHs)/ (5 alkylated PAHs) spill area. These samples were found to contain hydro-
to be smaller than 0.01 versus very scattered ratios of carbons from spill oil, diesel fuel, coal, combustion
phenanthrene/anthracene. Heavy oils (such as Cold products, creosote, products of natural biological pro-
Lake Bitumen and Orimulsion) and heavy fuels (such as cesses, and oil from natural oil seeps. Page et al. (1995,
IFO-180, A-02, IF-30, and Bunker C type oils) show 1996, 1999) and Boehm et al. (1998, 2001) used these
significantly higher ratios falling in the range of 0.01– advanced fingerprinting techniques, in particular distri-
0.05 (clusters 1 and 2). In sharp contrast, the average butions of PAHs and double ratios of C2D/C2P to C3D/
ratio for the six soot samples of 1994 Mobile Burns is as C3P, to identify and allocate four major PAH sources in
high as 1.13 (in a range of 0.8–2.0). Obviously, this the sediment samples: biological PAHs; combustion
product PAHs; natural petrogenic background PAHs
derived from oil seeps outside of the Sound; and pe-
troleum PAHs from ANS sources (i.e., either spilled oil
or diesel oil refined from ANS oil). Carbon isotopes and
terpane distribution distinguished Exxon Valdez resi-
dues found on shorelines from tars from other sources.
Diesel and diesel soot were identified by the absence of
alkylated chrysenes and a narrow distribution of n-alk-
anes, whereas pyrogenic products were distinguished by
the dominance of 4- to 6-ring PAHs over 2-to 3-ring
PAHs and by the dominance of non-alkylated over al-
kylated homologues of each PAH series. The back-
ground hydrocarbons present within Prince William
Sound and the Gulf of Alaska were identified to origi-
nate from natural oil seeps and eroding source rocks in
the same regions. Burns et al. (1997) developed a more
P
Fig. 12. Plot of the relative ratios of ðother 3–6 ring PAHsÞ= detailed method to more precisely allocate PAHs in
P
ð5-alkylated PAHsÞ over the relative ratios of phenanthrene/an- sediment samples to the PAH sources from which they
thracene for over 60 oils and petroleum products. Lighter petroleum came. This new approach uses principal component
P
products and P most crudes show the ratios of ðother 3–6 analysis to identify possible sources and a least-squares
ring PAHsÞ= ð5-alkylated PAHsÞ fall into a range of 0–0.01, while
heavy oils and heavy fuels show significantly higher ratios in the range model to find the source mix that gives the best fit of 36
of 0.01–0.05. The soot samples show the most striking increase in the PAH analytes (including chrysenes) in each sample. The
ratio, indicated by the right circle. method identified 18 possible PAH sources within PWS
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 449

of Alaska, including diesel oil, diesel soot, spilled crude hydrocarbon source identification and differentiation,
oil in various weathering states, natural background, when it is applied properly. However, in many cases,
creosote, and combustion products from human activi- particularly for complex hydrocarbon mixtures or ex-
ties and forest fires. The PCA and least-squares model- tensively weathered and degraded oil residues, there is
ling procedures can be generalized to allocate PAHs or no single fingerprinting technique which can meet the
other complex hydrocarbons to multiple sources in objectives of forensic investigation and quantitatively
other situations, including other spills. allocate hydrocarbons to their respective sources.
Short and co-worker from the NOAA Auke Bay Combined and integrated multiple ‘‘tools’’ are often
Laboratory cited evidence based on the presence of coal necessary under such situations.
nearby beaches and the extensive coal measures between Development in hydrocarbon fingerprinting tech-
Bering River and Ice Bay and agued that the back- niques will continue as analytical and statistical methods
ground hydrocarbons come from such deposit (Short evolve. It can be anticipated that these developments
et al., 1999). will further enhance the utility and defensibility of oil
Multivariate statistical techniques have become hydrocarbon fingerprinting.
available in recent years which allow signature to be
determined from source samples which can then be
quantitatively extracted from environmental samples. In
References
an attempt to resolve the issue of origin of background
hydrocarbons in the benthic sediments of PWS and the Aboul-Kassim, T.A.T., Simoneit, B.R.T., 1995a. Aliphatic and
Gulf of Alaska, Mudge (Mudge, 2002) uses this pow- aromatic hydrocarbons in particulate fallout of Alexandria, Egypt:
erful technique to reassess the percentage contribution sources and implications. Environ. Sci. Technol. 29, 2473–2483.
of coal, seep oil, shales and rivers to the hydrocarbon Aboul-Kassim, T.A.T., Simoneit, B.R.T., 1995b. Petroleum hydrocar-
loading in the Gulf of Alaska. The data analysis results bon fingerprinting and sediment transport assessed by molecular
biomarker and multivariate statistical analyses in the Eastern
suggests mixed sources whose contributions vary sig- Harbour of Alexandria, Egypt. Mar. Pollut. Bull. 30, 63–73.
nificantly across the sampling area. ASTM Method 3328-90 1997a. In: Annual Book of ASTM Standards,
Water (II), vol. 11.02. American Society for Testing and Materials,
13.2. Identification and allocation of multiple hydrocarbon Philadelphia, PA.
sources using the holistic approach and TOC ASTM Method 5037-90 1997b. In: Annual Book of ASTM Standards,
Water (II), vol. 11.02. American Society for Testing and Materials,
Philadelphia, PA.
Total organic carbon, (TOC) is one important tool in ASTM Method 5739-95 1997c. In: Annual Book of ASTM Standards,
holistic approach to hydrocarbon source fingerprinting Water (II), vol. 11.02. American Society for Testing and Materials,
(Boehm et al., 2002; Page et al., 2002). It can constrain Philadelphia, PA.
the contributions of specific sources and rule out in- Barakat, A.O., Mostafa, A., El-Gayar, M.S., Rullkotter, J., 1997.
Source-dependent biomarker properties of five crude oils from the
correct source allocation in cases where inputs are Gulf of Suez. Egypt Org. Geochem. 26, 441–450.
dominated by fossil organic carbon. In a study of a Barakat, A.O., Qian, Y., Kim, M., Kennicutt M.C., II. 2002.
benthic sediments from Prince William Sound (PWS) Compositional changes of aromatic steroid hydrocarbons in
and the Gulf of Alaska (GOA), Boehm et al. found naturally weathered oil residues in the Egyptian Western Desert.
excellent agreement between measured TOC and TOC In: Proceedings of the 25th Arctic and Marine Oil Spill Program
(AMOP) Technical Seminar. Environment Canada, Ottawa. pp.
calculated from hydrocarbon fingerprinting matches of 477–488.
PAH and biomarkers. Confirmation by two indepen- Bauza de Mirabo, F., Mateu, J., Forteza, R., Cerda, V., Colom, M.,
dent source indicators (TOC and fingerprint matches) Oms, M., 1997. Determination of aliphatic and aromatic com-
provides evidence that source allocation determined by pounds in aerosols collected at two semi-rural stations in Majorca
fingerprint matches are robust and that the major TOC (Spain). J. Environ. Sci. Health. A 32, 1415–1433.
Bence, A.E., Kvenvolden, K.A., Kennicutt II, M.C., 1996. Organic
sources have been correctly identified. They concluded geochemistry applied to environmental assessments of Prince
that the natural petrogenic hydrocarbon background in William Sound, Alaska, after the Exxon Valdez oil spill––a review.
sediments this area comes from eroding Tertiary shales Org. Geochem. 24, 7–42.
and associated oils seeps along the northern GOA coast. Bence, A.E., Burns, W.A., 1995. Fingerprinting hydrocarbons in the
biological resources of the Exxon Valdez spill area. In: Wells, P.G.,
Butler, J.N., Hughes, J.S. (Eds.), Exxon Valdez Oil Spill: Fate and
Effects in Alaska Waters. ASTM STP 1219. ASTM, Philadelphia,
14. Conclusions PA, pp. 84–140.
Benlahcen, K.T., Chaoui, A., Budzinski, H., Bellocq, J., Garrigues, P.,
The advances in petroleum hydrocarbon fingerprint- 1997. Distribution and sources of polycylic aromatic hydrocarbons
ing and data interpretation methods and approaches in in some Mediterranean coastal sediments. Mar. Pollut. Bull. 34,
298–305.
the last two decades have now allowed for detailed Bieger, T., Hello, J., Abrajano Jr., T.A., 1996. Petroleum biomarkers
qualitative and quantitative characterization of spilled as tracers of lubricating oil contamination. Mar. Pollut. Bull. 32,
oils. Chemical fingerprinting is a powerful ‘‘tool’’ for 270–274.
450 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

Bjøeseth, A., 1985. Sources of emissions of PAH. In: Bjøeseth, A., transported bitumen from the coastline of Australia. Org. Geo-
Ramdahl, T. (Eds.), Handbook of Polycyclic Aromatic Hydrocar- chem. 23, 729–737.
bons. Marcel Dekker, New York, pp. 1–20. EPA Methods for chemical analysis of water and wastes, EPA-600-4-
Blumer, M., Youngblood, W.W., 1975. Polycyclic aromatic hydrocar- 79-020, Environmental Monitoring and Support Laboratory.
bons in soils and recent sediments. Science 188, 53–55. Cincinnati, OH, 1983.
Boehm, P.D., Douglas, G.S., Burns, W.A., Mankiewicz, P.J., Page, EPA Methods for Chemical Analysis of Water and Wastes 1986. US
D.S., Bence, A.E., 1997. Application of petroleum hydrocarbon EPA, Washington, DC, 1986.
chemical fingerprinting and allocation techniques after the Exxon EPA test methods for evaluating solid waste (SW-846). Update III, US
Valdez oil spill. Mar. Pollut. Bull. 34, 599–613. EPA, Office of Solid Waste and Emergency Response, Washington,
Boehm, P.D., Page, D.S., Gilfillan, E.S., Bence, A.E., Burns, W.A., DC, 1997.
Mankiewicz, P.J., 1998. Study of the fate and effects of the Exxon EPA Method, 1664, Revision A: n-Hexane extractable material and
Valdez oil spill on benthic sediments in two bays in Prince William silica gel treated n-hexane extractable material by extraction and
Sound, Alaska. 1. Study design, chemistry and source fingerprint- gravimetry. EPA-821-R-98-002, US EPA, Office of Water, Wash-
ing. Environ. Sci. Technol. 32, 567–576. ington, DC, 1999.
Boehm, P.D., Page, D.S., Burns, W.A., Bence, A.E., Mankiewicz, P.J., Etkin, D.S. (Ed.), 1998. Nearly 300 million gallons of oil spilled into
Brown, J.S., 2001. Resolving the origin of the petrogenic hydro- US marine waters since 1960s. Oil Spill Intelligence ReportÕs White
carbon background in Prince William Sound. Alaska Environ. Sci. Paper Series, vol. 2(2), Cutter Information Corp.
Technol. 35, 471–479. Fayad, N.M., Overton, E., 1995. A unique biodegradation pattern
Boehm, P.D., Burns, W.A., Page, D.S., Bence, A.E., Mankiewicz, P.J., of the oil spilled during the 1991 Gulf war. Mar. Pollut. Bull. 30,
Brown, J.S., Douglas. G.S., 2002. Total organic carbon, an 239–246.
important tool in holistic approach to hydrocarbon source finger- Fingas, M., 1995. A literature review of the physics and predictive
printing. In: Proceedings of the 25th Arctic and Marine Oil Spill modelling of oil spill evaporation. J. Hazard. Mater. 42, 157–175.
Program (AMOP) Technical Seminar, Environment Canada, Gains, R.B., Frysinger, G.S., Hendrick-Smith, M.S., Stuart, J.D.,
Ottawa, pp. 413–428. 1999. Oil spill source identification by comprehensive two-dimen-
Bragg, J.R., Owens, E.H., 1994. Clay-oil flocculation as a natural sional gas chromatography. Environ. Sci. Technol. 33, 2106–
cleansing process following oil spill: Part 1––studies of shoreline 2112.
sediments and residues from past spills. In: Proceedings of the 17th Garrett, P.M., Pickerring, I.J., Haith, C.E., Prince, R.C., 1998.
Arctic and Marine Oil Spill Program (AMOP) Technical Seminar. Photooxidation of crude oils. Environ. Sci. Technol. 32, 3719–
Environment Canada, Ottawa, Ontario, pp. 1–25. 3723.
Brown, J.S., Boehm, P.D., 1993. The use of double ratio plots of Gonzalez-Vila, F.J., Lopez, J.L., Martin, F., del Rio, J.C., 1991.
polynuclear aromatic hydrocarbon (PAH) alkyl homologues for Determination in soils of PAH produced by combustion of
petroleum source identification. In: Proceedings of the 1993 biomass under different conditions. Frenes. J. Anal. Chem. 339,
International Oil Spill Conference, American Petroleum Institute, 750–753.
Washington, DC, pp. 799–801. Gough, M.A., Rowland, S.J., 1990. Characterization of unresolved
Burns, W.A., Mankiewicz, P.J., Bence, A.E., Page, D.S., Parker, K.R., complex mixture of hydrocarbons in petroleum. Nature 344, 648–
1997. A principal-component and least-squares method for allo- 650.
cating sediment hydrocarbons to multiple sources. Environ. Gough, M.A., Rowland, S.J., 1991. Characterization of unresolved
Toxicol. Chem. 16, 1119–1131. complex mixtures of hydrocarbons from lubricating oil feedstocks.
Butler, E.L., Douglas, G.S., Steinhauter, W.S., Prince, R.C., Axcel, T., Energy Fuels 5, 869–874.
Tsu, C.S., Bronson, M.T., Clark, J.R., Lindstrom, J.E., 1991. Hayes, J.M., Freeman, K.H., Popp, B.N., Hohan, C.H., 1990.
Hopane, a new chemical tool for measuring oil biodegradation. In: Compoundspecific isotope analysis: a novel tool fpr reconstruction
Hinchee, R.E., Olfenbuttel, R.F. (Eds.), On-site Reclamation. of ancient biogeochemical process. Org. Geochem. 16, 1115–1128.
Butterworth-Heinemann, Boston, MA, pp. 515–521. Hostettler, F.D., Rosenbauer, R.J., Kvenvolden, K.A., 1999. PAH
Chen, J., Fu, J., Sheng, G., Liu, D., Zhang, J., 1996. Diamondoid refractory index as a source discriminant of hydrocarbon input
hydrocarbon ratios: novel maturity indices from highly mature from crude oil and coal in Prince William Sound, Alaska. Org.
crude oils. Org. Geochem. 25, 179–190. Geochem. 30, 873–879.
Christensenand, L.B., Larsen, T.H., 1993. Method for determining the Howells, S., Patey, J., Turner, C., Dodd, N.M., 1986. Oil Pollution
age of diesel oil spills in the soil. Ground Water Monitor. Research Unit Report 27. Field Studies Council, Angle, Dyfed,
Remediat. Fall, 142–149. Wales.
Cretney, W.J., Green, D.R., Fowler, B.R., Humphrey, B., Fiest, D.L., Jordan, R.E., Payne, J.R., 1980. Fate and Weathering of Petroleum
Boehm, P.D., 1987. Hydrocarbon biogeochemical setting of the Baffin Spills in the Marine Environment: A Literature Review and
Island oil spill experimental site, I, Sediments. Arctic 40, 51–55. Synopsis. Ann Arbor Science Publishers, Ann Arbor, Michigan.
Dahl, J.E., Moldowan, J.M., Peters, K.E., Claypool, G.E., Rooney, Jokuty, P., Whiticar, S., Wang, Z.D., Fingas, M., Fieldhouse, B.,
M.A., Michael, G.E., Mello, M.R., Kohnen, M.L., 1999. Dia- Lambert, P., Mullin, J., 1999. Properties of Crude Oils and Oil
mondoid hydrocarbons as indicators of natural oil cracking. Products (Report Series No. EE-165). Environment Canada,
Nature 399, 54–57. Ottawa, ON.
Daling, P.S., Faksness, L.G., Hansen, A.B., Stout. S.A., 2002. Kaplan, I.R., Alimi, M., Galperin, Y., Lee, R.P., Lu, S., 1995. Pattern
Improved and standardized methodology for oil spill fingerprint- of chemical changes in fugitive hydrocarbon fuels in the environ-
ing. In: Proceedings of the 25th Arctic and Marine Oil Spill ment. In: Presented at the 1995 SPE/EPA Exploration and
Program (AMOP) Technical Seminar, Environment Canada, Production Environmental Conference, Houston, TX, March 27–
Ottawa, pp. 429–454. 29, 1995, SPE Paper No. 29754.
Douglas, G.S., Bence, A.E., Prince, R.C., McMillen, S.J., Butler, E.L., Kaplan, I.R., Galperin, Y., Alimi, H., Lee, R.P., Lu, S., 1996. Pattern
1996. Environmental stability of selected petroleum hydrocarbon of chemical changes during environmental alteration of hydrocar-
source and weathering ratios. Environ. Sci. Technol. 30, 2332– bon fuels. Ground Water Monitor. Remediat. Fall, 113–124.
2339. Kaplan, I.R., Galperin, Y., Lu, S., Lee, R.P., 1997. Forensic
Dowling, L.M., Boreham, C.J., Hope, J.M., Marry, A.P., Summons, environmental geochemistry differentiation of fuel-types, their
R.E., 1995. Carbon isotope composition of hydrocarbons in ocean sources, and release time. Org. Geochem. 27, 289–317.
Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452 451

Kennicutt II, M.C., 1998. The effect of biodegradation on crude oil human activity: a case study in Prince William Sound. Alaska Mar.
bulk and molecular composition. Oil Chem. Pollut. 4, 89–112. Pollut. Bull. 38 (4), 247–260.
Kolattukudy, P.E., 1976. Chemistry and Biochemistry of Natural Page, D.S., Bence, A.E., Burns, W.A., Boehm, P.D., Douglas. G.S.,
Waxes. Elsevier, New York. 2002. A holistic approach to hydrocarbon source allocation in the
Krahn, M.M., Ylitalo, G.M., Buzitis, J., Chan, S., Varanasi, U., 1993. subtidal sediments of Prince, William Sound, Alaska, Embay-
Rapid HPLC methods that screen for aromatic compounds in ments. In: Proceedings of the 25th Arctic and Marine Oil Spill
environmental samples. J. Chromatogr. 642, 15–32. Program (AMOP) Technical Seminar, Environment Canada,
Kvenvolden, K.A., Rapp, J.B., Bourell, J.H., 1985. In: Magoon, L.B., Ottawa, pp. 547–563.
Claypool G.E. (Eds.), Alaska North Slope Oil/Rock Correlation Peters, K.E., Moldowan, J.W., 1993. The Biomarker Guide : Inter-
Study, American Association of Petroleum Geologists Studies in preting Molecular Fossils in Petroleum and Ancient Sediments.
Geology, No. 20, pp. 593–617. Prentice-Hall, New Jersey.
Kvenvolden, K.A., Hostettler, F.D., Rapp, J.B., Carlson, P.R., 1993. Philp, R.P., 2002a. Application of stable isotopes and radioisotopes in
Hydrocarbon in oil residues on beaches of islands of Prince environmental forensics. In: Murphy, B.L., Morrison, R.D. (Eds.),
William Sound. Alaska Mar. Pollut. Bull. 26, 24–29. Introduction to Environmental Forensics. Academic Press, Lon-
Kvenvolden, K.A., Hostettler, F.D., Carlson, P.R., Rapp, J.B., don, pp. 99–136. Chapter 5.
Threlkeld, C.N., Warden, A., 1995. Ubiquitous tarballs with a Philp, R.P., Allen, J., Kuder, T., 2002. The use of isotopic composition
California-source signature on the shorelines of Prince William of individual compounds for correcting spilled oils and refined
Sound. Alaska Environ. Sci. Technol. 29, 2684–2694. products in the environment with suspected sources. In: Proceed-
Lake, J.L., Norwood, C., Dimock, C., Bowen, R., 1979. Origins of ings of the 25th Arctic and Marine Oil Spill Program (AMOP)
PAHs in estuarine sediments. Geochim. Cosmochim. Acta 43, Technical Seminar, Environment Canada, Ottawa, pp. 565–580.
1847–1854. Pond, K.L., Huang, Y., Wang, Y., Kulpa, C.F., 2002. Hydrogen
Lavine, B.K., Mayfield, H., Kromann, P.R., Faruque, A., 1995. isotopic composition of individual n-alkanes as an intrinsic tracer
Source identification of underground fuel spills by pattern recog- for bioremediation and source identification of petroleum contam-
nition analysis of high-speed gas chromatograms. Anal. Chem. 67, ination. Environ. Sci. Technol. 36, 724–728.
3846–3852. Prince, R.C., 1993. Petroleum Spill Bioremediation in Marine Envi-
Leahy, J.G., Colwell, R.R., 1990. Microbial degradation of hydrocar- ronment. Crit. Rev. Microbial 19, 217–242.
bons in the environment. Microbial Rev. 54, 305–315. Prince, R.C., Elmendorf, D.L., Lute, J.R., Hsu, C.S., Haith, C.E.,
Lundanes, E., Greibrokk, T., 1994. Separation of fuels, heavy Senius, J.D., Dechert, G.J., Douglas, G.S., Butler, E.L., 1994.
fractions, and crude oils into compound classes: a review. J. High 17a(H), 21b(H)-hopane as a conserved internal marker for
Resolut. Chromatogr. 17, 197–202. estimating the biodegradation of crude oil. Environ. Sci. Technol.
Mansuy, L., Philp, R.P., Allen, J., 1997. Source identification of oil 28, 142–145.
spills based on the isotopic composition of individual components Sasaki, T., Maki, H., Ishihary, M., Harayama, S., 1998. Vanadium as
in weathered oil samples. Environ. Sci. Technol. 31, 417–3425. an internal marker to evaluate microbial degradation of crude oil.
Mazeas, L., Budzinski, H., 2002. Molecular and stable carbon isotope Environ. Sci. Technol. 32, 3618–3621.
source identification of oil residues and oiled bird feathers samples Sauer, T.C., Boehm, P.D., 1991. The use of defensible analytical
along the Atlantic Coast of France after the Erika oil spill. chemical measurements for oil spill natural resource damage
Environ. Sci. Technol. 36, 130–137. assessment. In: Proceedings of the 1991 Oil Spill Conference.
McKirdy, D.M., Cox, R.E., Volkman, J.K., Howell, V.J., 1986. American Petroleum Institute, Washington DC, pp. 363–370.
Botryococcane in a new class of Australian non-marine crude oils. Sauer, T.C., Brown, J.S., Boehm, P.D., Aurand, D.V., Michel, J.,
Nature 320, 57–59. Hayes, M.O., 1993. Hydrocarbon source identification and weath-
Mudge, S.M., 2002. Reassessment of the hydrocarbons in Prince ering characterization of intertidal and subtidal sediments along
William Sound and the Gulf of Alaska: identifying the source using the Saudi Arabian coast after the Gulf war oil spill. Mar. Pollut.
partial least-squares. Environ. Sci. Technol. 36, 2354–2360. Bull. 27, 117–134.
Munoz, D., Guiliano, M., Doumenq, P., Jacquot, F., Scherrer, P., Sauer, T.C., Boehm, P.D., 1995. Hydrocarbon chemistry analytical
Mille, G., 1997. Long term evolution of petroleum biomarkers in methods for oil spill assessments. Technical Repot 95-032. Marine
mangrove soil. Mar. Pollut. Bull. 34, 868–874. Spill Response Corporation, Washington, DC.
Noble, R.A., Alexander, R., Kagi, R.I., Knox, J., 1986. Identification Sauer, T.C., Michel, J., Hayes, M.O., Aurand, D.V., 1998. Hydrocar-
of some diterpenoids hydrocarbons in petroleum. Org. Geochem. bon characterization and weathering of oiled intertidal sediments
10, 825–829. along the Saudi Arabian Coast two years after the Gulf war oil
OÕMalley, V.P., Abrajano, T.A., Hellou, J., 1994. Determination of the spill. Environ. Int. 24, 43–60.
13
C/12 C ratios of individual PAH from environmental samples: can Seifert, W.K., Moldowan, J.M., Demaison, G.J., 1984. Source
PAH sources be apportioned? Org. Geochem. 21, 809–822. correlation of biodegraded oils. Org. Geochem. 6, 633–643.
OÕMalley, V.P., Abrajano, T.A., Hellou, J., 1996. Stable carbon Shen, J., 1984. Minimization of interferences from weathering effects
isotopic apportionment of individual PAHs in St. JohnÕs harbour, and use of biomarkers in identification of spilled crude oils by
Newfoundland. Environ. Sci. Technol. 30, 634–639. gas chromatography/mass spectrometry. Anal. Chem. 56, 214–
Page, D.S., Boehm, P.D., Douglas, G.S., Bence, A.E., 1995. Identi- 217.
fication of Hydrocarbon Sources in the Benthic Sediments of Short, J.W., Jackson, T.J., Larsen, M.L., Wade, T.L., 1996. Analytical
Prince William Sound and the Gulf of Alaska Following the Exxon methods used for the analysis of hydrocarbons in crude oil tissues
Valdez Spill. In: Wells, P.G., Butler, J.N., Hughes, J.S. (Eds.), sediments and seawater collected for the natural resources damage
Exxon Valdez Oil Spill: Fate and Effects in Alaska Waters. ASTM, assessment of the Exxon Valdez oil spill. Am. Fish. Soc. Symp. 18,
Philadelphia, PA, pp. 41–83. 140–148.
Page, D.S., Boehm, P.D., Douglas, G.S., Bence, A.E., Burns, W.A., Short, J.W., Heintz, R.A., 1997. Identification of Exxon Valdez oil in
Mankiewicz, P.J., 1996. The natural petroleum hydrocarbon sediments and tissues from Prince William Sound and the
background in subtidal sediments of Prince William Sound, Northwestern Gulf of Alaska based on a PAH weathering model.
Alaska, USA. Environ. Toxicol. Chem. 15 (8), 1266–1281. Environ. Sci. Technol. 31, 2375–2384.
Page, D.S., Boehm, P.D., Douglas, G.S., Bence, A.E., Burns, W.A., Short, J.W., Kvenvolden, K.A., Carlson, P.R., Hostettler, F.D.,
Mankiewicz, P.J., 1999. Pyrogenic PAHs in sediments record past Rosenbauer, R.J., Wright, B.A., 1999. Natural hydrocarbon
452 Z. Wang, M.F. Fingas / Marine Pollution Bulletin 47 (2003) 423–452

background in benthic sediments of Prince William Sound, Alaska: linkage of tarballs from the coasts of Vancouver Island and
oil vs. coal. Environ. Sci. Technol. 33, 34–42. Northern California using GC/MS and isotopic techniques. J. High
Short, J.W., 2002. Oil identification based on a goodness-of-fit metric Resolut. Chromatogr. 21, 383–395.
applied to hydrocarbon analysis results. In: Proceedings of the 25th Wang, Z.D., Fingas, M., Blenkinsopp, S., Sergy, G., Landriault, M.,
Arctic and Marine Oil Spill Program (AMOP) Technical Seminar, Sigouin, L., Foght, J., Semple, K., Westlake, D.W.S., 1998c.
Environment Canada, Ottawa, pp. 581–592. Comparison of oil composition changes due to biodegradation and
Sicre, M.A., Marty, J.C., Salion, A., Aparicio, X., Grimalt, J., physical weathering in different oils. J. Chromatogr. A 809, 89–107.
Albaiges, J., 1987. Aliphatic and aromatic compounds in aerosols. Wang, Z.D., Fingas, M., Page, D., 1999a. Oil Spill Identification.
Atmos. Environ. 21, 2247–2259. J. Chromatogr. 843, 369–411.
Stout, S.A., Uhler, A.D., Naymik, T.G., McCarthy, K.J., 1998. Wang, Z.D., Fingas, M., Shu, Y.Y., Sigouin, L., Landriault, M.,
Environmental forensics: unravelling site liability. Environ. Sci. Lambert, P., 1999b. Quantitative characterization of PAHs in burn
Technol. 32, 260A–264. residue and soot samples and differentiation of pyrogenic PAHs
Stout, S.A., Uhler, A.D., McCarthy, K.J., 2001. A strategy and from petrogenic PAHs––the 1994 Mobile burn study. Environ. Sci.
methodology for defensibly correlating spilled oil to source Technol. 33, 3100–3109.
candidates. Environ. Forens. 2, 87–98. Wang, Z.D., Fingas, M., Landriault, M., Sigouin, L., Grenon, S., Zhang,
Stout, S.A., Uhler, A.D., McCarthy, K.J., Emsbo-Mattingly, S., 2002. D., 1999c. Source identification of an unknown spilled oil from
Chemical fingerprinting of hydrocarbons. In: Murphy, B.L., Quebec (1998) by unique biomarker and diagnostic ratios of ‘‘source-
Morrison, R.D. (Eds.), Introduction to Environmental Forensics. specific marker compounds’’. Environ. Technol. 20, 851–862.
Academic Press, London, pp. 137–260. Chapter 6. Wang, Z.D., Fingas, M., Owens, E.H., Sigouin, L., Brown, C.E., 2001.
Theobald, N., Rave, A., Jerzycki-Brandes, K., 1995. Input of Long-term fate and persistence of the spilled Metula oil in a marine
hydrocarbons into the North Sea by the river Elbe. Fresen. J. salt marsh environment: degradation of petroleum biomarkers.
Anal. Chem. 353, 83–87. J. Chromatogr. 926, 190–275.
Venkatesan, M.I., 1988. Occurrence and possible sources of perylene in Wang, Z.D., Fingas, M., Sigouin. L., 2002. Using multiple criteria for
marine sediments––a review. Mar. Chem. 25, 1–27. fingerprinting unknown oil samples having very similar chemical
Wang, Z.D., Fingas, M., Li, K., 1994a. Fractionation of ASMB oil, composition. In: Proceedings of the 25th Arctic and Marine Oil
identification and quantitation of aliphatic aromatic and biomar- Spill Program (AMOP) Technical Seminar, Environment Canada,
ker compounds by GC/FID and GC/MSD. J. Chromatogr. Sci. 32, Ottawa, pp. 639–660.
361–366 (Part I) and 367–382 (Part II). Whittaker, M., Pollard, S.J.T., Fallick, T.E., 1995. Characterization of
Wang, Z.D., Fingas, M., Sergy, G., 1994b. Study of 22-year-old Arrow refractory wastes at heavy oil-contaminated sites: a review of
oil samples using biomarker compounds by GC/MS. Environ. Sci. conventional and novel analytical methods. Environ. Technol. 16,
Technol. 28 (9), 1733–1746. 1009–1033.
Wang, Z.D., Fingas, M., Sergy, G., 1995a. Chemical characterization Whittaker, M., Pollard, S.J.T., Fallick, A.E., Preston, T., 1996.
of crude oil residues from an Arctic beach by GC/MS and GC/FID. Characterisation of refractory wastes at hydrocarbon-contami-
Environ. Sci. Technol. 29 (10), 2622–2631. nated sites 11: screening of reference oils by stable carbon isotope
Wang, Z.D., Fingas, M., 1995b. Use of methyldibenzothiophenes as fingerprinting. Environ. Pollut. 94, 195–203.
markers for differentiation and source identification of crude and Williams, J.A., Bjorøy, M., Dolcater, D.L., Winters, J.C., 1986.
weathered oils. Environ. Sci. Technol. 29 (11), 2841–2849. Biodegradation in South Texas Eocene oils: effects on aromatics
Wang, Z.D., Blenkinsopp, S., Fingas, M., Sergy, G., Landriault, M., and biomarkers. Org. Geochem. 10, 451–461.
Sigouin, L., Foght, J., Semple, K., Westlake, D.W.S., 1997a. Xu, H., Lesage, S., 1992. Separation of vanadyl and nickel petropor-
Chemical composition changes and biodegradation potentials of phyrins on an aminopropyl column by HPLC. J. Chromatogr. 607,
nine Alaska oil under freshwater incubation conditions. Symp. Am. 139–144.
Chem. Soc. 43 (3), 828–835 (preprints). Xu, H., Lesage, S., Brown, S., 1994. Petroporphyrins as chemical
Wang, Z.D., Fingas, M., Landriault, M., Sigouin, L., Feng, Y., indicators of soil contamination by crude oil. Chemosphere 28,
Mullin, J., 1997b. Using systematic and comparative analytical 1599–1609.
data to identify the source of unknown oil on contaminated birds. Zakaria, M.P., Horinouchi, A., Tsutsumi, S., Takada, H., Tanabe, S.,
J. Chromatogr. 775, 251–265. Ismail, A., 2000. Oil pollution in the Straits of Malacca, Malaysia:
Wang, Z.D., Fingas, M., Blenkinsopp, S., Sergy, G., Landriault, M., Application of molecular markers for source identification. Envi-
Sigouin, L., 1998a. Study of the 25-year-old Nipisi oil spill: ron. Sci. Technol. 34, 1189–1196.
persistence of oil residues and comparisons between surface and Zeng, E.Y., Vista, C.L., 1997. Organic pollutants in the coastal
subsurface sediments. Environ. Sci. Technol. 32 (15), 2222–2232. environment off San Diego, California.1. Source identification and
Wang, Z.D., Fingas, M., Landriault, M., Sigouin, L., Castle, B., assessment by compositional indices of polycyclic aromatic hydro-
Hostetter, D., Zhang, D., Spencer, B., 1998b. Identification and carbons. Environ. Toxicol. Chem. 16, 179–188.

Potrebbero piacerti anche