Open access peer-reviewed chapter

Polycyclic Aromatic Hydrocarbons (PAHs) and Their Influence to Some Aquatic Species

Written By

Ayoub Baali and Ahmed Yahyaoui

Submitted: 06 December 2018 Reviewed: 04 April 2019 Published: 27 September 2019

DOI: 10.5772/intechopen.86213

From the Edited Volume

Biochemical Toxicology - Heavy Metals and Nanomaterials

Edited by Muharrem Ince, Olcay Kaplan Ince and Gabrijel Ondrasek

Chapter metrics overview

1,596 Chapter Downloads

View Full Metrics

Abstract

Polycyclic aromatic hydrocarbons (PAHs) are ubiquitous environmental pollutants generated primarily during the incomplete combustion of organic materials (e.g., coal, oil, petrol, and wood). Many PAHs have toxic, mutagenic, and/or carcinogenic functions. PAHs are highly lipid soluble which lead to a fast absorption by the gastrointestinal tract of marine mammals. They are immediately distributed in a vast variety of tissues with a notable tendency for localization in body fat. Metabolism of PAHs is obtained via the cytochrome P450-mediated mixed function oxidase system with oxidation or hydroxylation as the first step. PAHs are environmental contaminants that pose significant risk to health of fish. The effect of PAHs on fish is a topic of rising attention in a lot of countries. Different studies using the bile metabolites separated by high-performance liquid chromatography with fluorescence detection were presented. The aim is to compare the levels of PAH metabolites in fish from different areas and fish species. The major metabolite present in all fish was 1-hydroxypyrene. The data confirm the importance of 1-hydroxypyrene as the key PAH metabolite in fish bile and suggest that the European eel is an ideal species for monitoring PAHs.

Keywords

  • PAHs
  • organic pollutants
  • metabolism
  • fish
  • 1-hydroxypyrene
  • European eel

1. Introduction

Aquatic ecosystems are susceptible to receiving and accumulating contaminants [1]. In particular, polycyclic aromatic hydrocarbons (PAHs) have been identified as general causes of the deterioration of aquatic ecosystems in recent decades [2].

Polycyclic aromatic hydrocarbons (PAHs) are ubiquitous and persistent environmental contaminants found in sediments and associated waters of urbanized estuaries and coastal areas [3, 4, 5]. They are a class of compounds found in crude oil and are everywhere in the aquatic ecosystem [6, 7, 8, 9, 10, 11, 12]. PAHs are the most toxic pollutants of crude oil and are remembered by the United States Environmental Protection Agency (EPA) as priority toxic components because of its persistence in the environment and are toxic to fishes [13, 14]; thus, PAHs are of special interest following oil spills and in environmental control. They come from natural and anthropogenic sources. The latter can be associated to pyrolysis and incomplete combustion of organic element [15]. Wastewater, atmospheric deposition, and petroleum spillage are some of the most important PAH sources. PAHs and their intermediate degradation products have the potential to generate toxic or mutagenic effects in fish [16, 17, 18] and humans [19]. PAH metabolites in the bile fluid are generally accepted as measures for PAH exposure in fish because of the rapid metabolism of PAH in most vertebrates [3]. Therefore, PAH metabolites in fish are recommended as monitoring parameters in European seas [20, 21].

In this chapter, we briefly review the origin, toxicity, and transformation of PAHs in the aquatic environment, highlighting their efficient metabolism in fish. We also review the presence of PAHs on fish bile and the works reported on that.

Advertisement

2. Organic contamination by polycyclic aromatic hydrocarbons (PAH)

2.1 PAH origin

PAHs are mainly formed during the incomplete combustion of organic matter and during the slow maturation of organic matter accumulated in deep sedimentary environments. These two origins present distinct formation mechanisms that are realized with different kinetics and induce variable molecular distributions (related to stability) [22].

2.1.1 Pyrolytic origin

Pyrolytic PAHs are generated by processes of incomplete combustion of organic matter at high temperatures. The mechanisms involved in their formation involve the production of free radicals by pyrolysis at high temperature (≥500°C) of the fossil material (oils, fuel oil, organic matter, etc.) under oxygen-deficient conditions. PAHs of pyrolytic origin come from the combustion of automotive fuel, domestic combustion (coal, wood, etc.), industrial production (steelworks, aluminum smelters, etc.), and energy production (power stations operating on oil or coal) or incinerators [23].

2.1.2 Petrogenic origin

The process of diagenesis can give rise to petroleum and other fossil fuels containing the so-called petrogenic PAHs. These PAHs are formed at low temperatures (150°C) over long periods of time. They result from exposure of organic matter to adequate conditions of temperature and pressure. The proportion of PAHs in oils varies according to their origin and level of refinement. In general, petrogenic PAH mixtures are marked by a predominance of low molecular weight PAHs, three cycles or less, and substituted PAHs [22].

PAHs represent between 20 and 40% by weight of crude oils which are mainly composed of saturated hydrocarbons. However, they are less than a few percent of the composition of refined gasoline (<0.5% by mass) or kerosene [24].

2.1.3 Biological origin

PAHs can also be formed by microorganisms from biogenic precursors such as diterpenes and triterpenes, steroids, pigments, or quinones in sediments or recent soils [25, 26]; these precursors can come from terrestrial or aquatic biological tissues (plants, animals, bacteria, macro- and microalgae) [27, 28].

2.2 PAH toxicity

The toxicity of several PAHs is a phenomenon that is well-known. Research has been conducted by several environmental groups such as the US Environmental Protection Agency-Toxic Substances Control Act (US EPA-TSCA) and the International Agency for Research on Cancer (IARC). The toxicity of PAHs to aquatic species is affected by metabolism and photooxidation. They are generally harmful in the presence of ultraviolet light. PAHs have moderate to high acute toxicity to aquatic organisms and birds. Mammals can absorb PAHs by various routes, e.g., dermal contact, inhalation, and ingestion [14, 29, 30]. The concentrations of PAHs found in fish are expected to be much higher than in the environment from which they were taken because of their bioaccumulation. Withal, metabolism of PAHs is sufficient to prevent biomagnifications [31, 32].

Teleost fish have an immense capacity to metabolize PAHs because of the enzymes cytochrome P450 in their tissues that oxidatively biotransform PAHs to hydroxylated metabolites [33].

The half-life times of PAHs in various biological tissues (bivalves, crustaceans, and fish) are of the order of a few days to 10 days and are about five times higher for heavy PAHs relative to lower PAHs.

The environmental matrices are moreover complex, containing numerous endogenous or exogenous, mineral or organic molecules between which interactions can take place. Synergistic toxic effects have been observed in particular between metals and PAH quinones [34].

Indeed, the carcinogenic nature of some of these molecules alone or in mixtures is proven. Twelve of these are classified by the International Agency for Research on Cancer (IARC) as probably carcinogenic to humans [35]. After contamination by these substances, they are biotransformed in the liver into different (poly)hydroxy-PAHs.

The risks of PAHs to fish and other aquatic organisms in natural systems are highly uncertain due to the occurrence of complex, incompletely characterized mixtures of these chemicals, large spatial heterogeneity in exposure concentrations, incomplete understanding of the importance of UV-activated PAH toxicity, the biological and physical controls on fish exposure to UV light, and the bioaccumulation of PAHs. These uncertainties are especially great for early-life-stage fish, which might be particularly susceptible to UV-activated because of their small size, lack of protective pigmentation and gill coverings, and ready accumulation of PAHs.

Advertisement

3. Transformations of PAHs in marine ecosystem

3.1 Physical and chemical degradation

Sediments contaminated with PAHs pose a real threat to all living organisms, even those that feed on the benthic prey.

PAHs with a low molecular weight can be found in all environmental matrices, since higher molecular weight compounds are more deeply associated (physically and chemically) with sediments/soils and particles than the other abiotic sample types. PAHs in air can be modified via chemical oxidation and photochemical processes [8], whereas in sediments/soils and the uppermost portion of the water column, degradation of PAHs, particularly lower molecular weight PAHs, occurs via photooxidation [6, 36]. In addition to parent PAHs, oxygenated PAH metabolites formed during these degradation processes can persist associating with sediments up to 6 months after initial addition to the water column and thus can endure in the environment for extended periods of time [37]. In water samples and sediment, some microorganisms (e.g., fungi, bacteria) have been demonstrated to mineralize PAHs under aerobic conditions, particularly those compounds that contain two- and three-fused rings (e.g., fluorene, naphthalene), to their basic elements or to biodegrade these compounds to more polar degradation products [3, 8]. More information on PAH microbial degradation pathways and identification of degradation can be found in Cerniglia and Heitkamp [38], Juhasz and Naidu [39], and Bamforth and Singleton [40]. Part of research studies have proved that pyrene can be mineralized by certain strains of bacteria (e.g., Mycobacterium) under optimum growing conditions in the laboratory, but it is uncertain if this occurs in the natural environment [41, 42]. In contrast, other higher molecular weight PAHs (e.g., five- and six-ring compounds) are not readily degraded by microbes and thus are more likely to accumulate in these environmental media (particularly in fine-grained sediments with high organic carbon content) [3, 39]. Under anoxic conditions, PAHs persist in sediments, particularly in organic sediments [42].

3.2 Biotransformation in the aquatic food web

Pyrolytic PAHs are the most common in aquatic environments. After the emission of pyrolytic PAHs into the atmosphere, the molecules fall back and settle on the surface of the water or soil [23]. Under the action of soil leaching, PAHs are transported to water bodies. These hydrophobic molecules then associate with other particles of the column of water and accumulate in the sediment [23].

In aquatic organisms, exposure to PAHs can occur through dermal exposure, respiration, or consumption of contaminated prey (e.g., annelids, crustaceans) or sediment [43]. Biotransformation of PAHs in aquatic organisms occurs to varying degrees depending on a number of factors, including the rate of uptake, metabolic capability, physical condition, feeding strategy, and age [43, 44]. Invertebrates are capable of PAH uptake from their environment and have been shown to have varying levels of PAH-metabolizing capability [44]. Mollusks generally have lower PAH-metabolizing capability than certain species of polychaetes, crustaceans, and fish [3].

PAH metabolism in fish is mainly conducted by inducible enzymes of the cytochrome P450 family, in particular CYP1A. These enzymes are localized in the membranes of the smooth endoplasmic reticulum, located mainly in the liver, but are also present in other organs. They are expressed and functional from the earliest stages of fish development. These enzymes catalyze the addition of an oxygen atom to the PAH molecule through an NADPH-dependent reaction. CYP1A protein is induced during exposure of the body to PAHs. CYP1A induction is rapid, and activity levels are often increased by a factor of 100 a few hours after exposure. These highly polar conjugated metabolites are then excreted into urine or the bile for rapid rejection over the gastrointestinal tract [14, 43, 45, 46]. Concerning the result of this rapid metabolism in fish, concentrations of parent PAHs are insignificant in muscle and other tissues. Thus, to determinate a recent exposure to PAHs in teleost fish, bile and urine are used, with the preference of bile because of its easy sampling. Otherwise, differences in the metabolism of benzo[a]pyrene (BaP), including differences in the types and proportions of metabolites formed, have been shown between two fish of freshwater [47, 48]. These differences could contribute to variations in susceptibility to these carcinogenic compounds among fish species. Differences in glutathione-S-transferases may also help explain differential susceptibility to chemically induced carcinogenesis among fish species [46]. A number of analytical methods have been developed to measure PAH metabolites in fish bile and are reviewed in [14]. Oil spills, including the Deepwater Horizon (DWH) spill in the Gulf of Mexico in 2010, show the necessity of the need for additional methods to determine PAH exposure in seafood or protected species (e.g., marine mammals) where species cannot be lethally collected. For example, a rapid, sensitive HPLC-fluorescence method was developed by the US Food and Drug Administration [49] during the DWH spill and was used by federal and Gulf state analytical laboratories as part of the seafood safety response [50]. The development of new analytical methods like those can provide important information on PAH exposure in aquatic organisms.

Advertisement

4. Effects of PAHs on aquatic species

PAHs are an important factor of contamination in the environment but also a risk to human health. In fact, the dangers related to PAHs vary according to their toxicity on the one hand and the many sources of exposure on the other. It has also been proven that carcinogenic and mutagenic effects related to a single compound of PAHs were found. A large number of effects have been identified [51]. In fact, genotoxicity and tumorigenesis observed in fish are linked to the presence of metabolites. Beyond genotoxicity, there are many other effects observed, for example, in behavior, reproduction, and growth.

Benzo[a]pyrene, for example, which is highly studied, leads to a decrease in weight [52] and growth [53], an increase in the gonado-somatic index (GSI) in Japanese medaka (Oryzias latipes) [52], DNA breaks in oysters (Crassostrea gigas) [54], and DNA adducts in zebrafish and on human liver cells (HepG2) [55]. Teratogenic effects in particular on the heart of sardine (Clupea pallasii) [56] and zebrafish [57, 58] have been observed as well as anemia in scorpion fish (Sebastes schlegelii) [59]. Benzo[a]pyrene affects the reproduction of isopods (Oniscus asellus and Asellus porcellio scaber) [53] and accumulates in oocytes in catfish (Ictalurus punctatus) [60]; it disrupts the expression of the aromatase (enzyme necessary for the conversion of androgen such as estrogen isosterone) in female mummichog (Fundulus heteroclitus) [61] and inhibits the synthesis of testosterone and estradiol in flounder (Platichthys flesus L.) [62].

The toxicity of a compound can be enhanced or reduced by endogenous and exogenous factors. For example, in fish, hypoxia [63], geographical origin and fish life history [64], and/or the various PAH compounds [65] can cause variations. The penetration time in the fish embryo and the depuration time can vary considerably. In addition, the effects produced by these molecules, tested individually, do not necessarily follow a dose-effect relationship [52].

4.1 Effects of PAHs on the survival

Survival is a commonly used variable. This variable has made it possible to develop standardized tests in toxicology such as LD50 (lethal dose) or LC50 (lethal concentration) calculations. Although protected in their chorions, fish eggs and then larvae are particularly exposed because, in most cases, they are unable to flee contaminated areas during their early-life stages [66]. It is during development, during the establishment of all organs and systems, that contaminants can pass this barrier. They can affect the development and have long-term consequences. Exposure to PAHs can lead to decreased survival in aquatic organisms during acute exposures. For example, there has been a decrease in survival after early exposure of salmon (Oncorhynchus gorbuscha) to dissolve PAHs [67].

In the case of chronic exposure in the early stages, similar effects can be observed. Impairment in survival has been observed in salmon (Oncorhynchus gorbuscha) exposed to crude oil [68], in minnow (Pimephales promelas) exposed to contaminated sediment [69], in Chanos chanos and capelin (Mallotus villosus) exposed to dissolved PAHs (anthracene, B[a]P, pyrene and heavy fuel oil) [66, 70], and in shrimp (Palaemonetes pugio) exposed to the pyrene feed which also shows reduced survival [71]. In other studies, survival is not affected by PAHs. This is the case, for example, in terrestrial isopods, where oral administration does not have significant effects on survival [53]. It is also not affected after exposure to BaP in mummichog (Fundulus heteroclitus) [72]. PAHs can affect survival in some cases and not in others. This extreme variable may not be the most sensitive for all species or all types of exposure.

4.2 Malformations and growth

PAHs induce malformations during development. They lead to a decrease in skeletal mineralization in bass (Dicentrarchus labrax) [73] and craniofacial deformities in scorpion fish (Sebastiscus marmoratus) [74]. Jaw malformations [75] in this same fish as well as in zebrafish [58] were also observed. The number of edemas is also increased in scorpion fish (Sebastes schlegelii), salmon (Oncorhynchus gorbuscha) [67, 76], and medaka (Oryzias latipes) [77], as well as the occurrence of hemorrhages in trout (Oncorhynchus mykiss) [78]. The impact of PAHs on growth is frequently reflected in a reduction in size and/or weight [67, 79, 80]. This reduction in growth is observed regardless of the mode of administration of PAHs, the concentrations used, and the duration of exposure [59, 70, 81]. Weight reduction is often proportional to contamination [81]. Unfortunately, these are not the only visible damage. A decrease in lipid reserves may be observed and result in a decrease in energy reserves [10, 79].

4.3 Metabolism and osmoregulation

In a PAH study, fish were exposed to the soluble fraction of a crude oil mixture. Structural lesions and morphological differences are noted on the gill [82]. These differences would be related to a metabolism aimed at reducing contact with the pollutant, which would reduce the gill surface and oxygen supply. A reduction in oxygen uptake could compromise the fish metabolism [82].

Osmoregulation problems following exposure of Sebastiscus marmoratus to dissolve PAHs have been observed [75]. PAHs would inhibit Na+/K+ activity in a dose-dependent manner and play a role in osmoregulation.

4.4 Effect on behavior

The behavioral response of an animal following exposure to stress and/or contaminant (s) is increasingly studied [83, 84, 85, 86, 87].

The behavior makes it possible to discriminate a large number of integrating variables from the changes induced by PAHs. Swimming activity can be assessed as well as other aspects such as lethargy, anxiety, social communication, eating behavior, flight response, learning, or reproductive behavior.

A reduced swimming activity in seabream (Sparus aurata) was observed following a 4 days of exposure to dissolved PAHs [88, 89]: phenanthrene, fluorine, and pyrene [89]. An increase in lethargy and a reduction in the number of surface surges have also been observed following exposure to dissolved PAHs in this species [88, 89].

These variables can also be used to evaluate the neurotoxic effects of a contaminant. The reduction of social interactions following exposure to phenanthrene [89] is proven.

The escape response, in the presence of fluoranthene, has been demonstrated in control fish [90]. The fish are placed in a double flow aquarium. A flow of control water and a flow of water containing fluoranthene are present. Fish that have never been exposed are fleeing fluoranthene. On the other hand, fish that have been previously exposed to a high dose of fluoranthene no longer leak the molecule [90].

Learning and exploration abilities are diminished after exposure to PAHs. For example, discrimination of a familiar object is altered in mice exposed to BaP. In the same vein, dietary exposure of the mother to a mixture containing the 16 priority PAHs of the USEPA leads to behavioral alteration in the next generation, especially in a new environment. The reproductive behavior can also be disturbed. The ability of a male to find a female can be altered, as is the case in amphipods, for example [91].

4.5 Effect on reproduction

PAHs are lipophilic molecules that are transported and found in the ovaries via vitellogenin and/or lipovitellin [60]. They can also result in inhibition of vitellogenin synthesis, as has already been shown in trout after exposure to β-naphthoflavone [92]. This exposure compromises the maturation of the ovaries and causes an increase in apoptosis in gonadal cells [91]. These pollutants lead, for example, to reproductive inhibition in shrimp exposed to pyrene [71]. A decrease in fecundity, number of breeding cycles, and larval survival is observed in different fish species. In mussels, gametes are deformed and are present in small numbers [91].

Females are not the only ones to be affected. Male sperm quality can also be altered after exposure to benzo[b]fluoranthene, as is the case in mice exposed via breast milk [59]. Sperm quality is reduced, and there is also an increase in testicular apoptosis.

Advertisement

5. PAHs in eels

High-performance liquid chromatography (HPLC) is generally used for the determination of PAH metabolites in considerable fish species [93, 94, 95, 96, 97, 98] and has been covered by an intercalibration exercise [22].

Although bile metabolites have been measured in many species of fish [13], those selected for biomonitoring programs tend to be common, long-lived species at the top of the food chain, with relatively sedentary life styles and benthic habits [99]. Consequently, the common eel (Anguilla anguilla) has been used in studies of PAH contamination [100, 101]. Pleuronectid flatfish are also well-suited to biomonitoring, and the European flounder (Pleuronectes flesus), an abundant flatfish in most European estuaries, has been frequently chosen to assess PAH contamination [102, 103].

The study conducted by Ruddock et al. [104] in the Severn Estuary showed that from the six metabolites of polycyclic aromatic hydrocarbons (PAHs) identified and quantified from the bile of Anguilla anguilla, Pleuronectes flesus, and Conger conger collected during 1997, the main metabolite present in all fish was 1-hydroxypyrene with lower proportions of 1-hydroxychrysene and 1-hydroxyphenanthrene and small concentration of three benzo[a]pyrene derivatives. The results approve the importance of 1-hydroxypyrene as the important PAH metabolite in fish bile and suggest that the A. anguilla is an excellent species for monitoring PAHs in estuarine ecosystems. 1-Hydroxypyrene is invariably the major metabolite present in the bile of fish exposed to PAH-contaminated sediments [105], and this was confirmed by the results of this work for fish in the Severn Estuary. Pyrene is produced by many petrogenic and pyrolytic processes [43] and has been detected in significant number in sewage outfalls [106]. It is regarded the best general indicator of PAH exposure in fish [100, 107]. The contribution of 1-OH-Phen to the total metabolites detected ranged from approximately 2% in flounders and conger eels to 8% in common eels. Phenanthrenes are released to the atmosphere during the combustion of fossil fuels, particularly coal, oil, and its refined products [43]. Like all PAHs with two to four benzene rings, phenanthrenes can remain suspended in airborne particles for long periods [108]. Compared to the other PAHs detected, BaP has a very low solubility in water and low bioavailability, but metabolites of BaP are especially important because of their potent mutagenic and carcinogenic properties [109, 110, 111].

A recent study conducted by Baali et al. [2] on bile metabolites of PAHs in 18 European eels (Anguilla anguilla), 7 moray (Muraenidae), and 28 conger eels (Conger conger) from Moroccan waters (Moulay Bousselham lagoon and Boujdour) shows the presence of two polycyclic aromatic hydrocarbon (PAH) metabolites, 1-hydroxypyrene (1-OH-Pyr) and 1-hydroxyphenanthrene (1-OH-Phen). The high-performance liquid chromatography with fluorescence detection method was used to separate the bile metabolites.

The goals of the present study were to compare the levels of PAH metabolites in eels from the lagoon and sea and also to compare levels of PAH metabolites between the different eels.

In this study the PAH metabolites (1-OH-Pyr and 1-OH-Phen) were detected in all species. The results of this investigation show that the concentration of 1-OH-Pyr was high for Anguilla anguilla than the other species (Figure 1). The conger eels represent the species with the lower concentration of 1-OH-Pyr. This result reflects the low degree of contamination in Boujdour coast (Figure 1). Thus, the presence of high concentration of 1-OH-Pyr and 1-OH-Phen in the bile of the European eels and morays reflects the high degree of contamination in the lagoon which is due to the anthropogenic activity in Moulay Bousselham lagoon. From the comparison between the contamination of the European eels and morays belonging to Moulay Bousselham lagoon, the results show that the first species present a higher concentration of PAH metabolites than the second one (Figure 1). This conclusion confirms that the Anguilla anguilla is more suitable species for monitoring PAH contamination. The European eels spend most of their life in muddy sediment which usually present a high PAH concentration levels. The pollutants in sediment are easily accumulated [112, 113, 114]. Accordingly, it is recognized that sediment contamination has a particular interest with regard to aquatic ecosystem quality. Sediment is an important source of pollutants and the factor with the high impact on the deterioration of the water quality. Although the feeding habit of the European eels may result in higher exposure to PAHs whence the high concentration of 1-OH-Pyr and 1-OH-Phen in the bile of this species [115], the accumulation of PAHs from the surrounding water is considered more efficient than impacted food [116]. The level of PAH metabolites in fish bile varies according to the area. The results show that Boujdour Sea is not a polluted site [117]. Moulay Bousselham lagoon is a semi-closed area; the concentration of pollutants in this site is higher than Boujdour Sea because of its lower water circulation. In the lagoon PAHs are easily accumulated than that in the sea [112]. Our results confirm that 1-OH-Pyr is the major metabolite present in fish bile [104, 105] and the best indicator of PAH exposure in fish [100, 107]. It was found that 1-OH-Pyr is the dominant compound in eel bile [118, 119, 120]. The results show that the levels of PAHs in Morocco are lower than those obtained in the other regions. As a conclusion of this study, the possible health risk of PAH contamination in Boujdour coast and Moulay Bousselham lagoon might be low compared to the other European sites.

Figure 1.

Bile metabolite 1-hydroxypyrene (a) and 1-hydroxyphenanthrene (b) concentrations detected in European eels (Anguilla anguilla) collected from different areas and eels from Morocco (conger, moray, and European eel) as mean (triangles) and range (panels).

The concentration of 1-OH-Pyr varies significantly with length (p < 0.05) for each species. The results obtained in this study [2] show that the concentration of PAH metabolites does not always increase with the size; there are obviously factors which can affect the exposure of this pollutant such as species differences, age, sex, maturity, and diet.

Advertisement

6. Conclusion

PAHs are originally organic compounds that are created from the partial combustion of organic elements or pyrolysis of organic material. These compounds are associated to the treatment of wood, oil, coal, and gas in order to produce the energy.

PAHs are transferred in the air in gas or particle aspect, and they are accumulated by wet and dry deposition. The transported elements play important role in the chemistry of the atmosphere. These particles also have significant impact in human health, because many PAHs are classified as probable human carcinogens.

The other faculty of PAHs is the capacity of degrading microorganism such as bacteria, fungi, and algae. It concerns the failure of organic compounds through biotransformation into less complex metabolites and through mineralization into inorganic minerals.

In this chapter, many effects on the biology of species following exposure to PAHs have been demonstrated. At the end of these organic studies on fish, it has been shown that the PAH biliary metabolites studied have the potential to describe the state of exposure of fish to organic pollutants (PAHs).

The study of a possible contamination of eels from different countries shows that 1-hydroxypyrene (1-OH-Pyr) is the dominant pollutant present in fish bile and is the best general indicator of PAH contamination.

Of the different eels investigated, European eels (Anguilla anguilla) contained the highest metabolite concentrations. This species looks like the most suitable for monitoring PAH contamination in the environment.

Using the studies conducted by several authors, we found that the rivers and lagoon contain PAH concentrations much higher than the coastal waters. These results appear normal in view that there is low water exchange in the rivers and lagoon ecosystems.

Finally we conclude that the quantification and identification of the metabolites in fish bile can give a rapid indication on the level of PAH contamination.

Advertisement

Conflict of interest

The authors declare that they have no competing interests.

References

  1. 1. Ahmad I, Pacheco M, Santos A. Anguilla anguilla L. oxidative stress biomarkers: An in situ study of freshwater wetland ecosystem (Pateira de Fermentelos, Portugal). Chemosphere. 2006;65:952-962
  2. 2. Baali A, Kammann U, Hanel R, El Qoraychy I, Yahyaoui A. Bile metabolites of polycyclic aromatic hydrocarbons (PAHs) in three species of fish from Morocco. Environmental Sciences Europe. 2016;28:25. DOI: 10.1186/s12302-016-0093-6
  3. 3. Meador J, Stein J, Reichert W, Varanasi U. Bioaccumulation of polycyclic aromatic hydrocarbons by marine organisms. Reviews of Environmental Contamination and Toxicology. 1995;143:79-165
  4. 4. Woodhead RJ, Law RJ, Matthiessen P. Polycyclic aromatic hydrocarbons in surface sediments around England and Wales, and their possible biological significance. Marine Pollution Bulletin. 1999;38:773-790
  5. 5. Blahová J, Havelková M, Kružíková K, Hilscherová K, Halouzka R. Assessment of contamination of the Svitava and Svratka rivers in the Czech Republic using selected biochemical markers. Environmental Toxicology and Chemistry. 2010;29(3):541-549
  6. 6. McElroy AE, Farrington JW, Teal JM. Bioavailability of PAH in the aquatic environment. In: Varanasi U, editor. Metabolism of Polycyclic Aromatic Hydrocarbons in the Aquatic Environment. Boca Raton, FL: CRC Press; 1989. pp. 1-40
  7. 7. Ramesh A, Walker SA, Hood DB, Guillén MD, Schneider K, Weyand EH. Bioavailability and risk assessment of orally ingested polycyclic aromatic hydrocarbons. International Journal of Toxicology. 2004;23(5):301-333
  8. 8. Lima ALC, Farrington JW, Reddy CM. Combustion-derived polycyclic aromatic hydrocarbons in the environment—A review. Environmental Forensics. 2005;6:109-131
  9. 9. Hylland K. Polycyclic aromatic hydrocarbon (PAH) ecotoxicology in marine ecosystems. Journal of Toxicology and Environmental Health, Part A. 2006;69:109-123
  10. 10. Meador JP, Sommers FC, Ylitalo GM, Sloan CA. Altered growth and related physiological responses in juvenile Chinook salmon (Oncorhynchus tshawytscha) from dietary exposure to polycyclic aromatic hydrocarbons (PAHs). Canadian Journal of Fisheries and Aquatic Sciences. 2006;63:2364-2376
  11. 11. Meador JP, Buzitis J, Bravo CF. Using fluorescent aromatic compounds in bile from juvenile salmonids to predict exposure to polycyclic aromatic hydrocarbons. Environmental Toxicology and Chemistry. 2008;27(4):845-853
  12. 12. Yanagida GK, Anulacion BF, Bolton JL, Boyd D, Lomax DP, Olson OP, et al. Polycyclic aromatic hydrocarbons and risk to threatened and endangered Chinook salmon in the lower Columbia River estuary. Archives of Environmental Contamination and Toxicology. 2012;62(2):282-295
  13. 13. Tuvikene A. Responses of fish to aromatic hydrocarbons. Annales Zoologici Fennici. 1995;32:295-309
  14. 14. Beyer J, Jonsson G, Porte C, Krahn MM, Ariese F. Analytical methods for determining metabolites of polycyclic aromatic hydrocarbon (PAH) pollutants in fish bile: A review. Environmental Toxicology and Pharmacology. 2010;30:224-244
  15. 15. Eisler R. Eisler’s Encyclopedia of Environmentally Hazardous Priority Chemicals. 1st ed. Amsterdam: Elsevier; 2007. 986 p
  16. 16. Brinkmann M, Hudjetz S, Cofalla C, Roger S, Kammann U, Zhang X, et al. A combined hydraulic and toxicological approach to assess re-suspended sediments during simulated flood events. Part I-multiple biomarkers in rainbow trout. Journal of Soils and Sediments. 2010;10:1347-1361
  17. 17. Brinkmann M, Eichbaum K, Kammann U, Hudjetz S, Cofalla C, Buchinger S, et al. Physiologically-based toxicokinetic models help identifying the key factors affecting contaminant uptake during flood events. Aquatic Toxicology. 2014;152:38-42
  18. 18. Monteiro PRR, Reis-Henriques MA, Coimbra J. Plasma steroid levels in female flounder (Platichthys flesus) after chronic dietary exposure to single polycyclic aromatic hydrocarbons. Marine Environmental Research. 2000a;49:453-467
  19. 19. Chen SC, Liao CM. Health risk assessment on human exposed to environmental polycyclic aromatic hydrocarbons pollution sources. Science of the Total Environment. 2006;366:112-123
  20. 20. OSPAR Commission. Co-ordinated environmental monitoring programme (CEMP). 2008. Available from: http://www.ospar.org/content/content.asp?menu=00900301400000 [Accessed: 27 December 2015]
  21. 21. HELCOM. Core indicators final report of the HELCOM CORESET project. Baltic Sea Environment Proceedings No. 136. 2013. 71 p. http://www.helcom.fi/Lists/Publications/BSEP136.pdf#search=core%20set [Accessed: 26 December 2015]
  22. 22. Wariaghly F. Etude écotoxicologique et parasitologique chez l’Anguille (Anguilla anguilla L.) dans les estuaires marocains : Sebou et Loukkos (Atlantique) [thesis]. Rabat: Université Mohammed V-Agdal, Faculté des Sciences; 2013
  23. 23. Abdel-shafy HI, Mansour MSM. A review on polycyclic aromatic hydrocarbons: Source, environmental impact, effect on human health and remediation. Egyptian Journal of Petroleum. 2016;25:107-123
  24. 24. Cole GM. Assessment and Remediation of Petroleum Contaminated Sites. 1st ed. Lewis Publishers; 1994. 384 p
  25. 25. Laflamme RE, Hites RA. The global distribution of polycyclic aromatic hydrocarbons in recent sediments. Geochimica et Cosmochimica Acta. 1978;42(3):289-303
  26. 26. Wakeham SG, Schaffner C, Giger W. Polycyclic aromatic hydrocarbons in recent lake sediments-I. Compounds having anthropogenic origins. Geochimica et Cosmochimica Acta. 1980;44:403-413
  27. 27. Ventakesan MI. Occurrence and possible sources of perylene in marine sediments—A review. Marine Chemistry. 1988;25(1):1-27
  28. 28. Wang X, Hong HS, Mu JL, Lin JQ , Wang SH. Polycyclic aromatic hydrocarbon (PAH) metabolites in marine fishes as a specific biomarker to indicate PAH pollution in the marine coastal environment. Journal of Environmental Science and Health, Part A. Toxic/Hazardous Substances and Environmental Engineering. 2008;43:219-226
  29. 29. Dong C, Chen C, Chen C. Determination of polycyclic aromatic hydrocarbons in industrial harbor sediments by GC-MS. International Journal of Environmental Research and Public Health. 2012;9:2175-2188
  30. 30. Veltman K, Huijbregts MAJ, Rye H, Hertwich EG. Including impacts of particulate emissions on marine ecosystems in life cycle assessment: The case of offshore oil and gas production. Integrated Environmental Assessment and Management. 2011;7:678-686
  31. 31. Inomata Y, Kajino M, Sato K, Ohara T, Kurokawa J, Ueda H, et al. Emission and atmospheric transport of particulate PAHs in Northeast Asia. Environmental Science & Technology. 2012;46(9):4941-4949. DOI: 10.1021/es300391w
  32. 32. Tudoran MA, Putz MV. Polycyclic aromatic hydrocarbons: From in cerebro to in silico eco-toxicity fate. Chemical Bulletin of Politehnica University of Timisoara. 2012;57(71):50-53
  33. 33. Choi MS. Effects of tributyltin (TBT) on the expression of cytochrome P4501A, aryl hydrocarbon receptor and vitellogenin genes[Master’s thesis]. Sunmoon University; 2012
  34. 34. El Morhit M, Fekhaoui M, Elie P, Girard P, Yahyaoui A, El Abidi A, et al. Heavy metals in sediment, water and the European glass eel Anguilla anguilla (Osteichthyes: Anguillidae), from Loukkos River estuary (Morocco, eastern Atlantic). Cybium. 2009;33:219-228
  35. 35. AESN (Agence de l’Eau Seine Normandie). Guide des profils de vulnérabilité des eaux de baignade. 2009. p. 84
  36. 36. McElroy AE, Farrington JW, Teal JM. Bioavailability of polycyclic aromatic hydrocarbons in the aquatic environment. In: Varanasi U, editor. Metabolism of Polycyclic Aromatic Hydrocarbons in the Aquatic Environment. Boca Raton, Florida: CRC Press, Inc.; 1989. pp. 2-33
  37. 37. Hinga KR, Pilson MEQ. Persistence of benz[a]anthracene degradation products in an enclosed marine ecosystem. Environmental Science & Technology. 1987;21:648-653
  38. 38. Cerniglia CE, Heitkamp MA. Microbial degradation of polycyclic aromatic hydrocarbons (PAH) in the aquatic environment. In: Varanasi U, editor. Metabolism of Polycyclic Aromatic Hydrocarbons in the Aquatic Environment. Boca Raton, FL: CRC Press; 1989. pp. 41-68
  39. 39. Juhasz AL, Naidu R. Extraction and recovery of organochlorine pesticides from fungal mycelia. Journal of Microbiological Methods. 2000;39:149-158
  40. 40. Bamforth SM, Singleton I. Bioremediation of polycyclic aromatic hydrocarbons: Current knowledge and future directions. Journal of Chemical Technology and Biotechnology. 2005;80(7):723-736
  41. 41. Heitkamp MA, Franklin W, Cerniglia CE. Microbial metabolism of polycyclic aromatic hydrocarbons: Isolation and characterization of a pyrene-degrading bacterium. Applied and Environmental Microbiology. 1988;54:2549-2555
  42. 42. Heitkamp MA, Cerniglia CE. Polycyclic aromatic hydrocarbon degradation by a Mycobacterium sp. in microcosms containing sediment and water from a pristine ecosystem. Applied and Environmental Microbiology. 1989;55:1968-1973
  43. 43. Varanasi U, Stein JE, Nishimoto M. Biotransformation and disposition of polycyclic aromatic hydrocarbons in fish. In: Varanasi U, editor. Metabolism of Polycyclic Aromatic Hydrocarbons in the Aquatic Environment. Boca Raton, FL: CRC Press; 1989. pp. 93-149
  44. 44. James MO. Biotransformation and disposition of PAH in aquatic invertebrates. In: Varanasi U, editor. Metabolism of Polycyclic Aromatic Hydrocarbons in the Aquatic Environment. Boca Raton, FL: CRC Press; 1989. pp. 70-88
  45. 45. Roubal WT, Collier TK, Malins DC. Accumulation and metabolism of carbon-14 labeled benzene, naphthalene, and anthracene by young coho salmon (Oncorhynchus kisutch). Archives of Environmental Contamination and Toxicology. 1977;5:513-529
  46. 46. Carls MG, Holland L, Larsen M, Collier TK, Scholz NL, Incardona JP. Fish embryos are damaged by dissolved PAHs, not oil particles. Aquatic Toxicology. 2008;88:121-127
  47. 47. Pangrekar J, Kandaswami C, Kole P, Kumar S, Sikka HC. Comparative metabolism of benzo(a)pyrene, chrysene and phenanthrene by brown bullhead liver microsomes. Marine Environmental Research. 1995;39:51-55
  48. 48. Sikka HC, Steward AR, Kandaswami C, Rutkowski JP, Zaleski J, Kumar S, et al. Metabolism of benzo(a)pyrene and persistence of DNA adducts in the brown bullhead (Ictalurus nebulosus). Comparative Biochemistry & Physiology. 1991;100C(1-2):25-28
  49. 49. Gratz S, Mohrhaus A, Gamble B, Gracie J, Jackson D, Roetting J, et al. Screen for the presence of polycyclic aromatic hydrocarbons in select seafoods using LC-fluorescence. Laboratory Information Bulletin. 2010;4475:1-39
  50. 50. Ylitalo GM, Krahn MM, Dickhoff WW, Stein JE, Walker CC, Lassitter CL, et al. Federal seafood safety response to the Deepwater Horizon oil spill. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(50):20274-20279
  51. 51. Hansen D, Di Toro D, McGrath J, Swartz R, Mount D, Burgess R, editors. Procedures for the Derivation of Equilibrium Partitioning Sediment Benchmarks (ESBs) for the Protection of Benthic Organisms: PAH Mixtures. Narragansett, RI: Duluth, MN: Newport, Or: USEPA; 2003. p. 175
  52. 52. Chikae M, Hatano Y, Ikeda R, Morita Y, Hasan Q , Tamiya E. Effects of bis(2-ethylhexyl) phthalate and benzo[a]pyrene on the embryos of Japanese medaka (Oryzias latipes). Environmental Toxicology and Pharmacology. 2004;16:141-145
  53. 53. Van Brummelen TC, Van Gestel CAM, Verweij RA. Long-term toxicity of five polycyclic aromatic hydrocarbons for the terrestrial isopods Oniscus Asellus and Porcellio Scaber. Environmental Toxicology and Chemistry. 1996;15:1199-1210
  54. 54. Wessel N, Rousseau S, Caisey X, Quiniou F, Akcha F. Investigating the relationship between embryotoxic and genotoxic effects of benzo[a]pyrene, 17[alpha]-ethinylestradiol and endosulfan on Crassostrea gigas embryos. Aquatic Toxicology. 2007;85:133-142
  55. 55. Miranda CL, Chung WG, Wang-Buhler JL, Musafia-Jeknic T, Baird WM, Buhler DR. Comparative in vitro metabolism of benzo[a]pyrene by recombinant zebrafish CYP1A and liver microsomes from [beta]-naphthoflavone-treated rainbow trout. Aquatic Toxicology. 2006;80:101-108
  56. 56. Incardona JP, Carls MG, Day HL, Sloan CA, Bolton JL, Collier TK, et al. Cardiac arrhythmia is the primary response of embryonic Pacific herring (Clupea pallasi) exposed to crude oil during weathering. Environmental Science & Technology. 2009;43:201-207
  57. 57. Hicken CE, Linbo TL, Baldwin DH, Willis ML, Myers MS, Holland L, et al. Sublethal exposure to crude oil during embryonic development alters cardiac morphology and reduces aerobic capacity in adult fish. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:7086-7090
  58. 58. Incardona JP, Collier TK, Scholz NL. Defects in cardiac function precede morphological abnormalities in fish embryos exposed to polycyclic aromatic hydrocarbons. Toxicology and Applied Pharmacology. 2004;196:191-205
  59. 59. Kim SG, Park DK, Jang SW, Lee JS, Kim SS, Chung MH. Effects of dietary benzo[a]pyrene on growth and hematological parameters in juvenile rockfish, Sebastes schlegeli (Hilgendorf). Bulletin of Environmental Contamination and Toxicology. 2008;81:470-474
  60. 60. Montverdi GH, Di Giulio RT. In vitro and in vivo association of 2,3,7,8-tetrachlorodibenzo-p-dioxin and benzo[a]pyrene with the yolk-precursor protein vitellogenin. Environmental Toxicology. 2000;19(10):2502-2511. DOI: 10.1002/etc.5620191016
  61. 61. Patel MR, Scheffler BE, Wang L, Willett KL. Effects of benzo(a)pyrene exposure on killifish (Fundulus heteroclitus) aromatase activities and mRNA. Aquatic Toxicology. 2006;77(3):267-278
  62. 62. Rocha Monteiro PR, Reis-Henriques MA, Coimbra J. Polycyclic aromatic hydrocarbons inhibit in vitro ovarian steroidogenesis in the flounder (Platichthys flesus L.). Aquat. Toxicology. 2000;48:549-559
  63. 63. Matson CW, Timme-Laragy AR, Di Giulio RT. Fluoranthene, but not benzo[a]pyrene, interacts with hypoxia resulting in pericardial effusion and lordosis in developing zebrafish. Chemosphere. 2008;74:149-154
  64. 64. Diekmann M, Nagel R. Different survival rates in zebrafish (Danio rerio) from different origins. Journal of Applied Ichthyology. 2005;21:451-454
  65. 65. Djomo JE, Garrigues P, Narbonn JF. Uptake and depuration of polycyclic aromatic hydrocarbons from sediment by the zebrafish (Brachydanio rerio). Environmental Toxicology and Chemistry. 1996;15:1177-1181
  66. 66. Frantzen M, Falk-Petersen IB, Nahrgang J, Smith TJ, Olsen GH, Hangstad TA, et al. Toxicity of crude oil and pyrene to the embryos of beach spawning capelin (Mallotus villosus). Aquatic Toxicology. 2012;108:42-52
  67. 67. Carls MG, Rice SD, Hose JE. Sensitivity of fish embryos to weathered crude oil: Part I. Low-level exposure during incubation causes malformations, genetic damage, and mortality in larval pacific herring (Clupea pallasi). Environmental Toxicology and Chemistry. 1999;18:481-493
  68. 68. Heintz RA, Short JW, Rice SD. Sensitivity of fish embryos to weathered crude oil: Part II. Increased mortality of pink salmon (Oncorhynchus gorbuscha) embryos incubating downstream from weathered Exxon valdez crude oil. Environmental Toxicology and Chemistry. 1999;18:494-503
  69. 69. Colavecchia MV, Backus SM, Hodson PV, Parrott JL. Toxicity of oil sands to early life stages of fathead minnows (Pimephales promelas). Environmental Toxicology and Chemistry. 2004;23:1709-1718
  70. 70. Palanikumar L, Kumaraguru AK, Ramakritinan CM, Anand M. Toxicity, feeding rate and growth rate response to sub-lethal concentrations of anthracene and benzo[a]pyrene in milkfish Chanos chanos (Forskkal). Bulletin of Environmental Contamination and Toxicology. 2013;90:60-68
  71. 71. Oberdörster E, Brouwer M, Hoexum-Brouwer T, Manning S, McLachlan JA. Long-term pyrene exposure of grass shrimp, Palaemonetes pugio, affects molting and reproduction of exposed males and offspring of exposed females. Environmental Health Perspectives. 2000;108(7):641-646
  72. 72. Dong W, Wang L, Thornton C, Scheffler BE, Willett KL. Benzo(a)pyrene decreases brain and ovarian aromatase mRNA expression in Fundulus heteroclitus. Aquatic Toxicology. 2008;88:289-300
  73. 73. Danion M, Deschamps MH, Thomas-Guyon H, Bado-Nilles A, Le Floch S, Quentel C, et al. Effect of an experimental oil spill on vertebral bone tissue quality in European sea bass (Dicentrarchus labrax L.). Ecotoxicology and Environmental Safety. 2011;74:1888-1895
  74. 74. Shi X, He C, Zuo Z, Li R, Chen D, Chen R, et al. Pyrene exposure influences the craniofacial cartilage development of Sebastiscus marmoratus embryos. Marine Environmental Research. 2012;77:30-34
  75. 75. Li R, Zuo Z, Chen D, He C, Chen R, Chen Y, et al. Inhibition by polycyclic aromatic hydrocarbons of ATPase activities in Sebastiscus marmoratus larvae: Relationship with the development of early life stages. Marine Environmental Research. 2011;71:86-90
  76. 76. Jee JH, Park KH, Keum YH, Kang JC. Effects of 7,12 dimethylbenz(a)anthracene on growth and haematological parameters in Korean rockfish, Sebastes schlegeli (Hilgendorf). Aquaculture Research. 2006;37:431-442
  77. 77. Le Bihanic F, Clérandeau C, Morin B, Cousin X, Cachot J. Developmental toxicity of PAH mixtures in fish early life stages. Part I: Adverse effects in rainbow trout. Environmental Science and Pollution Research. 2014;21:13720-13731. DOI: 10.1007/s11356-014-2804-0
  78. 78. Sundberg H, Ishaq R, Åkerman G, Tjärnlund U, Zebühr Y, Linderoth M, et al. A bio-effect directed fractionation study for toxicological and chemical characterization of organic compounds in bottom sediment. Toxicological Sciences. 2005;84:63-72
  79. 79. Gilliers C, Claireaux G, Galois R, Loizeau V, Le Pape O. Influence of hydrocarbons exposure on survival, growth and condition of juvenile flatfish: A mesocosm experiment. Journal of Life Sciences. 2012;4:113-122
  80. 80. Gundersen DT, Kristanto SW, Curtis LR, Al-Yakoob SN, Metwally MM, Al-Ajmi D. Subacute toxicity of the water-soluble fractions of Kuwait crude oil and partially combusted crude oil on Menidia beryllina and Palaemonetes pugio. Archives of Environmental Contamination and Toxicology. 1996;31:1-8
  81. 81. Moles A, Rice SD. Effects of crude oil and naphthalene on growth, caloric content, and fat content of pink salmon juveniles in seawater. Transactions of the American Fisheries Society. 1983;112:205-211
  82. 82. Agamy E. Impact of laboratory exposure to light Arabian crude oil, dispersed oil and dispersant on the gills of the juvenile brown spotted grouper (Epinephelus chlorostigma): A histopathological study. Marine Environmental Research. 2013;86:46-55
  83. 83. Ali S, Champagne DL, Richardson MK. Behavioral profiling of zebrafish embryos exposed to a panel of 60 water-soluble compounds. Behavioural Brain Research. 2012;228(2):272-283. DOI: 10.1016/j.bbr.2011.11.020
  84. 84. Chen TH, Wang YH, Wu YH. Developmental exposures to ethanol or dimethylsulfoxide at low concentrations alter locomotor activity in larval zebrafish: Implications for behavioral toxicity bioassays. Aquatic Toxicology. 2011;102:162-166
  85. 85. Egan RJ, Bergner CL, Hart PC, Cachat JM, Canavello PR, Elegante MF, et al. Understanding behavioral and physiological phenotypes of stress and anxiety in zebrafish. Behavioural Brain Research. 2009;205:38-44
  86. 86. López-Patiño MA, Yu L, Cabral H, Zhdanova IV. Anxiogenic effects of cocaine withdrawal in zebrafish. Physiology & Behavior. 2008;93:160-171
  87. 87. Sackerman J, Donegan J, Cunningham C, Nguyen N, Lawless K, Long A, et al. Zebrafish behavior in novel environments: Effects of acute exposure to anxiolytic compounds and choice of Danio rerio line. Journal of Comparative Psychology. 2010;23:43-61
  88. 88. Correia AD, Goncalves R, Scholze M, Ferreira M, Reis-Henriques MA. Biochemical and behavioral responses in gilthead seabream (Sparus aurata) to phenanthrene. Journal of Experimental Marine Biology and Ecology. 2007;347:109-122
  89. 89. Goncalves R, Scholze M, Ferreira AM, Martins M, Correia AD. The joint effect of polycyclic aromatic hydrocarbons on fish behavior. Environmental Research. 2008;108:205-213
  90. 90. Farr AJ, Chabot CC, Taylor DH. Behavioral avoidance of fluoranthene by fathead minnows (Pimephales promelas). Neurotoxicology and Teratology. 1995;17:265-271
  91. 91. Vignet C. Altération de la physiologie des poissons exposés à des hydrocarbures aromatiques polycycliques (HAP): Comportement et reproduction [thesis]. France: Université de La Rochelle; 2014
  92. 92. Anderson MJ, Miller MR, Hinton DE. In vitro modulation of 17-b-estradiolinduced vitellogenin synthesis: Effects of cytochrome P4501A1 inducing compounds on rainbow trout (Oncorhynchus mykiss) liver cells. Aquatic Toxicology. 1996;34:327-350
  93. 93. Holth TF, Nourizadeh-Lillabadi R, Blaesbjerg M, Grung M, Holbech H, Petersen GI, et al. Differential gene expression and biomarkers in zebrafish (Danio rerio) following exposure to produced water components. Aquatic Toxicology. 2008;90:277-291
  94. 94. Pikkarainen A. Ethoxyresorufin-O-deethylase (EROD) activity and bile metabolites as contamination indicators in Baltic Sea perch: Determination by HPLC. Chemosphere. 2006;65:1888-1897
  95. 95. Pj V, Keinänen M, Vuontisjärvi H, Barsiene J, Broeg K, Förlin L, et al. Use of biliary PAH metabolites as a biomarker of pollution in fish from the Baltic Sea. Marine Pollution Bulletin. 2006;53:479-487
  96. 96. Kammann U, Lang T, Vobach M, Wosniok W. Ethoxyresorufin-O-deethylase (EROD) activity in dab (Limanda limanda) as biomarker for marine monitoring. Environmental Science and Pollution Research. 2005;12:140-145
  97. 97. Harman C, Thomas K, Tollefsen KE, Meier S, Bøyum O, Grung M. Monitoring the freely dissolved concentrations of polycyclic aromatic hydrocarbons (PAH) and alkylphenols (AP) around a Norwegian oil platform by holistic passive sampling. Marine Pollution Bulletin. 2009;58:1671-1679
  98. 98. Zm T, Amb G, Hansen R, Andersen O. 1-Hydroxypyrene as a biomarker of PAH exposure in the marine polychaete Nereis diversicolor. Marine Environmental Research. 2009;67:38-46
  99. 99. McMaster ME, Van Der Kraak GJ, Munkittrick KR. An epidemiological evaluation of the biochemical basis for steroid hormonal depressions in fish exposed to industrial wastes. Journal of Great Lakes Research. 1996;22:153-171
  100. 100. Van der Oost R, Beyer J, Vermeulen NPE. Fish bioaccumulation and biomarkers in environmental risk assessment: A review. Environmental Toxicology and Pharmacology. 2003;13:57-149
  101. 101. Pointet K, Milliet A. PAHs analysis of fish whole gall bladders and livers from the Natural Reserve of Camargue by GCy MS. Chemosphere. 2000;40:293-299
  102. 102. Eggens ML, Opperhuizen A, Boon JP. Temporal variation of CYP1A indices, PCB and 1-OH pyrene concentration in flounder, Platichthys flesus, from the Dutch Wadden Sea. Chemosphere. 1996;33:1579-1596
  103. 103. Hylland K, Sandvik M, Skasre JU, Beyer J, Egaas E, Goksøyr A. Biomarkers in flounder (Platichthys flesus): An evaluation of their use in pollution monitoring. Marine Environmental Research. 1996;42:223-227
  104. 104. Ruddock PJ, Bird DJ, McCalley DV. Bile metabolites of PAHs in three species of fish from the Severn Estuary. Ecotoxicology and Environmental Safety. 2002;51:97-105
  105. 105. Ariese F, Kok SJ, Verkaik M, Gooijer C, Velthorst NH, Hofstraat JW. Synchronous fluorescence spectrometry of fish bile: A rapid screening method for the biomonitoring of PAH exposure. Aquatic Toxicology. 1993;26:273-286
  106. 106. Zhou JL, Fileman TW, Evans S, Donkin P, Llewellyn C, Readman JW, et al. Fluoranthene and pyrene in the suspended particulate matter and surface sediments of the Humber Estuary, UK. Marine Pollution Bulletin. 1998;36:587-597
  107. 107. Lin ELC, Cormier SM, Racine RN. Synchronous fluorometric measurement of metabolites of polycyclic aromatic hydrocarbons in the bile of brown bullhead. Environmental Toxicology and Chemistry. 1994;13:707-715
  108. 108. Neff JM. Sources of PAH in the aquatic environment. In: Polycyclic Aromatic Hydrocarbons in the Aquatic Environment. London: Applied Science Publishers; 1979. pp. 7-43
  109. 109. Cavalieri E, Rogan E. Role of radical cations in aromatic hydrocarbon carcinogenesis. Environmental Health Perspectives. 1985;64:69-84
  110. 110. Johnston EP, Baumann PC. Analysis of fish bile with HPLC fluorescence to determine environmental exposure to benzo(a)pyrene. Hydrobiologia. 1989;188(1):561-566
  111. 111. Penning TM, Burczynski ME, Hung CF, McCoull KD, Palackal NT, Tsuruda L. Dihydrodiol dehydrogenases and polycyclic aromatic hydrocarbon activation: Generation of reactive and redox active O-quinones. Chemical Research in Toxicology. 1999;12:1-18
  112. 112. Hisano T, Hayase T. Countermeasures against water pollution in enclosed coastal seas in Japan. Marine Pollution Bulletin. 1991;23:479-484
  113. 113. Moriarty C. Eels: A Natural and Unnatural History. Londres: David & Charles; 1978. 192 p
  114. 114. Van Schooten FJ, Maas LM, Moonen EJC, Kleinjans JCS, van der Oost R. DNA dosimetry in biological indicator species living on PAH-contaminated soils and sediments. Ecotoxicology and Environmental Safety. 1995;30(2):171-179
  115. 115. Wheeler A. The eels. The Fishes of the British Isles and North-West Europe. London: Macmillan & Co.; 1969. pp. 223-230
  116. 116. Sandvik M, Horsberg TE, Skaare JU, Ingebrigtsen K. Comparison of dietary and waterborne exposure to benzo[a]pyrene: Bioavailability, tissue disposition and CYP1A1 induction in rainbow trout (Oncorhynchus mykiss). Biomarkers. 1998;3:399-410
  117. 117. Kammann U, Akcha F, Budzinski H, Burgeot T, Gubbins MJ, Lang T, et al. PAH metabolites in fish bile: From the Seine Estuary to Iceland. Marine Environmental Research. 2017;124:41-45. DOI: 10.1016/j.marenvres. 2016.2.014
  118. 118. Kammann U, Brinkmann M, Freese M, Pohlmann JD, Stoffels S, Hollert H, et al. PAH metabolites, GST and EROD in European eel (Anguilla anguilla) as possible indicators for eel habitat quality in German rivers. Environmental Science and Pollution Research. 2014;1:2519-2530
  119. 119. Ruddock PJ, Bird DJ, McEvoy J, Peters LD. Bile metabolites of polycyclic aromatic hydrocarbons (PAHs) in European eels Anguilla anguilla from United Kingdom estuaries. Science of the Total Environment. 2003;301:105-117
  120. 120. Wariaghli F, Kammann U, Hanel R, Yahyaoui A. PAH metabolites in bile of European Eel (Anguilla anguilla) from Morocco. Bulletin of Environmental Contamination and Toxicology. 2015;95(6):740-744

Written By

Ayoub Baali and Ahmed Yahyaoui

Submitted: 06 December 2018 Reviewed: 04 April 2019 Published: 27 September 2019