GET THE APP

Towards A Convenient Procedure to Characterize Polycyclic Aromati
Journal of Pollution Effects & Control

Journal of Pollution Effects & Control
Open Access

ISSN: 2375-4397

+44 1223 790975

Research Article - (2015) Volume 3, Issue 1

Towards A Convenient Procedure to Characterize Polycyclic Aromatic Hydrocarbons in Sediments Receiving Industrial Effluents

Nadia Morin-Crini*, Coline Druart, Caroline Amiot, Frédéric Gimbert, Etienne Chanez and Grégorio Crini
Department of Chrono-Environment, UMR UFC/CNRS 6249 USC INRA, University of Franche-Comté, 16 Route de Gray, 25030 Besançon cedex, France
*Corresponding Author: Nadia Morin-Crini, Department of Chrono-Environment, UMR UFC/CNRS 6249 USC INRA, University of Franche-Comté, 16 Route De Gray, 25030 Besançon Cedex, France, Tel: +33 (0) 381 665 786, Fax: +33 (0) 381 666 083; Email:

Abstract

Using pressurized liquid extraction and GC-MS/MS, Polycyclic Aromatic Hydrocarbons (PAH) levels were determined in pond and river sediments receiving effluents from a Chemicals Installation (CI) and a Surface Treatment Installation (STI), respectively. Maximum values were obtained for the STI site with total PAH concentrations of 3000-4000 ng g-1 compared to 200-2500 ng g-1 for the CI site. Furthermore, in the river (STI), for two PAHs (phenanthrene and acenaphthylene), one sample presented concentrations that exceeded the probable effect concentrations leading to this sediment being classified as toxic. However, PAH levels were higher upstream of the STI discharge water than downstream, indicating sediment PAH accumulation was not exclusively due to this industrial activity. At the CI site, the concentrations found at different points showed that in the pond, PAHs were derived from industrial activities but were rapidly diluted in the water mass. PAH fingerprinting following various guidelines, revealed the prevalence of a pyrolytic origin.

Keywords: Freshwater sediments; Hydrocarbons; Industrial sites; GC-MS/MS; Pyrolytic origin; Pressurized liquid extraction

Introduction

Polycyclic Aromatic Hydrocarbons (PAHs) are organic pollutants with two or more fused aromatic rings. They are introduced into the environment via natural and anthropogenic processes and contaminate all environmental compartments. Indeed, they have already been found in food or food supplements [1], marine organisms [2-4], water [5-7], sewage sludge [8], dust particles, soils and sediments [9-12]. Due to their high toxicity to humans [13,14] and aquatic organisms [15], they have become a focus for scientific research. Consequently, it was necessary to develop analytical methods able to quantify this pollutant family. Numerous methods have already been developed for PAH extraction in various environmental matrices, such as Pressurized Liquid Extraction (PLE), Supercritical Fluid Extraction (SFE), subcritical water extraction, dispersive liquid-liquid microextraction (DLLME), Solid-Phase Extraction (SPE) or microextraction (SPME), Stir Bar Sorptive Extraction (SBSE) or soxhlet extraction [16-20].

Due to their chemical properties such as low water solubility, PAHs released into the aquatic environment tended to sorb onto suspended particles and sediments and could therefore impact both benthic organisms, directly in contact with sediments, and pelagic microphagous organisms feeding on suspended particles [21,22]. It could be interesting to determine the origin of PAHs which could be pyrogenic (incomplete combustion of organic material from anthropogenic activity or natural source such as forest fire), petrogenic (crude oil or coal) or biogenic (diagenesis) sources in freshwater sediments, given that their presence has already been shown in industrial discharge waters [3,23,24] usually flushed into the freshwater environment. Thus, it is necessary to develop selective and sensitive methods of extraction and analysis of freshwater sediments from industrial sites, because of the simultaneous presence of a large quantity of various pollutant types (salts, metal ions, organics) as described by Morin-Crini et al. [23].

Firstly, the present work aimed to develop, in freshwater sediments, a specific, rapid and sensitive analytical method for the quantification of the 16 PAHs which have been selected by the US Environmental Protection Agency (US EPA) as priority pollutants (Figure 1). Secondly, we investigated the presence of PAHs in sediments in contact with discharge waters from industrial sites carrying out chemistry and surface treatment activities. Finally, we propose to determine the PAH compositional origin through three indicators based on the number of rings, isomeric ratios and total index (based on weighted-isomeric ratios).

pollution-effects-Chemical-structure

Figure 1: Chemical structure of the 16 PAHs defined as priority pollutants by the US EPA.

Materials and Methods

Chemicals

Sixteen PAHs and three deuterated internal standards (ISTD) were purchased from Supelco Sigma Aldrich (Saint Quentin Fallavier, France) and Interchim (Montluçon, France): naphthalene (NAP), acenaphthene (ACE and ACE-d10), acenaphthylene (ACY), fluorene (FLU), phenanthrene (PHE), anthracene (ANT), fluoranthene (FLT), pyrene (PYR and PYR-d10), benzo[a]anthracene (BaANT), chrysene (CHY), benzo[b]fluoranthene (BbFLT), benzo[k]fluoranthene (BkFLT), benzo[a]pyrene (BaPYR and BaPYR-d12,), dibenzo[a,h] anthracene (dBahANT), indeno[1,2,3-cd]pyrene (IcdPYR), benzo[g,h,i]perylene (BghiPL). n-hexane and methanol Distol, a range of products for organic trace analysis, were obtained from Fisher Scientific (Illkirch, France). Acetone (analytical grade, Fisher Scientific, Illkirch, France), Decon-90 (Fisher Scientific, Illkirch, France) and ultra-pure water (Millipore Milli-Q Integral 3 system, Molsheim, France) were used for glassware clean-up.

PAH extraction

Different experiments were performed to improve PAH extraction efficiencies using sediment from a river located in a rural (forest) and weakly anthropized watershed, referred to as blank. The blank sediment was spiked at 50 ng g-1 dry mass. However, to allow equilibrium to be reached between PAHs and matrix, a contact time of one night in the dark and at 4°C (to prevent photodegradation) was applied. Comparison between PLE and ultrasonic extraction (USE) was investigated. For USE, 5 g of sediment was sonicated with 50 mL of solvent (hexane or methanol) for 15 min, conditions adapted from Martinez et al. [25]. The extract was centrifuged (5 min at 4500 rpm) then the sediment was retrieved and sonicated a second time in the same conditions. Both fractions were concentrated and injected separately into the GCMS/ MS apparatus and then summed. Pressurized Liquid Extraction (PLE) was performed using a Büchi (Rungis, France) SpeedExtractor E-914 system equipped with stainless steel extraction cells (volume 40 mL). The extraction cell was filled from the bottom to the top with the following materials: first, a glass fiber filter, then a few grams of pure quartz sand followed by the dried sediment sample (5.000 ± 0.001 g), covered by glass beads and finally, a cellulose filter. The sediment samples were extracted with hexane or methanol under the following conditions: temperature: 100°C; pressure: 10 MPa; cycles: 1 or 2; heatup: 1 min; hold: 10 min; discharge: 2 min; flush with solvent: 1 min; flush with nitrogen gas: 2 min. In both cases, USE or PLE, the extracts were concentrated using a Syncore Analyst (Büchi, Rungis, France). In the case of PLE, it was the same vessel which was used both for collected extracts and their concentration, avoiding the loss of organic compound during transfer. Moreover, a locally cooled appendix on the sample vessel prevented evaporation to dryness (residual volume of roughly 0.3 mL). The concentrated extracts of PLE and USE were transferred to a 1 mL volumetric flask and made up with hexane.

PAH analysis

Sample extracts were analyzed in the GC-MS/MS apparatus (Agilent, Massy, France) which included a 7890A GC system, a 7000 GC triple quadrupole mass spectrometer and an 80 GC Combipal autosampler. 1 μL of the extract was injected at a temperature of 300°C in splitless mode under a constant He flow (purity 99.9999 %) at 1.5 mL min-1 followed by a purge flow to split vent after 0.5 min. We used a (5 % phenyl)-methylpolysiloxane HP5MS column (30 m x 0.25 mm i.d, 0.25 μm film thickness, Agilent 19091J-433). The temperature gradient was from 45°C (hold 0.5 min) to 100°C at 30°C min-1 then from 100°C to 325°C at 10°C min-1 (hold 5 min). Total run time was 30 min. The temperatures of the transfer line, ion source and quadrupoles were 300, 300 and 150°C respectively. An electron-impact ion source at 70 eV was required. Nitrogen with a purity of 99.9 % was produced by a NiGen LC-MS 40-1 Claind generator (Gengaz, Wasquehal, France) and used as collision gas at 1.5 mL min-1. For data analysis, Agilent MassHunter B.05.02.1032 software was used. For apparatus performance controls, an autotune was performed weekly at m/z 69, 264 and 502 from the electron-impact of perfluorotributylamine (PFTBA).

Validation of extraction method

A calibration graph was established for each of the 16 compounds using blank sediment spiked at five increasing concentrations (10, 20, 40, 60 and 100 ng g-1 dry mass). The spiked calibration samples were treated following the validated extraction procedure: PLE in hexane with 2 cycles. Calibration curves were constructed by plotting the analyte/ISTD peak area ratios versus the analyte concentration. Five experimental replicates for each concentration were carried out to validate the method through the following parameters: linearity, Limit of Detection (LOD), precision and trueness.

Field samples

The method was applied to field samples of sediments receiving industrial effluents. Sediments were manually collected at two sites using a modified pickaxe which sampled 0.05 m2 to a depth of 5 cm. At the Chemical Installation (CI) site, sediments were sampled at six locations in the pond in which the wastewater was released (Figure 2). Points 2 to 6 were situated along a transect starting from the nearest to the discharge to the furthest whereas point 1 was isolated upstream of the discharge water by a dike (Figure 2). At the Surface Treatment Installation (STI) site, sediments were sampled at four locations in the river in which the discharge water was released. Points Cup and Dup were situated upstream of the discharge water, close (110 m) and distant (645 m) respectively, whereas Cdown and Ddown were situated downstream of the discharge water, close (130 m) and distant (545 m), respectively. An aliquot of each sample was immediately lyophilized until constant mass and then sieved through a 0.5 mm mesh. The other aliquot was stored at -20°C in case of need for further analyses. Indeed, additional analysis was performed on sediment from point 2 (site CI) which had been centrifuged for 15 minutes at 6100 rpm to remove interstitial water, before being lyophilized.

pollution-effects-sampling-points

Figure 2: Location of the six sampling points along the length of the chemical industry pond.

Regulatory standards to sediment toxicity classification

The concentrations determined in field samples were compared to different types of reference values which defined threshold concentrations regarding biological effects (Table 1). The effects range low (ERL) and effects range median (ERM) provided chemical concentration ranges that are rarely, occasionally, or frequently associated with adverse biological effects [26,27]. The threshold and probable effect concentrations (TEC and PEC, respectively) proposed by Zeng et al. [28] and derived from several guidelines [29-31] allowed us to judge whether sediment was toxic or nontoxic.

  ERL ERM TEC PEC
NAP 160 2100 176 561
ACY 44 640 5.87 128
ACE 16 500 6.71 88.9
FLU 19 540 77.4 536
PHE 240 1500 204 1170
ANT 85.3 1100 57.2 845
FLT 600 5100 423 2230
PYR 665 2600 195 1520
BaANT 261 1600 108 1050
CHY 384 2800 166 1290
BbFLT nd nd nd nd
BkFLT nd nd 240 13400
BaPYR 430 1600 150 1450
IcdPYR nd nd 200 3200
dBahANT 63.4 260 33 135
BghiPL nd nd 170 3200
Σ PAHs 4022 44792 1610 22800
nd non defined

Table 1: Reference values in sediments (ng g-1) [26-31] and associated biological effects (rarely adverse effect < ERL < occasionally < ERM < frequently; nontoxic < TEC < neither toxic nor nontoxic < PEC < toxic).

Methods for determining the composition and hence the origin of the PAHs

Several approaches were used to determine the source of PAH pollution (petrogenic, biogenic and pyrogenic). The first, which was described by several authors [5,32,33] stated that pyrogenic PAHs were dominated by heavier molecules with 4 to 6 rings. The second approach used analyses of selected PAH isomer ratios [22,34,35]:

where [ANT], [PHE], [FLT], [PYR], [BaANT], [CHY], [IcdPYR], [BghiPL] were PAH concentrations.

Ratios above 0.10 at mass 178, 0.50 at mass 202, 0.35 at mass 228 and 0.20 at mass 276 represented dominance of pyrogenic origin.

The third approach used the total index defined by Orecchio [36] which stated that a value above 4 indicated a pyrogenic process:

where ratios 178 to 276 were defined in equations (1) to (4).

Results and Discussion

Investigations of extraction parameters

Whatever the solvent used (hexane or methanol) and the PLE cycle number, the extraction yields were still higher with PLE than with USE. For instance, in hexane and with two extraction cycles, an average efficiency increase of 36% was obtained with the use of PLE compared to both fractions summed in USE. Thus, all results presented hereunder were realized with the PLE method. Investigations on the choice of the extraction solvent, hexane or methanol, showed that, except for BaPYR, IcdPYR, dBahANT and BghiPL, the signal to noise ratio (S/N) was increased with the use of hexane up to 500-fold at a given concentration. These results are in accordance with those of Martinez et al. [25] and Foan and Simon [37] who indicated that the use of polar solvent, such as methanol, caused chromatographic interference due to the co-extraction of polar compounds. Finally, it was demonstrated that performing the extraction in two cycles instead of only one increased the average efficiency by 13%.

Method validation

Correlation coefficients of each compound were always higher than 0.9 except for NAP and IcdPYR (Table 2). These results are in agreement with those of Burkhardt et al. [38] who obtained the lowest R2 value for NAP. Precision, which was assessed through the Residual Standard Deviations (RSD), gave suitable values for all PAHs (RSD values ≤ 30%; Table 2). The trueness was acceptable, since recoveries were determined between 98 % (BghiPL) and 141 % (NAP) (Table 2). High sensitivity was obtained for most PAHs (LOD between 0.2 and 5.5 ng g-1 for 1 μL injection volume, 5 g sediment, Table 2). However, NAP and PHE gave higher LOD values (18.6 and 10.6 ng g-1 respectively). The LODs obtained in this study were better than those observed by the GC-MS analysis of 1 μL of extract obtained by PLE and SPE of 25 g of sediment [38]. Similar values were obtained for an injection volume of 10 μL and a mass of extracted sample of 0.5 g, by USE and SPE coupled to reversed phase liquid chromatography with UV detection [39] and for Soxhlet extraction of 3 g sediment and an injection volume of 1 μL in GC-MS [36]. Nevertheless, very high sensitivity was obtained by a longer, multi-step method composed of PLE followed by stir bar sorptive extraction coupled with thermal desorption and GC-MS/MS analysis [40].

  Correlation coefficient LOD (ng g-1) RSD (%) Recovery (%)
NAP 0.634 18.6 27 141
ACY 0.915 1.1 19 119
ACE 0.957 2.1 14 114
FLU 0.961 2.0 14 116
PHE 0.941 10.6 15 123
ANT 0.935 1.7 17 122
FLT 0.935 5.5 16 121
PYR 0.951 3.5 15 111
BaANT 0.944 0.5 16 113
CHY 0.924 1.9 16 112
BbFLT 0.967 1.4 18 102
BkFLT 0.941 1.7 15 106
BaPYR 0.943 0.5 16 122
IcdPYR 0.886 0.6 27 106
dBahANT 0.938 0.2 26 119
BghiPL 0.941 0.7 24 98

Table 2: Correlation coefficient (R2), Detection Limits (LOD), Relative Standard Deviations (RSD) and recoveries for the determination of the 16 PAHs.

Analysis of field samples

The validated method was applied to PAH determination in sediments from two stations linked to industrial activities. At the CI site, note that among the 16 PAHs studied, 14 were systematically found in all samples (Table 3). The highest values were obtained for point 2, located closest to the discharge water release point from the chemical industry with a total PAH content of 2467 ng g-1. Because two compounds (BaANT and CHY) presented concentrations higher than ERL (Tables 1 and 3), the sediment from point 2 could cause adverse biological effects. In return, considering TEC and PEC (Table 1), this sample was judged to be neither toxic nor nontoxic firstly because concentrations of 6 individual PAHs (BaANT, CHY, BkFLT, BaPYR, dBahANT and BghiPL) and secondly the sum of PAHs were between the TEC and PEC values. To determine if the high concentrations at point 2 came from the sediment itself or from the interstitial water, the sample was analyzed a second time with the addition of a centrifugation step to separate the liquid and solid phases of the sediment. The results indicated that concentrations decreased by a factor of 15 to 30 for BaANT, CHY, BbFLT, BkFLT, BaPYR, IcdPYR, dBahANT and BghiPL. Thus, PAHs present at point 2, were mainly concentrated in the interstitial water in colloidal or dissolved forms. The low concentrations found at point 1 (Table 3), always below ERL and TEC (Table 1), were expected since this sample was isolated from the discharge water by the dike. Thus, point 1 could be defined as a relative field control sample. Taking these considerations into account, the sample concentrations at points 3 to 6 seemed to regain this base level, reflecting a sharp decrease in the sum of PAH concentrations from the input point. Various abiotic and biotic phenomena may explain these results, from simple PAH dilution in the pond to more complex biogeochemical degradation processes [41,42].

  Point 6 Point 5 Point 4 Point 3 Point 2 Point 1
  NAP   31.12   < LOD   < LOD   < LOD   62.49   23.12
ACY 1.91 1.73 2.01 2.23 2.46 2.09
ACE 3.26 < LOD < LOD < LOD 5.10 < LOD
FLU 19.73 8.85 5.88 8.36 15.17 16.81
PHE 48.40 30.45 25.76 32.02 54.33 44.18
ANT 5.82 4.29 4.37 5.48 7.24 5.46
FLT 35.47 35.05 35.80 42.56 56.24 40.16
PYR 33.10 28.12 29.60 34.75 48.93 30.92
BaANT 11.38 12.81 13.53 15.69 286.88 12.77
CHY 21.35 23.39 24.10 28.46 528.68 24.01
BbFLT 15.55 15.56 16.59 18.87 344.30 15.10
BkFLT 16.86 17.31 19.23 21.19 359.22 18.51
BaPYR 13.68 12.91 13.7 15.96 275.61 13.74
IcdPYR 7.59 8.25 7.93 8.32 138.07 6.22
dBahANT 2.01 2.96 5.22 2.32 44.63 2.00
BghiPL 12.88 11.87 12.03 13.95 237.75 11.27
Σ PAHs 280.1 213.6 215.8 250.2 2467 266.3
Ratio 178 0.11 0.12 0.15 0.15 0.12 0.11
Ratio 202 0.52 0.55 0.55 0.55 0.53 0.57
Ratio 228 0.35 0.35 0.36 0.36 0.35 0.35
Ratio 276 0.37 0.41 0.40 0.37 0.37 0.36
Total index 6.05 6.36 6.48 6.41 6.18 6.14

Table 3: Concentrations (ng g-1), isomeric ratios and total index of PAHs from pond sediments from chemical installation site.

In the STI site, all samples were highly contaminated, notably in comparison with the CI site (Tables 3 and 4). Indeed, they all exceed the ERL values for at least one compound and considering the sum of PAH concentrations, they were defined as neither toxic nor nontoxic (Tables 1 and 4). However, with regard to PHE (1337 ng g-1) and ACY (231 ng g-1) concentrations, Dup sediment can be classified as toxic since it overshot PEC values for these compounds (Table 1). Note that this point was the most contaminated among the four locations along the river. However, this sample originated from a location 545 meters upstream of the STI site, so its high levels of contamination cannot be attributed to the STI activities. Therefore, at least a portion of the PAH levels measured downstream of the discharge waters must result from this upstream sediment contamination. However, considering the differences in PAH concentrations in sediments just before and just after the discharge water (Cup and Cdown, respectively, Table 4), the STI activities also contributed to the PAHs released in the river (+23% for the sum of PAHs).

  Dup Cup Cdown Ddown
NAP 294.00
ACY 231.00 46.00 39.80 37.30
ACE 46.50 3.70 4.40 6.90
FLU 225.00 18.10 16.70 21.70
PHE 1336.5 255.20 259.40 225.80
ANT 467.50 79.70 69.80 60.80
FLT 1728.30 718.70 879.40 672.90
PYR 1496.50 468.70 578.80 441.50
BaANT 668.34 368.43 390.37 252.03
CHY 456.75 301.60 348.98 216.40
BbFLT 1122.70 391.20 498.80 399.50
BkFLT 929.20 388.67 432.47 329.87
BaPYR 1037.30 367.30 390.90 390.20
IcdPYR 296.37 107.40 115.50 105.70
dBahANT 227.37 81.64 88.14 72.80
BghiPL 365.40 145.90 172.30 143.10
Σ PAHs 10929 3742 4286 3377
Ratio 178 0.26 0.24 0.21 0.21
Ratio 202 0.54 0.61 0.60 0.60
Ratio 228 0.59 0.55 0.53 0.54
Ratio 276 0.45 0.42 0.40 0.42
Total index 9.47 9.05 8.65 8.80

Table 4: Concentrations (ng g-1), isomeric ratios and total index of PAHs from river sediments from surface treatment industrial site.

As the two sites studied were close to industrial areas, we can assume that most PAHs were of anthropogenic pyrogenic origin (furnaces, industrial production, etc.) rather than from a petroleum source. It can be observed that PAHs with 4-6 rings made up the major part of total PAHs in both sites (Figure 3), which was 61-94% and 76- 91% for CI and STI sites, respectively, confirming that the PAH arose from combustion processes. The average ratios 178 were 0.12 and 0.23 for sediments from CI and STI sites, respectively (Tables 3 and 4) stressing the dominance of combustion origin. For ratio 202, the values were situated in the range 0.52-0.61 for both sites (Tables 3 and 4) indicating rather biomass or coal combustion. An average value of 0.35 was obtained for ratio 228 at site CI (Table 3), corresponding to the limit between petrogenic and pyrolytic sources. The average value of 0.55 obtained at site STI (Table 4) pointed to a combustion origin. For CI and STI sediments, ratio 276 gave average values of 0.38 and 0.42 respectively, suggesting combustion processes. However, whatever the site and the location, all total indexes were above 4 (Tables 3 and 4), indicating high temperature combustion processes and thus confirming a mainly pyrolytic origin of the PAHs found in the sediments impacted by both industrial sites.

Conclusion

In this work we developed and validated a method based on PLEGC- MS/MS for the analysis of PAHs, allowing their identification and quantification at ultra-trace levels in freshwater sediments. The results showed large concentration differences between the chemical installation and surface treatment installation sites, STI samples having total PAH values situated in the range 210-2500 ng g-1 against 3400- 11000 ng g-1 for CI samples. After checking against the regulatory values which classified sediments as toxic or not, the present study showed that these media sometimes reached hazardous concentrations for the aquatic environment. Following fingerprinting to determine the PAH origin, the total index indicated that whatever the location, the PAHs were of pyrolytic origin.

Acknowledgements

The authors thank the Agence de l’Eau Rhône Méditerranée Corse, Conseil Régional de Franche-Comté and the FEDER (Fonds Européens de Développement Régional) for financial support. The authors wish to thank Dr Peter Winterton (University Toulouse III, Toulouse, France) for his English grammar and syntax review and his critical reading.

Conflict of Interests

The authors declare that they have no conflict of interest.

References

  1. Danyi S, Brose F, Brasseur C, Schneider YJ, Larondelle Y, et al. (2009) Analysis of EU priority polycyclic aromatic hydrocarbons in food supplements using high performance liquid chromatography coupled to an ultraviolet, diode array or fluorescence detector. Anal Chim Acta 633: 293-299.
  2. Jaouen-Madoulet A, Abarnou A, Le Guellec AM, Loizeau V, Leboulenger F (2000) Validation of an analytical procedure for polychlorinated biphenyls, coplanar polychlorinated biphenyls and polycyclic aromatic hydrocarbons in environmental samples. J Chromatogr A 886: 153-173.
  3. Naes K, Axelman J, Naf C, Broman D (1998) Role of soot carbon and other carbon matrices and the distribution of PAHs among particles, DOC, and the dissolved phase in the effluent and recipient waters of an aluminium reduction plant. Environ Sci Technol 32: 17^86-1792.
  4. Serpe FP, Esposito M, Gallo P, Serpe L (2010) Optimisation and validation of an HPLC method for determination of polycyclic aromatic hydrocarbons (PAHs) in mussels. Food Chem 122: 920-925.
  5. Bihari N, Fafandel M, Piskur V (2007) Polycyclic aromatic hydrocarbons and ecotoxicological characterization of seawater, sediment, and mussel Mytilus galloprovincialis from the Gulf of Rijeka, the Adriatic Sea, Croatia. Arch Environ Contam Toxicol 52: 379-387.
  6. Law RJ, Dawes VJ, Woodhead RJ, Matthiessen P (1997) Polycyclic aromatic hydrocarbons (PAH) in seawater around England and Wales. Mar Pollut Bull 34: 306-322.
  7. Oros DR, Ross JR, Spies RB, Mumley T (2007) Polycyclic aromatic hydrocarbon (PAH) contamination in San Francisco Bay: a 10-year retrospective of monitoring in an urbanized estuary. Environ Res 105: 101-118.
  8. Santos JL, Aparicio I, Alonso E (2007) A new method for the routine analysis of LAS and PAH in sewage sludge by simultaneous sonication-assisted extraction prior to liquid chromatographic determination. Anal Chim Acta 605: 102-109.
  9. Christensen JH, Tomasi G, de Lemos Scofield A, de Fatima Guadalupe Meniconi M (2010) A novel approach for characterization of polycyclic aromatic hydrocarbon (PAH) pollution patterns in sediments from Guanabara Bay, Rio de Janeiro, Brazil. Environ Pollut 158: 3290-3297.
  10. Nemirovskaya IA (2010) The concentration and composition of hydrocarbons in water, particulate matter, and bottom sediments of the Kara sea. Oceanology 50: 716-728.
  11. Schulze T, Seiler TB, Streck G, Braunbeck T, Hollert H (2012) Comparison of different exhaustive and biomimetic extraction techniques for chemical and biological analysis of polycyclic aromatic compounds in river sediments. J Soil Sediment 12: 1419-1434.
  12. Smith JN, Lee K, Gobeil C, Macdonald RW (2009) Natural rates of sediment containment of PAH, PCB and metal inventories in Sydney Harbour, Nova Scotia. Sci Total Environ 407: 4858-4869.
  13. Roos PH, Tschirbs S, Pfeifer F, Welge P, Hack A, et al. (2004) Risk potentials for humans of original and remediated PAH-contaminated soils: application of biomarkers of effect. Toxicology 205: 181-194.
  14. Sánchez-Guerra M, Pelallo-Martínez N, Díaz-Barriga F, Rothenberg SJ, Hernández-Cadena L, et al. (2012) Environmental polycyclic aromatic hydrocarbon (PAH) exposure and DNA damage in Mexican children. Mutat Res 742: 66-71.
  15. Cachot J, Geffard O, Augagneur S, Lacroix S, Le Menach K, et al. (2006) Evidence of genotoxicity related to high PAH content of sediments in the upper part of the Seine estuary (Normandy, France). Aquat Toxicol 79: 257-267.
  16. Fatoki OS, Ximba BJ, Opeolu BO (2011) Polycyclic aromatic hydrocarbons (PAHs) in food and environmental samples: an overview. Fresenius Environ Bull 20: 2012-2020.
  17. Hawthorne SB, Grabanski CB, Martin E, Miller DJ (2000) Comparisons of soxhlet extraction, pressurized liquid extraction, supercritical fluid extraction and subcritical water extraction for environmental solids: recovery, selectivity and effects on sample matrix. J Chromatogr A 892: 421-433.
  18. Itoh N, Numata M, Aoyagi Y, Yarita T (2008) Comparison of low-level polycyclic aromatic hydrocarbons in sediment revealed by Soxhlet extraction, microwave-assisted extraction, and pressurized liquid extraction. Anal Chim Acta 612: 44-52.
  19. Kim JH, Moon JK, Li QX, Cho JY (2003) One-step pressurized liquid extraction method for the analysis of polycyclic aromatic hydrocarbons. Anal Chim Acta 498: 55-60.
  20. Wang XW, Lin L, Luan TG, Yang LH, Tam NFY (2012) Determination of hydroxylated metabolites of polycyclic aromatic hydrocarbons in sediment samples by combining subcritical water extraction and dispersive liquid-liquid microextraction with derivatization. Anal Chim Acta 753: 57-63.
  21. Cailleaud K, Forget-Leray J, Souissi S, Hilde D, Lemenach K, et al. (2007) Seasonal variations of hydrophobic organic contaminant concentrations in the water-column of the Seine Estuary and their transfer to a planktonic species Eurytemora affinis (Calanoïda, copepoda). Part 1: PCBs and PAHs. Chemosphere 70: 270-280.
  22. Readman JW, Fillmann G, Tolosa I, Bartocci J, Villeneuve JP, et al. (2002) Petroleum and PAH contamination of the Black Sea. Mar Pollut Bull 44: 48-62.
  23. Morin-Crini N, Druart C, Gavoille S, Lagarrigue C, Crini G (2013) Analytical monitoring of the chemicals present in the discharge water generated by the surface treatment industry. J Environ Protection 4: 53-60.
  24. Umbuzeiro Gde A, Kummrow F, Roubicek DA, Tominaga MY (2006) Evaluation of the water genotoxicity from Santos Estuary (Brazil) in relation to the sediment contamination and effluent discharges. Environ Int 32: 359-364.
  25. Martinez E, Gros M, Lacorte S, Barceló D (2004) Simplified procedures for the analysis of polycyclic aromatic hydrocarbons in water, sediments and mussels. J Chromatogr A 1047: 181-188.
  26. Mai BX, Fu JM, Sheng GY, Kang YH, Lin Z, et al. (2002) Chlorinated and polycyclic aromatic hydrocarbons in riverine and estuarine sediments from Pearl River Delta, China. Environ Pollut 117: 457-474.
  27. McCauley DJ, DeGraeve GM, Linton TK (2000) Sediment quality guidelines and assessment: overview and research needs. Environ Sci Policy 3, Supplement 1: 133-144.
  28. Zeng S, Zeng L, Dong X, Chen J (2013) Polycyclic aromatic hydrocarbons in river sediments from the western and southern catchments of the Bohai Sea, China: toxicity assessment and source identification. Environ Monit Assess 185: 4291-4303.
  29. Hong SH, Yim UH, Shim WJ, Oh JR, Viet PH, et al. (2008) Persistent organochlorine residues in estuarine and marine sediments from Ha Long Bay, Hai Phong Bay, and Ba Lat Estuary, Vietnam. Chemosphere 72: 1193-1202.
  30. Ingersoll CG, MacDonald DD, Wang N, Crane JL, Field LJ, et al. (2001) Predictions of sediment toxicity using consensus-based freshwater sediment quality guidelines. Arch Environ Contam Toxicol 41: 8-21.
  31. MacDonald DD, Ingersoll CG, Berger TA (2000) Development and evaluation of consensus-based sediment quality guidelines for freshwater ecosystems. Arch Environ Contam Toxicol 39: 20-31.
  32. Boitsov S, Jensen HK, Klungsøyr J (2009) Natural background and anthropogenic inputs of polycyclic aromatic hydrocarbons (PAH) in sediments of South-Western Barents Sea. Mar Environ Res 68: 236-245.
  33. Culotta L, De Stefano C, Gianguzza A, Mannino MR, Orecchio S (2006) The PAH composition of surface sediments from Stagnone coastal lagoon, Marsala (Italy). Mar Chem 99: 117-127.
  34. Baumard P, Budzinski H, Garrigues P (1998) PAHs in Arcachon Bay, France: Origin and biomonitoring with caged organisms. Mar Pollut Bull 36: 577-586.
  35. Yunker MB, Macdonald RW, Vingarzan R, Mitchell RH, Goyette D, et al. (2002) PAHs in the Fraser River basin: a critical appraisal of PAH ratios as indicators of PAH source and composition. Org Geochem 33: 489-515.
  36. Orecchio S (2010) Contamination from polycyclic aromatic hydrocarbons (PAHs) in the soil of a botanic garden localized next to a former manufacturing gas plant in Palermo (Italy). J Hazard Mater 180: 590-601.
  37. Foan L, Simon V (2012) Optimization of pressurized liquid extraction using a multivariate chemometric approach and comparison of solid-phase extraction cleanup steps for the determination of polycyclic aromatic hydrocarbons in mosses. J Chromatogr A 1256: 22-31.
  38. Burkhardt MR, Zaugg SD, Burbank TL, Olson MC, Iverson JL (2005) Pressurized liquid extraction using water/isopropanol coupled with solid-phase extraction cleanup for semivolatile organic compounds, polycyclic aromatic hydrocarbons (PAH), and alkylated PAH homolog groups in sediment. Anal Chim Acta 549: 104-116.
  39. Sun FS, Littlejohn D, Gibson MD (1998) Ultrasonication extraction and solid phase extraction clean-up for determination of US EPA 16 priority pollutant polycyclic aromatic hydrocarbons in soils by reversed-phase liquid chromatography with ultraviolet absorption detection. Anal Chim Acta 364: 1-11.
  40. Camino-Sánchez FJ, Zafra-Gómez A, Pérez-Trujillo JP, Conde-González JE, Marques JC, et al. (2011) Validation of a GC-MS/MS method for simultaneous determination of 86 persistent organic pollutants in marine sediments by pressurized liquid extraction followed by stir bar sorptive extraction. Chemosphere 84: 869-881.
  41. Bertilsson S, Widenfalk A (2002) Photochemical degradation of PAHs in freshwaters and their impact on bacterial growth - influence of water chemistry. Hydrobiologia 469: 23-32.
  42. Haritash AK, Kaushik CP (2009) Biodegradation aspects of polycyclic aromatic hydrocarbons (PAHs): a review. J Hazard Mater 169: 1-15.
Citation: Crini NM, Druart C, Amiot C, Gimbert F, Chanez E, et al. (2014) Towards a Convenient Procedure to Characterize Polycyclic Aromatic Hydrocarbons in Sediments Receiving Industrial Effluents. J Pollut Eff Cont 3:128

Copyright: © 2014 Crini NM, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top