Vous êtes sur la page 1sur 8

a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44

available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/aca

Analysis of emerging contaminants in sewage effluent and


river water: Comparison between spot and passive sampling

Zulin Zhang, Andrew Hibberd, John L. Zhou ∗


Department of Biology and Environmental Science, School of Life Sciences, University of Sussex,
Falmer, East Sussex, Brighton BN1 9QG, UK

a r t i c l e i n f o a b s t r a c t

Article history: Passive sampling is highly complimentary to spot sampling in environmental analysis. A
Received 30 August 2007 polar organic chemical integrative sampler (POCIS) was extensively tested to optimize its
Received in revised form performance under both controlled and field conditions. The passive sampler was subse-
8 November 2007 quently used for the sampling and analysis of estrone, 17␤-estradiol, 17␣-ethynylestradiol,
Accepted 15 November 2007 bisphenol A, propranolol, sulfamethoxazole, meberverine, thioridazine, carbamazepine,
Published on line 21 November 2007 tamoxifen, indomethacine, diclofenac and meclofenamic acid in sewage effluent and river
water. Under laboratory conditions, the kinetics of compound uptake by POCIS were linear
Keywords: during 10-day of exposure. POCIS sampling rates of the target compounds were signifi-
Passive sampling cantly greater by using polyethersulfone instead of polysulfone membrane, and enhanced
Pharmaceuticals with increasing sorbent exposure area. Together with spot water sampling, the optimized
Endocrine disrupting chemicals POCIS was deployed in the River Ouse, West Sussex, UK to obtain field-derived sampling
Liquid chromatography–tandem rates which, for endocrine disrupting chemicals (EDCs), were significantly higher than those
mass spectrometry from laboratory experiments. Both spot and passive sampling demonstrated that most of
Gas chromatography–mass the target chemicals were frequently detected in sewage effluent and river waters, and that
spectrometry the daily changes in the pollutant concentrations were greater for pharmaceuticals than
for EDCs. The aqueous concentrations of all compounds were elevated at a sewage outfall,
which is confirmed to be an important source of the target compounds in the river. The
validated POCIS was then successfully used to estimate the concentrations of the target
compounds in effluent and river water, which were in good agreement with those from spot
sampling for pharmaceuticals.
© 2007 Elsevier B.V. All rights reserved.

1. Introduction 17␣-ethynylestradiol (EE2) is the main component of the oral


contraceptive pill, and carbamazepine and diclofenac are used
Water pollution from emerging pollutants such as endocrine as anti-epileptic and anti-inflammatory drugs, respectively
disrupting chemicals (EDCs), pharmaceuticals and personal [3–5]. In addition, EDCs can be naturally occurring in the envi-
care products (PPCPs) is one of the important aspects of cur- ronment, e.g. female hormones estrone (E1) and 17␤-estradiol
rent environmental research due to their interference with (E2) are both excreted by women and hence ubiquitous in the
the endocrine system of wildlife and human [1,2]. Most aquatic environment. Such compounds may not be removed
EDCs and PPCPs are man-made organic chemicals being by sewage treatment works (STW) and are reactivated during
introduced to the environment by anthropogenic inputs, e.g. those processes.


Corresponding author. Tel.: +44 1273 877318; fax: +44 1273 678937.
E-mail address: j.zhou@sussex.ac.uk (J.L. Zhou).
0003-2670/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.aca.2007.11.024
38 a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44

Considering the potential impacts of EDCs and PPCPs, they and behavior of emerging contaminants in sewage effluent
should be routinely monitored in aquatic systems. The most and river waters. Robust extraction and analytical methods
widely used technique for performing such monitoring is spot have been developed based on solid-phase extraction (SPE)
sampling followed by laboratory-based extraction and analy- followed by gas chromatography–mass spectrometry (GC–MS)
sis [6–8]. This approach, however, yields only an instantaneous for EDCs [20,21] and liquid chromatography–tandem mass
measurement of pollutant levels and suffers from the uncer- spectrometry (LC–MS–MS) for pharmaceuticals [22]. Labora-
tainty of short- and long-term concentration variations. An tory calibration was performed to evaluate the sampling rate
increase in sampling frequency or the use of flow- and time- of the modified POCIS for the target compounds. The effects
weighted automatic samplers may reduce such uncertainty; of different conditions (e.g. size of sampler) on POCIS sam-
however, the associated increase in costs may prove unfeasi- pling rates were evaluated. The POCIS sampling rates derived
ble [9,10]. from laboratory conditions were compared with those from
There has been rapid development in the use of passive field experiments. Estimation of the ambient aqueous concen-
sampling devices that allow continuous monitoring of aque- tration of target chemicals was achieved by using field-based
ous pollutants without the disadvantages of using organisms sampling rates.
[9–12]. Of the various passive sampling devices, the most
widely used is the semi-permeable membrane device (SPMD)
2. Experimental
consisting of a tubular polyethylene membrane containing
a film of lipids such as triolein [13,14]. In comparison to
traditional water sampling, SPMD can be standardized and 2.1. Materials
deployed over a long period, can detect low levels of organic
contaminants, and mimics the bioconcentration of pollutants The solvents used including methanol and ethyl acetate were
in aquatic organisms [15]. However, SPMD is not suitable of distilled-in-glass grade from Rathburns Chemicals Ltd.
for polar organics, as either the membrane is impermeable (Scotland). Compounds E1, E2, EE2, E2-d2 , propranolol (pro),
to such compounds or accumulation is thermodynamically sulfamethoxazole (sul), meberverine, thioridazine, carba-
unfavorable due to the low affinity of the receiving phase for mazepine (carb), tamoxifen, indomethacine (indo), diclofenac
such analytes [16]. As it is the polar compounds (e.g. EE2) that (diclo) and meclofenamic acid were purchased from Sigma,
are primarily responsible for the estrogenic activities, recent UK, and bisphenol A (BPA) and BPA-d16 were supplied by
development has been focused on their passive sampling, Aldrich (Dorset, UK). Diuron-d6 and 13 C-phenacetin were
e.g. polar organic chemical integrative sampler (POCIS) and purchased from Cambridge Isotope Laboratories, USA. Sep-
Chemcatcher [17,18]. POCIS comprises a solid receiving phase arate stock solutions of individual compounds (1000 mg L−1 )
(sorbent) sandwiched between two microporous polyethersul- were prepared by dissolving an appropriate amount of each
fone (PES) membranes. POCIS samples from water and thereby substance in methanol. From these standards, a mixture
enables the chemical concentration to be estimated as follows of working standards or internal standards (BPA-d16 , E2-
[12,17]: d2 , diuron-d6 , 13 C-phenacetin) containing each compound
at 10 mg L−1 was prepared by diluting the stock solution in
Ms = Cw Rs t (1) methanol. All the standard solutions were stored at −18 ◦ C
prior to use. Ultrapure water was supplied by a Maxima Unit
where Ms is the mass of analytes in the receiving phase at from USF Elga, UK.
time t, Cw represents time-weighted average concentration in The passive sampling device was similar to POCIS [17]
water during the deployment period. Rs is the sampling rate of except the holder which supports both the diffusion-limiting
the system, which may be interpreted as the volume of water membrane and sorbent and seals them in place, was made of
cleared of analyte per unit of exposure time by the device [19]. PTFE rather than stainless steel (Fig. 1). In addition, the device
Chemcatcher uses a polytetrafluoroethylene (PTFE) sup- of different exposure diameters (27, 38, 54 mm) was tested.
port device to protect a layer of membrane which covers a solid The Oasis HLB sorbent (100 mg) from Waters Ltd., UK was
receiving phase (e.g. C18 Empore disk). Accumulation rates chosen because it sorbs a wider range and polar compounds
and selectivity are regulated by the choice of both the mem- better than C18 [20,22]. The Oasis sorbent was spread evenly
brane and the solid receiving material; both are supported between the membranes. The PES membrane (0.1 ␮m pore
and sealed in place by an inert plastic housing [12,18]. Both of size) and polysulfone (PS) membrane (0.2 ␮m pore size) were
these samplers have advantages over traditional methods of
sampling; however, further improvement and validation are
needed. For example, existing calibration experiments were
only performed in small systems, e.g. 1 L microcosms [17].
Certain factors (e.g. pollutant concentration, salinity) which
may affect the sampling rate were not reported [17,18]. Fur-
thermore, in order to estimate water concentration more
accurately, it is essential to validate laboratory-derived with
in situ sampling rates.
This paper focused on the validation and application of
POCIS for emerging contaminants in water, including evalua- Fig. 1 – The schematic diagram of the POCIS device (1) PTEE
tion of spot and passive sampling for studying the occurrence holder, (2) screw, (3) membrane, (4) sorbent and (5) hole.
a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44 39

provided by Pall Gelman Sciences (VMR International, UK). (29 January to 9 February 2007). The samplers were suspended
PTFE used to construct the sampler was from Aquarius Plastics vertically at a water depth of 50–100 cm by attaching to a float-
Ltd., UK. Peristaltic pumps (Watson Marlow 401U/DM2) and ing ball. POCIS fitted with PES membrane and Oasis sorbent
tubes (0.5 mm × 1.6 mm and 3.2 mm × 1.6 mm) for controlling as the receiving phase with an exposure diameter of 54 mm,
the flow-through system were from Fisher, UK. were deployed at each site. Three spot water samples (1 L)
and POCIS were sampled each day. Water samples were fil-
2.2. Oasis sorbing capability for target chemicals tered through pre-combusted GF/F filters (0.7 ␮m), spiked with
100 ng of internal standards and extracted by SPE. The ana-
Before the sorbents were used as the receiving phase of POCIS, lytes were eluted from the cartridges with 10 mL methanol,
it was necessary to determine their sorbing capacity for the which were concentrated to 0.3 mL under N2 , and analyzed
target chemicals. Samples of 50 mL of water spiked with 5 ␮g of by LC-tandem-MS for pharmaceuticals and GC-MS for EDCs
each compound were passed through glass columns contain- [20–22]. The results were used to derive field-based sampling
ing 100 mg sorbent. The concentrations of target compounds rates for POCIS. Based on the field POCIS blanks, the limits of
were selected at levels above typical sequestered chemical quantification (LOQ) for selected target compounds in water,
residues from passive samplers to ensure future applications as calculated from Eq. (1), were from 6 to 487 pg L−1 .
of POCIS. Analysis of permeate from the sorbent beds showed The validated POCIS was deployed in the River Ouse (out-
concentrations of the chemicals below limits of detection fall, upstream, downstream) during 23–27 October 2006, for
(LOD), indicating that target chemicals had been retained by comparing performance with spot sampling. Again water
the sorbents. Recovery of the target compounds from the sor- samples were taken daily for monitoring the target com-
bents was achieved by eluting the sorbent beds with 20 mL pounds, while POCIS was sampled at the end of the week-long
of methanol, followed by concentration under N2 , and GC- experiment. In addition, water properties such as tempera-
MS analysis [20,21]. Good recovery (80–87%) of the target ture, pH and dissolved oxygen at each site were recorded at
compounds was achieved from replicate sampling and anal- each visit.
ysis, confirming that the sorbent has sufficient capacity for
trapping target compounds and can be used in POCIS for mon-
itoring purpose. 3. Results and discussion

2.3. Exposure in a flow-through system 3.1. Flow-through exposure under controlled


conditions
To determine Rs , a flow-through tank allowing the continu-
ous addition of water and target compound solutions was The performance of POCIS was tested by exposure to a
used throughout the experiments. It is hypothesized that relatively constant concentration of target compounds in
laboratory-derived Rs will be different from field-derived Rs , the exposure tank (R.S.D. = 2.1–11.1%). Characteristic analyte
due primarily to different hydrodynamic regimes. Deionized uptake curves for the sampler are shown in Fig. 2, demon-
water and a water solution of the target compounds were strating that the uptake of selected target compounds in POCIS
pumped simultaneously into the exposure tank by the peri- follows a linear pattern. Based on uptake kinetics, the mean
staltic pumps. The tank was always filled with water (30 L) at sampling rate varied from 0.036 L day−1 for BPA to 0.069 L day−1
approximately 15 ◦ C. The rates of addition of distilled water for E1. Alvarez et al. reported that laboratory-derived Rs val-
and target compound solution were controlled at approxi- ues for some other compounds such as diuron, isoproturon,
mately 20 and 1 mL min−1 , respectively. The concentration of azithromycin and omeprazole varied from 0.005 to 0.12 L day−1
analytes in the exposure tank was checked daily (100 mL, n = 3), [17]. Matthiessen et al. [23] estimated a sampling rate from
followed by SPE (Oasis HLB), and instrumental analysis [20,21]. 0.09 to 0.129 L day−1 for E2 when calibrating in a glass beaker
The results show a relatively constant chemical concentration with stirring. Such Rs values are comparable to ours for target
(R.S.D. = 2.1–11.1%) in the exposure tank and a replacement chemicals although they are for different compounds.
time for the system of 23.8 h. Following exposure, the POCIS The concentrations of organic micropollutants in aquatic
(n = 3) was removed and dismantled. The sorbent in POCIS environments are highly variable, both temporally and spa-
was removed, spiked with 100 ng of internal standards, and tially. To investigate the effect of compound concentrations on
then extracted with 10 mL of methanol 3 times. The extracts the sampling rate of POCIS, the experiments were performed
were concentrated to 0.3 mL under nitrogen, and analyzed by with chemical concentrations varying from 10 to 1000 ng L−1 .
GC-MS for EDCs [20,21]. In addition, a minimum of 3 blanks The results (Table 1) show that Rs for each compound was not
for water, membrane and sorbent were analyzed during cal- significantly influenced by compound concentrations, which
ibration experiments to determine the initial levels of target are consistent with other reports that the sampling rates for
compounds in various matrices, the results show that none of SPMD are independent of environmental concentration [24].
the blanks contained the target compounds above LOD. The findings suggest that POCIS can be applied for monitoring
the target compounds of different concentrations.
2.4. Sampler validation and application in the field
3.2. Membrane evaluation
Two sites in the River Ouse, West Sussex, England, includ-
ing the outfall of Sheffield Park STW and its downstream The microporous membrane in POCIS acts as a semi-
(700 m), were chosen for field validation of POCIS for 2 weeks permeable barrier between the sorbent and the surrounding
40 a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44

Fig. 2 – Accumulation of BPA (1.63 ␮g L−1 ), E1 (1.41 ␮g L−1 ), E2 (2.14 ␮g L−1 ) and EE2 (1.37 ␮g L−1 ) from water by the passive
sampler under laboratory conditions.

aquatic environment. It allows polar organic chemicals to tors such as the biodegradability, pore size, strength and
pass through to the sorbent, while particulates with cross- durability.
sectional diameters greater than the membrane pore size will In this study, the performance of POCIS made by PES and
be excluded selectively. Without the membrane, direct contact PS membrane was compared, by calculating the sampling
of these excluded materials with the sorbent can result in a rates of POCIS following exposure. As shown in Fig. 3, the
site-specific bias of apparent contaminant concentrations in sampling rate of the target compounds varied from 0.033 to
the sorbent, reduce uptake due to enhanced biofouling in the 0.048 L day−1 (R.S.D. = 12.5–19.1%) for PES membrane, and from
membrane and sorbent, and potentially interfere with sam- 0.004 to 0.017 L day−1 (R.S.D. = 6.8–13.4%) for PS membrane. As
ple processing and analysis. Kingston et al. [18] suggested the PES membrane showed significantly higher Rs than PS
that PS membrane was suitable for polar organic compounds.
Alvarez et al. [17] however concluded that PES membrane
was preferred to sample polar organic contaminants, after
studying many types of membrane and considering fac-

Table 1 – Sampling rate (Rs (L day−1 )) of POCIS with an


exposure diameter of 27 mm at different EDC
concentrations
Concentration (ng L−1 ) BPA E1 E2 EE2

10 0.047 0.034 0.041 0.047


20 0.040 0.039 0.037 0.065
50 0.053 0.040 0.028 0.049
100 0.028 0.026 0.029 0.046
250 0.038 0.040 0.039 0.053
500 0.042 0.039 0.042 0.051
1000 0.033 0.064 0.045 0.046

Average 0.040 0.040 0.037 0.051


Fig. 3 – Comparison of the sampling rates for POCIS
S.D. 0.008 0.012 0.007 0.007
R.S.D. (%) 21.1 28.4 17.6 13.0
(diameter = 27 mm) using PES and PS membranes, for
compounds BPA, E1, E2 and EE2.
a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44 41

sampling rate was determined following 3-day exposure


in the flow-through system. It was found that Rs did not
vary significantly with changing salinity, with R.S.D. below
12%. Hence the sampler can be applied to many types
of natural waters, such as fresh water, estuarine water or
seawater.

3.5. Occurrence of emerging contaminants in water


measured by spot sampling

The distribution of the target compounds in sewage effluent


and receiving river water was monitored daily, to under-
stand their temporal changes. As shown in Fig. 5, overall,
the mean daily concentrations of BPA, E1 and E2 were
Fig. 4 – Relationship between the sampling rates and the relatively invariable from Monday to Friday, whilst for phar-
exposure surface area of POCIS using PES membrane. maceuticals (propranolol, sulfamethoxazole, carbamazepine,
indomethacine and diclofenac) their concentrations were
more variable, in particular for carbamazepine.
membrane, it was therefore selected as the membrane for For EDCs BPA, E1 and E2, their concentrations were clearly
POCIS for further experiments. elevated at the sewage outfall, suggesting that it is the source
of EDCs in the area. The mean concentrations for E1, E2 and
3.3. Sampler size optimization BPA in the upstream were 9.9, 11.8 and 19.5 ng L−1 , respectively,
which are lower than those (up to 158 ng L−1 of BPA) found in
The kinetics of trapping the target compounds in passive sam- New Orleans, LA, USA [26]. Similarly Kolpin et al. [27] reported
plers are expected to be dependent on the size, and perhaps E2 to be present at average concentrations of 9 ng L−1 in US
more importantly, the exposure surface area of such devices. surface water.
POCIS with three different exposure areas of 5.72, 11.33 and For pharmaceuticals, propranolol, sulfamethoxazole, car-
22.89 cm2 , respectively, was examined for the uptake of tar- bamazepine, indomethacine and diclofenac were regularly
get compounds. The relationship between the sampling rate detected in water samples, while meberverine, thioridazine,
and sampler exposure surface area (Fig. 4) shows that Rs tamoxifen and meclofenamic acid were lower than their LOD
had a positive relationship with the exposure surface area in all water samples. Carbamazepine is one of the most fre-
of the sampler, with the correlation coefficient (r2 ) of the quently detected compounds with the highest concentration
regression from 0.82 (EE2) to 1.0 (BPA). The trend is similar of 652 ng L−1 from sewage effluent. Similarly up to 1075 ng L−1
to what was described by Huckins et al. [24] and Vrana et al. of carbamazepine was found in Berlin waters [28], and even
[25] that Rs was proportional to the surface area of the sam- higher concentrations of up to 7100 ng L−1 were detected in the
pling device. As a result, POCIS with a sorbent diameter of River Elbe and its tributaries [29]. Wiegel et al. [29] suggested
54 mm and a sampling surface area to sorbent mass ratio of that such levels were due to the high persistency of carba-
229 cm2 g−1 was chosen as the optimum for further exposure mazepine, which was therefore chosen as an tracer substance
experiments. for pharmaceutical agents. For the other compounds, the lev-
els of propranolol (1.7–85.2 ng L−1 ) in this study were similar
3.4. Effects of environmental condition on sampling to those (10–93 ng L−1 ) in Canada [30]. The concentrations of
rate diclofenac (1.5–91 ng L−1 ) in the Ouse were lower than those
(<20–599 ng L−1 ) detected by Ashton et al. [3] in southeast UK.
Generally, the speciation of weakly acidic compounds in For sulfamethoxazole, its concentrations (12.4–25.1 ng L−1 ) in
aqueous solutions depends on the solution pH. A series of the STW effluent were also lower than those (243–871 ng L−1 )
experiments for investigation of pH effect on sampling rate in Canadian cites [30].
were performed, and the results show that Rs for target com-
pounds remains relatively similar among pH 4–10, with R.S.D. 3.6. Validation of POCIS by field trials
less than 5%. The pKa values of the chemicals are all higher
than 10, varying from 10.2 for BPA to 10.5 for EE2. At pH As field environment is very different from laboratory con-
value less than 10, the test chemicals will stay predominantly ditions, POCIS performance should be validated in situ. To
as neutral molecules. At pH 10, between 32 and 63% of the understand such differences, POCIS were deployed in the field
compounds will be in dissociated form. The findings there- to determine Rs values in situ, where target compounds were
fore suggest that both neutral and ionized forms of the target measured simultaneously in water and POCIS. As shown in
chemicals can be accumulated in POCIS. Fig. 6a, field-derived sampling rates for EDCs were signifi-
Natural waters can have different salinity. It is well cantly greater than those from laboratory experiments, which
known that the aqueous solubility of many organic com- is due to significantly high water flow and associated water
pounds decreases with increasing salt concentration, thus turbulence in the field, hence increasing the mass transfer of
their absorption efficiency in the sorbent of POCIS may EDCs from water to POCIS. During laboratory-based experi-
increase. The effect of water salinity (0, 18‰, 35‰) on POCIS ments, relatively small flow rates were used with no mixing,
42 a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44

Fig. 6 – (a) Comparison of the POCIS (diameter = 54 mm)


sampling rates for BPA, E1, E2 between field and
laboratory-derived data; (b) sampling rates for
pharmaceuticals derived from field data. DW: downstream
river water; EW: STW effluent; Lab: laboratory-derived
sampling rates.

compound this difference could be several-fold between the


two conditions [17,19,32,33]. The results therefore suggest the
need for appropriate field validation of such sampling device.
Under field conditions, the sampling rates for BPA, E1 and E2
were higher at downstream site than those from STW efflu-
ent; whilst for pharmaceuticals, the difference between sites
was less clear-cut (Fig. 6b).

3.7. Comparison between spot and passive sampling


Fig. 5 – Daily variations of emerging contaminant
concentrations in sewage effluent and receiving river water The POCIS was deployed and sampled daily at three sites
in the River Ouse, West Sussex, UK between 23–27 October in River Ouse for a period of 5 days. Following sam-
2006. pling, the target compounds in POCIS and water were
determined. The field-derived sampling rates for the tar-
get chemicals in POCIS were used in the calculation of
which explain the lower sampling rates than those from the the predicted ambient water concentrations by Eq. (1). The
field, where factors such as fluctuations in water temperature sampling rates derived from downstream river water were
and in particular water turbulence would have increased the applied to estimate pollutant concentration in water at
uptake kinetics of the compounds in POCIS [17,18,31]. It has upstream and downstream sites, while Rs derived from
been reported that Rs from turbulent environments will be effluent was used to predict the compound concentration
much higher than that from quiescent conditions; for some in the effluent. Then the predicted compound concentra-
a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44 43

Fig. 8 – Correlation between spot sampling and POCIS


sampling for all compounds from all sites including
upstream (UW), STW effluent (EW) and downstream (DW).

closer. For the pharmaceuticals their predicted concentrations


from POCIS are similar to those by spot sampling for most
measurements. The mean aqueous concentrations measured
by spot sampling for propranolol, sulfamethoxazole, carba-
mazepine, indomethacine and diclofenac, varied between 3.0
and 45.6 ng L−1 , <LOD and 17.6 ng L−1 , 16.6 and 539 ng L−1 ,
0.4 and 7.2 ng L−1 and 2.4 and 65.2 ng L−1 , respectively; while
their concentrations predicted by POCIS were between 2.8 and
40.5 ng L−1 , <LOD and 18.2 ng L−1 , 26.3 and 427 ng L−1 , 0.5 and
11.9 ng L−1 and 4.4 and 165 ng L−1 , respectively. It is appar-
ent (Fig. 8) that overall, for all compounds from all sites, the
predicted concentrations from POCIS are lower than those
by spot sampling. For carbamazepine in the upstream and
diclofenac in the effluent, the predicted and measured values
differ significantly. When using POCIS in the River Ravens-
bourne and the Thames Tideway in England, Alvarez et al.
[17] showed that the predicted ambient water concentrations
from POCIS for diuron were much higher than those from
spot sampling while for isoproturon an opposite trend was
observed. They suggested that field validation of such sam-
pling device was needed, which has been achieved in this
Fig. 7 – Comparison of the mean (n = 15, 5-day) emerging
study.
contaminant concentrations from spot water sampling with
those predicted by POCIS (diameter = 54 mm) using
field-derived Rs values at three sites in the River Ouse, West
4. Conclusions
Sussex, UK.

The monitoring of emerging pollutants in the aquatic envi-


ronment has been realized by two approaches: spot sampling
tions in water were compared with those measured by spot and time-weighted average passive sampling. As demon-
sampling. strated in this study, passive sampling device should be
As shown in Fig. 7, the concentrations of BPA, E1 and E2 thoroughly assessed both in the laboratory and in the field,
predicted from POCIS are lower than those by spot sampling as its performance can differ between the two environments.
except for E1 in the effluent site where the two values are much There is a good agreement between pharmaceutical con-
44 a n a l y t i c a c h i m i c a a c t a 6 0 7 ( 2 0 0 8 ) 37–44

centrations obtained using spot sampling and those from [13] J.N. Huckins, G.K. Manuweera, J.D. Petty, D. Mackay, J.A. Lebo,
passive sampling which has been validated in situ, high- Environ. Sci. Technol. 27 (1993) 2489–2496.
lighting the potential benefits of using passive sampling for [14] J.D. Petty, J.N. Huckins, C.E. Orazio, J.A. Lebo, B.C. Poulton,
R.W. Gale, C.S. Charbonneau, E.M. Kaiser, Environ. Sci.
water quality monitoring (e.g. low maintenance, high repro-
Technol. 29 (1995) 2561–2566.
ducibility). However, the agreement between the two sets [15] J.D. Petty, B.C. Poulton, C.S. Charbonneau, J.N. Huckins, S.B.
of values for EDCs is not satisfactory, and further work is Jones, J.T. Cameron, H.F. Prest, Environ. Sci. Technol. 32
needed to improve its performance as a complimentary tech- (1998) 837–842.
nique to spot sampling for the integrative analysis of emerging [16] J.D. Petty, J.N. Huckins, D.A. Alvarez, W.G. Brumbaugh, W.L.
pollutants. Granor, R.W. Gale, A.C. Rastall, T.L. Jones-Lepp, T.J. Leiker,
C.E. Rostad, E.T. Furlong, Chemosphere 54 (2004) 695–705.
[17] D.A. Alvarez, J.D. Petty, J.N. Huckins, T.L. Jones-Lepp, Environ.
Acknowledgements Toxicol. Chem. 23 (2004) 1640–1648.
[18] J.K. Kingston, R. Greenwood, G.A. Mills, G.M. Morrison, L.B.J.
Persson, J. Environ. Monit. 2 (2000) 487–495.
The project was funded by the 6th Framework Programme of
[19] B. Vrana, G.A. Mills, E. Dominiak, R. Greenwood, Environ.
the European Commission (MIF1-CT-2004-510012) and an EU Pollut. 142 (2006) 333–343.
Interreg grant (162/039/096). [20] R. Liu, J.L. Zhou, A. Wilding, J. Chromatogr. A 1022 (2004)
179–189.
[21] Z.L. Zhang, A. Hibberd, J.L. Zhou, Anal. Chim. Acta 577 (2006)
Appendix A. Supplementary data 52–61.
[22] Z.L. Zhang, J.L. Zhou, J. Chromatogr. A 1154 (2007) 205–213.
Supplementary data associated with this article can be found, [23] P. Matthiessen, D. Arnold, A.C. Johnson, T.J. Pepper, T.G.
in the online version, at doi:10.1016/j.aca.2007.11.024. Pottinger, K.G.T. Pulman, Sci. Total Environ. 367 (2006)
616–630.
[24] J.N. Huckins, J. Petty, H. Prest, R. Clark, D. Alvarez, C. Orazio,
references J. Lebo, W. Cranor, B. Johnson, A Guide for the Use of
Semipermeable Membrane Devices (SPMDs) as Samplers of
Waterborne Hydrophobic Organic Contaminants. API
[1] J.P. Sumpter, Acta Hydrochim. Hydrobiol. 33 (2005) 9–16. Publication 4690, American Petroleum Institute,
[2] B.H. Sorensen, S.N. Nielsen, P.F. Lanzky, F. Ingerslev, H.C.H. Washington, DC, 2002.
Lutzhoft, S.E. Jorgensen, Chemosphere 36 (1998) 357–393. [25] B. Vrana, G.A. Mills, R. Greenwood, J. Knutsson, K. Svensson,
[3] D. Ashton, M. Hilton, K.V. Thomas, Sci. Total Environ. 333 G. Morrison, J. Environ. Monit. 7 (2005) 612–620.
(2004) 167–184. [26] G.R. Boyd, J.M. Palmeri, S.Y. Zhang, D.A. Grimm, Sci. Total
[4] S.D. Richardson, Anal. Chem. 76 (2004) 3337–3364. Environ. 333 (2004) 137–148.
[5] M. Petrović, M.D. Hernando, M.S. Diaz-Cruz, D. Barceló, J. [27] D.W. Kolpin, E.T. Furlong, M.T. Meyer, E.M. Thurman, S.D.
Chromatogr. A 1067 (2005) 1–14. Zaug, L.B. Barber, H.T. Buxton, Environ. Sci. Technol. 36
[6] R. Jeannot, H. Sabik, E. Sauvard, T. Dagnac, K. Dohrendorf, J. (2002) 1202–1211.
Chromatogr. A 974 (2002) 143–159. [28] T. Herberer, Toxicol. Lett. 131 (2002) 5–17.
[7] S.D. Richardson, T.A. Ternes, Anal. Chem. 77 (2005) [29] S. Wiegel, A. Aulinger, R. Brockmeyer, H. Harms, J. Loffler, H.
3807–3838. Reincke, R. Schmidt, B. Stachel, W.V. Tumpling, A. Wanke,
[8] T.A. Ternes, M. Stumpf, J. Mueller, K. Haberer, R.D. Wilken, Chemosphere 57 (2004) 107–126.
M. Servos, Sci. Total Environ. 225 (1999) 81–90. [30] X.S. Miao, F. Bishay, M. Chen, C.D. Metcalfe, Environ. Sci.
[9] A. Kot, B. Zabiegala, J. Namiesnik, Trends Anal. Chem. 19 Technol. 38 (2004) 3533–3541.
(2000) 446–459. [31] G.S. Ellis, J.N. Huckins, C.E. Rostad, C.J. Schmitt, J.D. Petty, P.
[10] T. Gorecki, J. Namiesnik, Trends Anal. Chem. 21 (2002) Maccarthy, Environ. Sci. Technol. 14 (1995) 1875–1884.
276–291. [32] K. Booij, H.M. Sleiderink, F. Smedes, Environ. Toxicol. Chem.
[11] B. Vrana, P. Popp, A. Paschke, G. Schuurmann, Anal. Chem. 17 (1998) 1236–1245.
73 (2001) 5191–5200. [33] B. Vrana, G. Schuurmann, Environ. Sci. Technol. 36 (2002)
[12] B. Vrana, G.A. Mills, I.J. Allan, E. Dominiak, K. Svensson, J. 290–296.
Knutsson, G. Morrison, R. Greenwood, Trends Anal. Chem.
24 (2005) 845–868.

Vous aimerez peut-être aussi