Sei sulla pagina 1di 9

Science of the Total Environment 601–602 (2017) 812–820

Contents lists available at ScienceDirect

Science of the Total Environment

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

Bioaccumulation of pharmaceutically active compounds and endocrine


disrupting chemicals in aquatic macrophytes: Results of hydroponic
experiments with Echinodorus horemanii and Eichhornia crassipes
N. Pi, J.Z. Ng, B.C. Kelly ⁎,1
Department of Civil and Environmental Engineering, National University of Singapore, 5A Engineering Drive 1, 117411, Singapore

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• PhACs and EDCs were readily taken up


E. crassipes and E. horemanii.
• Diphenhydramine and triclosan exhibit-
ed the highest degree of bioaccumula-
tion.
• BPA, E1, E2 and warfarin exhibited rela-
tively low bioaccumulation potential.
• No clear relationship was observed be-
tween physicochemical properties and
bioaccumulation behaviour.

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

Article history: Information regarding the bioaccumulation behaviour of pharmaceutically active compounds (PhACs) and endo-
Received 7 March 2017 crine disrupting chemicals (EDCs) in aquatic plants is limited. The present study involved controlled hydroponic ex-
Received in revised form 14 May 2017 periments to assess uptake and elimination rate constants (ku, ke), bioconcentration factors (BCFs) and translocation
Accepted 15 May 2017
factors (TFs) of several PhACs and EDCs in two aquatic macrophyte species, including one submerged species
Available online xxxx
(Echinodorus horemanii) and one free-floating species (Eichhornia crassipes). The results revealed that the studied
Editor: Jay Gan compounds are readily taken up in these aquatic plants. While bioconcentration factors (BCFs) and translocation
factors (TFs) of the test compounds varied substantially, no discernible relationship with physicochemical proper-
Keywords: ties such as octanol-water distribution coefficient (Dow), membrane-water distribution coefficient (Dmw) and or-
Pharmaceutically active compounds ganic carbon-water partition coefficient (Koc). Diphenhydramine and triclosan exhibited the highest degree of
Endocrine disrupting chemicals uptake and bioaccumulation potential. For example, the whole-plant BCF of triclosan in E. horemanii was
Aquatic macrophytes 4390 L/kg, while the whole-plant BCF of diphenhydramine in E. crassipes was 6130 L/kg. BCFs of 17β-estradiol
Bioaccumulation factors (E2), 17α-ethinylestradiol (EE2), estrone (E1) and bisphenol A (BPA) were relatively low (2–150 L/kg). BCFs
Translocation factors
were generally higher in free-floating aquatic macrophyte species compared to the submerged species. For the
free-floating species, E. crassipes, the majority of PhACs and EDCs were more allocated in roots compared to leaves,
with TFs b 1. However, some compounds such as caffeine, atrazine, diphenhydramine, E2 and carbamazepine were
more allocated in leaf tissue (TFs N 1). The study findings may be useful for design and implementation of
phytoremediation systems, as well as aid future modeling and risk assessment initiatives for these emerging organic
contaminants.
© 2017 Elsevier B.V. All rights reserved.

⁎ Corresponding author.
E-mail address: bckelly@nus.edu.sg (B.C. Kelly).
1
Department of Civil and Environmental Engineering, National University of Singapore, Block E1A, #07-03, No.1 Engineering Drive 2, 117576, Singapore.

http://dx.doi.org/10.1016/j.scitotenv.2017.05.137
0048-9697/© 2017 Elsevier B.V. All rights reserved.
N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820 813

1. Introduction 2. Material and methods

Emerging organic contaminants (EOCs) such as pharmaceutically 2.1. Chemicals


active compounds (PhACs) and endocrine disrupting chemicals
(EDCs) are widely used and produced substances, with annual produc- Nineteen PhACs and EDCs were examined in this study (Table 1, and
tion volumes typically in the kiloton to megaton range (EC, 2011; WHO, Table S1, Supporting information), due to their wide distribution in
2012; USEPA, 2014). PhACs can enter environment as the parent phar- aquatic environment detected in organisms at different trophic levels
maceutical compound or pharmacologically active metabolites (Bayen et al., 2013; Crane et al., 2006; Huerta et al., 2016; Jayasiri
(Halling-Sorensen et al., 1998). Aquatic environments are considered et al., 2013; Kannan et al., 2005; Nakata, 2005). High purity chemicals
to be a major sink for PhACs and EDCs, since these chemicals are not of the native compounds were purchased. Atenolol, caffeine, linuron
fully removed by conventional wastewater treatment technologies and sulfamethoxazole were obtained from Wako Pure Chemical
(Petrie et al., 2013). (Wako Pure Chemical Inc., Japan), atrazine, bisphenol A (BPA),
Due to high polarity and low volatility, most PhACs and EDCs tend to carbamazapine, diclofenac, dilantin, diphenhydramine, estrone (E1),
be easily transported in aquatic environments (Breton and Boxall, 17β-estradiol (E2), 17α-ethinylestradiol (EE2), gemfibrozil, ibuprofen,
2003). More than eighty PhACs/EDCs have been detected in aquatic eco- naproxen, thiabendazole, triclosan and warfarin were supplied by
systems (Crane et al., 2006; Huerta et al., 2016; Jayasiri et al., 2013), Sigma-Aldrich (Sigma-Aldrich, Saint Louis Mo, USA). Isotopically la-
some of which are known to be relatively persistent, such as carbamaz- beled compounds were purchased and used as the internal standards.
epine and ibuprofen (Crane et al., 2006; Monteiro and Boxall, 2009). Atenolol-d7 (N99%), atrazine-d5 (N 99%), carbamazepine-d10 (N99%),
PhACs and EDCs and/or their metabolites, upon entering the water en- EE2-d4 (N 99%), ibuprofen-d3 (N99%), 13C12-triclosan (N98%) and
vironment, can pose potential ecological and human health risks, even warfarin-d5 (N99%) were obtained from CDN Isotopes (C/D/N Isotopes
at trace levels ranging from ng/L to μg/L (Carlsson et al., 2006; Inc., Canada), bisphenol A-d6 (98%), 13C3-caffeine (99%), E2-d4 (95–
Corcoran et al., 2010; Daughton and Ternes, 1999; la Farre et al., 97%), gemfibrozil-d6 (N 98%), 13C6-sulfamethoxazole (N98%) and 13C6-
2008). Aquatic organisms within wastewater impacted surface waters thiabendazole (N98%) were supplied by Cambridge Isotope Laboratory
are particularly susceptible to PhACs and EDCs (Garcia-Rodriguez (Cambridge Isotope Laboratory, Inc., USA). HPLC-grade solvents were
et al., 2014; Morley, 2009). obtained from Fisher Scientific (UK) and Tedia (USA).
In recent years, aquatic plant-based phytoremediation systems
such as constructed wetlands have been proposed for the removal
of various waterborne contaminants (Matamoros et al., 2008; 2.2. Hydroponic experiments
Zhang et al., 2014). Aquatic plants play a significant role in removing
waterborne contaminants, via direct adsorption and/or absorption The hydroponic experiments were conducted at Singapore's Van
(Dordio and Carvalho, 2013; Li et al., 2014; Pi et al., 2011). Water- Kleef Aquatic Science Centre, which offers a covered outdoor environ-
borne contaminants can be taken up and translocated (Felizeter ment, allowing good growing conditions avoiding rain-water interfer-
et al., 2012), accumulated (Felizeter et al., 2012; Kannan et al., ence. Nine plastic tanks, each had a dimension of 28 × 18 × 18 cm
2005; Nakata, 2005) and/or degraded via the metabolic transforma- (total volume was 9072 cm3), were prepared and divided into three
tion (Li et al., 2014). Uptake and translocation of organic contami- sets, E. crassipes, E. horemanii and control, with triplicates in each set.
nants in the whole plant system are highly dependent on the plant Thirty E. crassipes and E. horemanii each (previously grown for
species (Garcia-Rodriguez et al., 2014; Zarate et al., 2012), contami- 3 months), were bought from Green Park Tropical Fish Farm and
nant concentrations (Dordio and Carvalho, 2013; Garcia-Rodriguez allowed to acclimate in the tanks for about one week, with 10 seedlings
et al., 2014), and physicochemical properties of the contaminants, per tank, the remaining tanks engaged as control were no plants pres-
including hydrophobicity (Garcia-Rodriguez et al., 2014), polarity ent. Each tank was fully covered with aluminum foil on the sides to pre-
(Dordio and Carvalho, 2013; Garcia-Rodriguez et al., 2014) and vent the photo degradation of PhACs and EDCs, and filled with 6 L of tap
water solubility (Dordio and Carvalho, 2013; Garcia-Rodriguez water, which pH value was adjusted to 7.0 ± 0.05 using hydrochloric
et al., 2014). acid and was previously passed through the column containing activat-
Previous studies have demonstrated that aquatic macrophytes can ed carbon to remove chlorine. The water was renewed every 2 days for
be effective at accumulating and/or removing a variety of waterborne another one week, to allow the plants acclimatized to new condition.
contaminants, including nutrients (Pi et al., 2011), heavy metals (Pi After acclimation, the hydroponic experiment was started with ex-
et al., 2011), and organic compounds such as perfluoroalkyl acids posure and depuration phase, from day 0 to 14 and day 15 to 28, respec-
(Felizeter et al., 2012). However, studies of PhACs and EDCs bioaccumu- tively. In exposure phase, the mixture of native PhACs and EDCs listed in
lation in aquatic macrophytes are limited (Li et al., 2014). Several stud- Table S1 were spiked into the tap water, with amount of each com-
ies have reported that the presence of plants in constructed wetlands pound in each tank at 120 μg and the final concentration was 20 μg/L.
played a positive role in the removal of various PhACs, including To keep the concentration constant at 20 μg/L over the exposure
diclofenac, ibuprofen, ketoprofen, naproxen, salicylic acid, amoxicillin, phase, the water with PhACs and EDCs was renewed every 2 days.
ampicillin, erythromycin, sulfadiazine, sulfamethazine, sulfamethoxa- From day 15 onwards, water was still renewed every 2 days but without
zole, atenolol, clofibric acid, carbamazepine and caffeine chemicals spiking, to keep the concentration of PhACs and EDCs in each
(Hijosa-Valsero et al., 2016; Li et al., 2014; Matamoros et al., 2012; tank constant at 0.
Zhang et al., 2012a; Zhang et al., 2012b). However, these studies have Before starting the exposure experiment, the possible decay of target
mainly focused on assessing contaminant removal efficiencies rather pollutants in water was considered and tested in the present study.
than chemical bioaccumulation behaviour (Hijosa-Valsero et al., 2016; Three plastic tanks, filled with the same medium (a mixture of test com-
Matamoros et al., 2012; Zhang et al., 2012a; Zhang et al., 2012b). pounds with amount of 120 μg each were spiked into the tank contain-
The objective of the present study was to conduct controlled hydro- ing 6 L of tap water, giving a final concentration of 20 μg/L for individual
ponic experiments to assess uptake and elimination kinetics, transloca- PhACs and EDCs) as that used for the exposure experiment but without
tion and bioaccumulation behaviour of several PhACs and EDCs in plants, were prepared and covered with aluminum foil on the sides, to
aquatic macrophytes. Two aquatic macrophytes species were studied, preclude the photodegradation of target compounds. The water concen-
including a free-floating species (Eichhornia crassipes (Mart.) Solms) trations of target compounds in each tank were then analyzed immedi-
and a submerged species (Echinodorus horemanii Rataj). The target ately (Cw0) and two days later (Cw2). The difference between Cw0 and
PhACs and EDCs exhibited a wide range of physicochemical properties. Cw2 for all the target compounds was lower than 5%.
814 N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820

Table 1
List of target compounds and corresponding physicochemical properties.

Compound Formula MW Group Class/Use Ionizable functional pKab % ionized Water Log Log Log Log
group molecules solubility Kow,Nb Kocb Dowc Dmwd
(pH 7) (mg/L)b (pH 7) (pH 7)

Atenolol C14H22N2O3 266.3 PhACs Beta blocker Basic 9.6 99.7 9.54 × 105 0.2 1.8 −2.3 −0.2
Atrazine C8H14ClN5 215.7 EDCs Pesticide Basic 1.6 0.00 214.0 2.6 2.4 2.6 2.7
BPA C15H16O2 228.3 EDCs Plasticizer Acidic 9.5 0.05 172.7 3.3 4.6 3.3 3.5
Caffeine C8H10N4O2 194.2 PhACs Stimulant Acidic 0.6 86.3 11,308 −0.1 1.0 -1.0 1.2
Carbamazapine C15H12N2O 236.3 PhACs Antiepileptic Neutral 15.4 0.00 112.0 2.4 3.1 2.4 2.5
Diclofenac C14H11Cl2NO2 296.2 PhACs NSAIDa Acidic 4.0 99.9 4.52 4.5 2.7 1.8 2.7
Dilantin C15H12N2O2 252.3 PhACs Antiepileptic Acidic 8.3 4.77 2.15 2.5 3.1 2.5 2.6
Diphenhydramine C17H21NO 255.4 PhACs Antihistamine Basic 9.0 99.0 362.7 3.3 3.9 1.3 2.2
E1 C18H22O2 270.4 EDCs Steroid estrogen Acidic 10.8 0.02 19.9 3.1 4.4 3.1 3.3
E2 C18H24O2 272.4 EDCs Steroid estrogen Acidic 10.8 0.02 50.1 4.0 4.2 4.0 4.2
EE2 C20H24O2 296.4 EDCs Synthetic Acidic 10.4 0.04 13.2 3.7 4.7 3.7 3.9
estrogen
Gemfibrozil C15H22O3 250.3 PhACs Lipid regulator Acidic 4.7 99.7 8.42 4.8 2.6 2.4 3.1
Ibuprofen C13H18O2 206.3 PhACs NSAIDa Acidic 4.4 98.4 41.1 4.0 2.6 2.2 2.5
Linuron C9H10Cl2N2O2 249.1 EDCs Pesticide Neutral 12.1 0.00 44.3 3.2 2.5 3.2 3.4
Naproxen C14H14O3 230.3 PhACs NSAIDa Acidic 4.2 99.9 44.1 3.2 2.5 0.5 1.4
Sulfamethoxazole C10H11N3O3S 253.3 PhACs Antibiotic Acidic 6.1 88.8 869.5 0.9 2.4 −0.05 0.1
Thiabendazole C10H7N3S 201.3 EDCs Pesticide Basic 4.7 0.50 482.1 2.5 3.6 2.5 2.6
Triclosan C12H7Cl3O2 289.5 PhACs Antibacterial Acidic 7.8 13.7 9.29 4.8 4.4 4.7 4.9
Warfarin C19H16O4 308.3 PhACs Anticoagulant Acidic 5.0 99.0 4.09 2.7 2.6 0.7 2.1
a
NSAID: nonsteroidal anti-inflammatory drugs.
b
Compiled from US EPA's Estimation Programs Interface (EPI) Suite™.
c
Octanol-water distribution coefficient (Dow) at pH 7 was determined as Dow (pH 7) = fN × Kow, N + fI × Kow, I, where fN and fI are fraction of neutral and ionic species at pH 7, re-
spectively, as predicted from the Henderson-Hasselbach equation; Kow, N and Kow, I are the octanol-water partition coefficients (Kow) of the neutral and ionic species, respectively. Kow, I was
determined by assuming Kow, N was approximately 3 log units higher (i.e., Δow = 3.1).
d
Membrane-water distribution coefficient (Dmw) at pH 7 was determined from liposome-water partition coefficients following Armitage et al. (2013). Specifically, the liposome-water
partition coefficient for neutral species was determined as Log Kmw, N = 1.01 × Log Kow + 0.12, the liposome-water partition coefficient for ionic species was determined as Log Kmw, I
= Log Kmw,N − Δmw. Subsequently, Dmw was determined as, Dmw (pH 7) = fN × Kmw, N + fI × Kmw, I where fN is the fraction of chemical in neutral form and fI is the fraction of chemical
in charged form at pH 7, as predicted by the Henderson-Hasselbalch equation.

During the whole experiment, one seedling of each plant species For macrophytes, extraction and cleanup were based on previous
was randomly collected from each tank every 4 days, washed and proc- methods for PhACs and EDCs in biota (Bayen et al., 2013; Holling
essed into leaves and roots. After being weighed, the leaves and roots et al., 2012; Metcalfe et al., 2010; Vanderford and Snyder, 2006),
samples were stored at − 20 °C before chemical analysis. Part of the USEPA method 1694 (USEPA, 2007)), with some modifications and
leave and root samples were dried in an oven at 70 °C for 48 to 72 h the actual method used in present study was described below. Specifi-
and the oven-dried weight of plant samples were measured, to obtain cally, 5 g of wet plant sample were weighed, grinded to dryness with an-
the plant dry biomass. Water samples were collected on day 0, 2, 4, 14 hydrous sodium sulfate, and transferred to 50 mL polypropylene tube.
(the day 0 of the depuration phase), 16 and 18, just before and immedi- 20 mL of methanol (HPLC grade, fisher, UK) was added to each sample
ately after water renewal, to determine chemical concentrations in tube and the mixture of labeled PhACs and EDCs was spiked, with
water. amount of each compound listed in Table S2. The contents of each
tube were then mixed with a vortex for 20 s, prior to the ultrasonic-
2.3. Extraction and cleanup of water and macrophyte samples extraction operated in an ultrasonic cleaner (Elma, S 60 H, Alphasonics
Pte., Ltd.) at room temperature for 15 min. The extract was centrifuged
Water samples were processed based on the previous studies at 1100 gravity for 5 min and the supernatant was then decanted to a
(Bayen et al., 2013; Togola and Budzinski, 2008; Wu et al., 2010b) and round bottom flask. The procedures were repeated for three times. In
USEPA method 1694 (USEPA, 2007), with some modifications in pres- order to increase the extraction efficiency of triclosan, acetone was
ent study and the actual method used was described below in detail. used at the second round instead of methanol, as acetone is more effec-
Water samples with volume of 500 mL each were filtered through a tive than methanol to extract triclosan from biota samples (USEPA,
glass fiber filter (diameter 47 mm, particle retention 1 μm) and the pH 2007). The extract was combined and the final 60 mL of extract in
was adjusted to 7.0 ± 0.05 using 1 M of hydrochloric acid. Prior to SPE each flask was then concentrated to approximately 2 mL using a rotary
extraction, the mixture of labeled PhACs and EDCs was spiked, with evaporator at 120 rpm and a temperature of 40 °C. Then 50 mL of Milli-Q
amount of each compound listed in Table S2, and equilibrated in the water was added to each flask, mixed and prepared for the clean-up.
dark at 4 °C for about 14–18 h. Samples collected during exposure The pH of the mixture was adjusted to 7.0 ± 0.05 using diluted hydro-
phase (day 0, 2 and 4) were then directly injected into LC/MS/MS for chloric acid at concentration of 1 M.
chemical analysis. Samples taken during depuration phase (day 14, 16 Solid Phase Extraction (SPE) clean-up was performed using HLB car-
and 18) were further extracted using Hydrophilic Lipophilic Balance tridges (200 mg, 6 mL, Phenomenex). Cartridges were pre-conditioned
(HLB) cartridges (60 mg, 3 mL, Phenomenex), pre-conditioned with by passage of 5 mL of methanol, followed by 5 mL of Milli-Q water and
5 mL of methanol, 5 mL of Milli-Q water and 5 mL of Milli-Q water 5 mL of Milli-Q water with pH of 7.0. Then samples obtained above were
with pH at 7.0. After the samples were loaded, the cartridges were passed through the pre-conditioned cartridges at a rate of 1 drop/s.
washed with 5 mL of MilliQ water, and vacuum-dried for 15 min. After all the samples were loaded, the cartridge was washed with
Analytes were eluted with 12 mL of methanol followed by 6 mL of ace- 5 mL of Milli-Q water and vacuum-dried for approximately 15 min.
tone and methanol (1:1 v/v). Combined elution was concentrated to a The cartridges were then eluted with 12 mL of methanol, followed by
final volume of 0.5 mL under nitrogen stream, transferred to 2 mL of 6 mL of acetone and methanol (1:1 v/v). The combined elution was con-
GC vial, and spiked with 50 ng of 13C6-TCPAA and Atrazine-d5. The pre- centrated under a gentle stream of nitrogen at 5 bar and 35 °C, to a final
pared samples were stored at −20 °C before LC/MS/MS analysis. volume of 0.5 mL. The resulting solution was transferred to 2 mL of GC
N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820 815

vial and 50 ng of instrument internal standards (13C6-TCPAA and Uptake and elimination rate constants (ku, ke) of target compounds
Atrazine-d5) were spiked to each vial. The vials were then stored at were determined using their observed concentrations in plant tissues
− 20 °C before sending to liquid-chromatography electrospray (Cp) during exposure and depuration phase, respectively:
ionization-tandem mass spectrometry (LC-ESI-MS/MS) for the chemical
analysis. C P ¼ C 0 e−ke t d ð2Þ

2.4. Instrumental analysis using LC-ESI-MS/MS " #


ku h i
CP ¼ Cw   1−e−ðke þkg Þt u ð3Þ
A Shimadzu LCMS-8030 Tandem Quadruple Liquid Chromatograph ke þ kg
Mass Spectrometer (LC/MS/MS) was used for the identification and
quantification of target analytes. A Poroshell 120 Stable bond SB-C18 where Co is the initial chemical concentration in plant at the start of
column (2.1 × 150 mm, 2.7-micron) preceded by a Poroshell 120 Stable depuration, Cw is the chemical concentrations in water, td is the time
bond SB-C18 guard column (2.1 × 5 mm, 2.7-micron) was selected for measured from the start of depuration phase, and tu is the time mea-
analysis. The mobile phase comprised 5 mM of ammonium acetate in sured from the start of the experiment until the end of exposure phase.
Milli-Q water (A) and ACN/MeOH (1:1 v/v, B) and was delivered at a Kinetically-derived bioconcentration factors (BCFk) values were de-
flow rate of 300 μL/min. The injection volume was 10 μL. The mass spec- termined as ku/(ke + kg). Apparent steady-state bioconcentration fac-
trometer was operated in multiple reaction monitoring (MRM) mode, tors (BCFss) were determined as the ratio of the chemical
with some compounds in positive ion and the others in negative ion concentration observed in plants (Cp) and water (Cw) at the end of
(Table S1, Supporting Information). The linear gradient program for the exposure phase (i.e., Cp/Cw). Translocation factors (TFs) were deter-
positive ion mode was started with mobile phase B at 10%, and in- mined as the ratio of the chemical concentration (ng/g dry weight) ob-
creased to 40% at 4 min, to 100% at 10 min, and kept at 100% until served in leaf and roots. Separate TF values were determined for the
19 min, then decreased to 10% at 20 min and kept until 22 min. The lin- exposure and depuration phase on day 14 and 28, respectively. Physico-
ear gradient program for negative ion mode was started with B at 10% chemical properties of individual PhACs and EDCs were used to evaluate
and kept until 0.5 min, increased to 50% at 2 min, to 100% at 6 min, uptake, translocation and bioaccumulation behaviour patterns.
kept at 100% until 10 min, then decreased to 10% at 10.5 min and kept Physicochemical properties of the target PhACs and EDCs are shown
till 14 min. The transitions (Q1 and Q3) for each compound were listed in Table1. Octanol-water partition coefficients of the neutral com-
in Table S1. pounds (Kow,N) and organic carbon-water partition coefficients (Koc)
were compiled the US EPA's Estimation Programs Interface (EPI)
2.5. Quality assurance/quality control Suite™. Octanol-water distribution coefficient (Dow), which is an esti-
mate of the ratio of the sum of the concentrations of all chemical species
All samples were divided into several batches for the treatment and (ionized + un-ionized) in octanol and water, was determined from the
analysis. Each batch consisted of 1 procedural blank and 10–12 samples. acid dissociation constant (pKa) at pH 7. Membrane-water distribution
Analytes were identified and quantified based on the following criteria: coefficient (Dmw) values were estimated from observations of
1) two MRM transitions of each analyte were detected during the entire liposome-water partition coefficients, following the approach used by
chromatographic run; 2) retention time (RT) of the peak of analyte was Armitage et al. (2013). The compiled information was used to better as-
within 3 s of the RT obtained from analysis of authentic compounds in sess the influence of physicochemical properties on transport and
the calibration standards; 3) the difference of the observed ratio of the partitioning behaviour of these ionizable organic compounds in neutral
two monitored ions for each compound was within 20% between the lipids, phospholipids and non-lipid organic matter of aquatic
analytes and calibration standards; and 4) the signal-to-noise (S/N) macrophytes.
ratio resulted from the peak response of the two corresponding ions
was ≥ 10 for proper quantification of the analyte. Internal standard 3. Results and discussions
isotope-dilution method was used for the calculation of concentrations
of analytes, using mean relative response factors (RRF) derived from 7 3.1. PhAC and EDC concentrations in water
calibration standards with relative standard deviation (RSD) below
20%. Method detection limits (MDLs) were set equal to the instrument Concentrations of test compounds in experimental and control tank
detection limit considering the matrix effects of each sample, deter- water ranged between 19.6 and 20.3 μg/L and 19.7 to 20.4 μg/L on day 0,
mined from each analyte peak response with a S/N ratio of 5. The between 18.0 and 18.6 μg/L and 19.0 to 20.5 μg/L on day 2, and between
MDLs and absolute recoveries for target PhACs and EDCs are provided 18.1 and 18.9 μg/L and 19.1 to 20.6 μg/L on day 4. The test compounds in
in Table S3 and Table S4 respectively. Isotope labeled surrogate com- water were below MDLs during the depuration phase (day 14, 16 and
pounds and the isotope dilution quantification approach were utilized, 18). The mass balance of chemicals in water was mostly between
which inherently corrected for exact losses in every sample. Thus, all 90.1% and 103.1%, indicating that fluctuations were negligible. There-
samples were recovery corrected. Compounds with blank levels lower fore, it can be assumed that the 2-day refilling of water with PhACs
than 10% of extracted sample matrix were further calculated and statis- and EDCs was able to maintain a constant water concentration of
tically analyzed without blank correction. chemicals and no further collection of water samples was needed.

2.6. Data analysis 3.2. Uptake and elimination of PhACs and EDCs

The approach for data treatment and analysis were similar as that Uptake of PhACs and EDCs in plants was observed for both species,
described in Pi et al. (2017). In brief, plant growth rate constant (kg) with amount and rate specific in individual parts, plant species and
was determined using the measured dry biomass during exposure compounds. Observed concentrations of individual PhACs and EDCs in
phase: plant root and leaf tissues during the experimental period are shown
in Figs. 1 and 2, and Figs. S1 and S2 (see Supporting information). During
Mt ¼ M 0 ekg t ð1Þ the exposure phase, concentrations of tested chemicals in plant roots
and leaves of both species typically increased rapidly between day 0
where Mt. and M0 are the total plant dry biomass at time t and 0 and 8, continuing to increase until an apparent steady state was
respectively. reached, typically between day 8 and 14. During the depuration phase,
816 N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820

concentrations of chemicals absorbed in root and leaf decreased rapidly 3.3. Bioconcentration factors (BCFs)
between day 14 and 21.
The uptake and elimination rate constants (ku, ke) for the various The BCFs of individual PhACs and EDCs in E. horemanii and
PhACs and EDCs are summarized in Table 2 and Table S5 (see E. crassipes, varied substantially (Table 3). On a whole-plant basis,
Supporting Information). On a whole-plant basis (leaf + root), ku values PhACs exhibited BCF values ranging from 10.1 (warfarin) to 4390 L/kg
in some PhACs and EDCs, including atenolol, BPA, caffeine, (triclosan) and from 9.4 (warfarin) to 6130 L/kg (diphenhydramine)
carbamazapine, dilantin, diphenhydramine and thiabendazole, were in E. horemanii and E. crassipes respectively. BCFs of EDCs in
found much higher in E. crassipes, ku values of sulfamethoxazole and E. horemanii and E. crassipes ranged from 5.8 (BPA) to 291 L/kg (thiaben-
warfarin were comparable between the two plant species, while ku dazole) and from 2.2 (E2) to 763 L/kg (thiabendazole). BCFs of individ-
values of the remaining compounds were observed higher in ual PhACs and EDCs were different between the two plant species.
E. horemanii. ku values ranged from 0.9 L/kg·d for warfarin to Atrazine, diclofenac, E1, E2, EE2, gemfibrozil, ibuprofen, naproxen, sulfa-
843 L/kg·d for triclosan in E. horemanii, and ranged from 0.4 L/kg·d methoxazole and triclosan exhibited higher BCFs in the submerged spe-
for E1 to 1550 L/kg·d for diphenhydramine in E. crassipes. Between cies, E. horemanii. Atenolol, BPA, caffeine, carbamazapine, dilantin,
the two plant species, whole-plant ke values for E. horemanii ranged ap- diphenhydramine, linuron and thiabendazole exhibited higher BCFs in
proximately from 0.07 d− 1 (atenolol and BPA) to 0.47 d− 1 the free-floating species, E. crassipes.
(carbamazapine). ke values observed for E. crassipes were between BCFs based on leaf and root tissue concentrations were clearly differ-
0.10 d−1 (E1) and 0.42 d−1 (dilantin). ent between the two macrophyte species E. crassipes, the floating

Fig. 1. Observed concentrations of individual PhACs and EDCs in roots and leaf tissue of E. horemanii during the 28-day experiment. Data are represented as mean ± standard deviation (n = 3).
N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820 817

Fig. 2. Observed concentrations of individual PhACs and EDCs in roots and leaf tissue of E. crassipes during the 28-day experiment. Data are represented as mean ± standard deviation (n = 3).

species, generally exhibited higher root BCFs compared to E. horemanii E. horemanii, the submerged species, with exceptions of diphenhydra-
for most compounds tested. For example, the root BCF for atenolol in mine and E2. Most tested compounds in E. horemanii were found to al-
E. crassipes (2660 L/kg) was nearly 20 times higher than that of locate more in leaf compared to root (i.e., TFs N 1), whereas, most
E. horemanii (118 L/kg). BCFs in E. horemanii were found to be higher compounds in E. crassipes were observed to allocate more in root, com-
in leaf, compared to root, for most compounds tested. Conversely, pared to leaf (i.e., TFs b 1). The observed TFs further highlight that PhACs
BCFs for most compounds in E. crassipes were higher in root, compared and EDCs are more efficiently accumulated in leaf tissue of E. horemanii
to leaf. These results can be attributed to the submerged nature of due to the submerged nature of this species. For E. crassipes, a free-
E. horemanii, which allows direct chemical exposure and partitioning floating species with foliage not directly exposed to the water, efficient
between water and leaf tissue. internal transport from roots into foliage is required for translocation to
occur.
3.4. Translocation factors (TFs) At the end of depuration phase (day 28), the majority of the PhACs
and EDCs in E. crassipes were allocated more in roots, compared to leaf tis-
The observed differences between the two types of macrophytes sue, which was the same as that in exposure phase (day 14) (i.e., TFs b 1).
were further shown by the determination of TFs (Table S6, Supporting However, for E. horemanii, the allocation of some compounds in leaf and
information). The difference of TFs was observed among the com- root was observed to be different between exposure and depuration pe-
pounds tested and between species. At the end of exposure phase riod. For example, the TF of caffeine in exposure period was N1 (4.98)
(day 14), TFs for the tested compounds were generally higher in and nearly 5 times higher than that in depuration period (0.91).
818 N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820

Table 2
Whole-plant uptake rate constants (ku, L/kg·d) and elimination rate constants (ke, d−1) of
individual PhACs and EDCs in E. horemanii (submerged macrophyte) and E. crassipes (free-
floating macrophyte).

Compound E. horemanii E. crassipes


ku (L/kg·d) ke (d−1) ku (L/kg·d) ke (d−1)

Atenolol 79.1 ± 6.6 0.07 ± 0.01 503 ± 25.5 0.12 ± 0.01


Atrazine 90.0 ± 8.3 0.25 ± 0.02 38.4 ± 2.9 0.25 ± 0.02
BPA 1.1 ± 0.1 0.07 ± 0.01 5.6 ± 0.5 0.12 ± 0.01
Caffeine 17.5 ± 1.3 0.27 ± 0.03 34.7 ± 2.9 0.31 ± 0.02
Carbamazapine 124 ± 10.7 0.47 ± 0.04 176 ± 12.9 0.39 ± 0.04
Diclofenac 20.0 ± 1.7 0.23 ± 0.02 5.7 ± 0.5 0.19 ± 0.02
Dilantin 55.1 ± 4.4 0.32 ± 0.02 102 ± 8.2 0.42 ± 0.03
Diphenhydramine 410 ± 29.9 0.36 ± 0.03 1552 ± 121 0.13 ± 0.01
E1 2.0 ± 0.2 0.11 ± 0.01 0.4 ± 0.0 0.10 ± 0.02
E2 2.2 ± 0.2 0.10 ± 0.01 0.7 ± 0.1 0.15 ± 0.01
EE2 38.2 ± 2.7 0.23 ± 0.02 17.7 ± 1.4 0.15 ± 0.01
Gemfibrozil 62.2 ± 4.2 0.17 ± 0.02 14.5 ± 1.2 0.18 ± 0.02
Ibuprofen 20.3 ± 1.6 0.20 ± 0.02 6.1 ± 0.5 0.19 ± 0.02
Linuron 76.2 ± 5.9 0.35 ± 0.03 129 ± 9.4 0.32 ± 0.02
Naproxen 77.9 ± 6.1 0.13 ± 0.01 62.0 ± 5.1 0.21 ± 0.02
Sulfamethoxazole 3.3 ± 0.3 0.13 ± 0.01 3.0 ± 0.3 0.17 ± 0.02
Thiabendazole 104 ± 7.4 0.32 ± 0.03 259 ± 17.7 0.25 ± 0.02
Triclosan 843 ± 51.6 0.12 ± 0.01 385 ± 28.9 0.24 ± 0.02
Warfarin 0.9 ± 0.1 0.08 ± 0.01 1.1 ± 0.1 0.14 ± 0.01

3.5. Influence of physicochemical properties

Fig. 3 illustrates plots of log BCFss values of the individual PhACs and
EDCs in whole plant of aquatic macrophyte versus chemical log Dow and
log Dmw. Similarly, Fig. 4 shows plots of the corresponding log TF values
of the studied compounds at the end of exposure phase versus chemical
log Dow and log Dmw. Recent studies of uptake and depuration of Fig. 3. Plots showing the relationship between the whole-plant bioconcentration factors
perfluoroalkyl substances (PFASs) in aquatic macrophytes demonstrat- (log BCFss, L/kg) of test chemicals versus (A) octanol-water distribution coefficient (log
ed strong relationships between bioconcentration and translocation po- Dow) and (B) membrane-water distribution coefficient (log Dmw).
tential and chemical Dow, Dmw and Koc (Pi et al., 2017). In particular, the
more hydrophobic long-chain PFASs, with higher Dow, Dmw and Koc, ex- Log BCFs/Log TFs and the organic carbon-water partition coefficient
hibited relatively lower TFs, but higher whole-plant BCFs. The data indi- (Koc), which is a representation of chemical partitioning in the organic
cate a higher degree of partitioning of those more hydrophobic carbon based constituents (Fig. S3). The lack of a discernible relation-
compounds into neutral and phospholipids and non-lipid organic mat- ship between BCFs and TFs and hydrophobicity may be due, in part, to
ter of aquatic macrophytes for more hydrophobic compounds. In con- variability in degradation of the different compounds. It is also impor-
trast, the results of the present study of PhACs and EDCs do not tant to note that the studied PhACs and EDCs comprise a range of differ-
indicate any correlation between log BCFs/log TFs and log Dow/log ent chemical classes, exhibiting different ionizable function groups,
Dmw values (Fig. 3, Fig. 4). Similarly, there was no clear trend between which may also influence uptake and partitioning behaviour.

Table 3
Observed steady-state bioconcentration factors (BCFss, L/kg) of individual PhACs and EDCs in leaf, roots and whole-plant of E. horemanii (submerged macrophyte) and E. crassipes (free-
floating macrophyte).

Compound E. horemanii E. crassipes


Leaf Root Whole plant Leaf Root Whole plant

Atenolol 463 ± 31.9 118 ± 8.9 402 ± 29.9 1560 ± 112.7 2660 ± 177 2050 ± 133
Atrazine 319 ± 23.3 30.3 ± 2.5 259 ± 18.3 138 ± 12.5 56.9 ± 4.3 106 ± 8.4
BPA 7.9 ± 0.7 16.4 ± 1.1 5.8 ± 0.5 25.6 ± 2.2 71.4 ± 5.9 19.8 ± 1.2
Caffeine 62.5 ± 5.6 8.5 ± 0.7 52.0 ± 5.1 144 ± 12.0 29.7 ± 2.0 98.4 ± 6.3
Carbamazapine 370 ± 31.1 98.3 ± 7.4 330 ± 21.4 458 ± 29.8 313 ± 21.6 402 ± 28.4
Diclofenac 75.9 ± 6.4 96.1 ± 7.0 78.9 ± 7.0 5.2 ± 0.4 42.3 ± 24.3 19.4 ± 1.4
Dilantin 173 ± 14.6 85.8 ± 7.8 161 ± 13.3 203 ± 15.2 266 ± 17.7 233 ± 15.2
Diphenhydramine 1100 ± 97.4 274 ± 19.6 939 ± 57.3 9350 ± 365 1290 ± 102 6130 ± 313
E1 15.4 ± 1.2 5.9 ± 0.6 13.4 ± 1.2 0.1 ± 0.01 9.7 ± 0.9 4.1 ± 0.4
E2 12.3 ± 1.0 1.2 ± 0.1 10.1 ± 0.9 3.3 ± 0.3 0.3 ± 0.04 2.2 ± 0.2
EE2 153 ± 13.8 47.5 ± 3.2 133 ± 11.1 57.0 ± 4.1 140 ± 10.9 57.9 ± 4.4
Gemfibrozil 351 ± 31.1 91.8 ± 7.9 314 ± 23.5 12.9 ± 1.2 126 ± 10.3 46.7 ± 4.7
Ibuprofen 120 ± 10.6 24.3 ± 1.9 105.9 ± 9.1 10.5 ± 0.9 38.5 ± 3.2 21.0 ± 1.7
Linuron 244 ± 19.6 40.9 ± 3.0 214 ± 15.4 222 ± 18.2 432 ± 32.4 307 ± 21.7
Naproxen 532 ± 33.4 95.6 ± 8.1 468 ± 33.6 114 ± 9.2 345 ± 26.3 207 ± 16.3
Sulfamethoxazole 23.7 ± 1.7 21.4 ± 1.9 23.4 ± 1.9 6.7 ± 0.6 15.5 ± 1.2 10.3 ± 0.8
Thiabendazole 287 ± 22.2 315 ± 24.3 291 ± 21.5 284 ± 21.1 1480 ± 117 763 ± 56.5
Triclosan 4570 ± 321 643 ± 43.8 4390 ± 296 809 ± 33.5 1580 ± 125 1050 ± 89.2
Warfarin 7.7 ± 0.6 21.8 ± 1.7 10.1 ± 0.9 2.4 ± 0.2 24.7 ± 2.0 9.4 ± 0.8
N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820 819

and translocation of these compounds within plants can be simply driv-


en by diffusion (Dietz and Schnoor, 2001; Dordio et al., 2011; Dordio
and Carvalho, 2013; Li et al., 2014). Previous studies have pointed out
that uptake and translocation of organic contaminants in aquatic sys-
tems caused by the diffusion process can be affected by physicochemical
properties of compounds such as water solubility, hydrophobicity (log
Kow), pKa, as well as pH and exposure concentration in the water
(Dordio and Carvalho, 2013; Garcia-Rodriguez et al., 2014; Hurtado
et al., 2016; Prosser et al., 2014; Tsao, 2003). It has been suggested
that moderately hydrophobic chemicals (log Kow range: 0.5 to 3.5) are
sufficiently lipophilic to move through the lipid bilayer of plant cell
membranes and water soluble enough to travel into the cell fluids of
plants, hence are more likely to be taken up and translocated to leaves
(Dordio and Carvalho, 2013; Li et al., 2014; Pilon-Smits, 2005). For ex-
ample, uptake and translocation of carbamazepine (log Dow = 2.4) in
plants has been attributed to efficient transport across root membranes
and cell fluids (Garcia-Rodriguez et al., 2014; Holling et al., 2012; Paz
et al., 2016; Shenker et al., 2011; Wu et al., 2010a; Zhang et al.,
2013a). Dordio et al. reported that carbamazepine can be readily
taken up by the roots of Typha spp. and then transported from roots to
stems and leaves (Dordio et al., 2011). Similarly, previous studies have
shown caffeine (Log Dow = −1.0) and sulfamethoxazole (Log Dow =
−0.1) could be easily taken up and translocated in cabbage and soybean
plants (Holling et al., 2012; Wu et al., 2010a).
Overall, the results from the present study are consistent with previ-
ous observations of PhAC and EDC bioaccumulation in plants. The re-
sults suggest that bioaccumulation and translocation of PhACs and
EDCs in aquatic plants is not easily predicted by physicochemical prop-
erties. Less hydrophobic compounds such as carbamazapine and di-
phenhydramine are more readily taken up and translocated in leaf
tissue. Triclosan, a weak acid (pKa = 7.8) with a log Dow of 4.7, exhibits
minimal translocation, but a relatively high degree of bioaccumulation
potential on a whole-plant basis. Further empirical and/or modeling
Fig. 4. Plots showing the relationship between the translocation factors (log TFs) of test studies are required to better understand the role of physicochemical
chemicals at the end of exposure phase versus (A) octanol-water distribution coefficient properties in the uptake and distribution of these ionizable organic con-
(log Dow) and (B) membrane-water distribution coefficient (log Dmw). The dashed line taminants in aquatic macrophytes.
represents a TF equal to 1.

3.6. Implications for phytoremediation and risk assessment

Alternatively, the hydrophobicity range investigated (log Dow − 2.3 to The results from the present study suggest that PhACs and EDCs
4.7) may be too narrow to observe clear differences. can be readily taken up and accumulated in aquatic macrophytes.
In the present study, triclosan (log Dow = 4.7) exhibited relatively The bioaccumulation potential and the distribution between roots
high bioaccumulation potential, with whole-plant BCFss values equal and leaf tissue vary widely among for these compounds. The
to 1050 and 4390 L/kg in E. crassipes and E. horemanii, respectively bioaccumulation behaviour of PhACs and EDCs in submerged and
(Table 2, and Table S5, Supporting information). However, less hydro- free-floating macrophytes is markedly different, as indicated by the
phobic compounds such as diphenhydramine (log Dow = 1.3) also ex- observed ku and BCF values in leaf and root tissues. For submerged
hibited very high bioconcentration potential (BCFss N 6000 L/kg). macrophytes such as E. horemanii, PhACs and EDCs are more prefer-
Similar to previous studies, carbamazapine (log Dow = 2.4) was found entially allocated in leaf tissue compared to root, due to the fact
to exhibit a high degree of bioaccumulation in both E. horemanii (BCFss leaves are in direct contact with the waterborne compounds. Con-
of 330 L/kg) and E. crassipes (BCFss of 402 L/kg). Similar results were ob- versely, for free-floating species such E. crassipes, PhACs and EDCs
served for atenolol (log Dow = −2.3), which exhibited a whole-plant tend to be more are allocated more in roots compared to leaf tissue,
BCFss of 2050 L/kg in E. crassipes. Surprisingly, more hydrophobic com- with the exception of more hydrophilic compounds that can be ef-
pounds such as BPA, E1, E2 and EE2 exhibited very low bioaccumulation fectively translocated.
potential in these aquatic macrophytes. These findings may be useful for future design and implementation
Regarding translocation, the results show that carbamazapine (log of phytoremediation systems, as well as environmental risk assessment
Dow = 2.4) is effectively transported from root to leaf, with TFs values initiatives. Phytoremediation involves the use of plant species to re-
in E. horemanii and E. crassipes of 3.76 and 1.46, respectively, which is move environmental contaminants via uptake, translocation, accumula-
consistent with previous studies (Dordio et al., 2011). Caffeine (log tion and/or degradation. This study provides new information regarding
Dow = − 1.0) was also shown to be readily taken up translocated bioaccumulation behaviour of PhACs and EDCs in aquatic macrophytes,
from root to leaf for both species, with TFs values much higher than 1, which can be used for removing waterborne contaminants in wastewa-
consistent with previous studies (Dettenmaier et al., 2009; ter impacted lagoons or surface waters, as well as in constructed wet-
Garcia-Rodriguez et al., 2014; Matamoros et al., 2012; Zhang et al., land systems, through direct sorption and/or translocation. In
2013a; Zhang et al., 2013b). addition, the information regarding uptake and elimination kinetics of
In the cell membranes of plant roots, non-specific transporters can PhACs and EDCs in aquatic macrophytes may aid future bioaccumula-
transport PhACs and EDCs into the plants tissues, however, the uptake tion modeling and risk assessment initiatives involving PhACs and EDCs.
820 N. Pi et al. / Science of the Total Environment 601–602 (2017) 812–820

Acknowledgements Kannan, K., Reiner, J.L., Yun, S.H., Perrotta, E.E., Tao, L., Johnson-Restrepo, B., et al., 2005.
Polycyclic musk compounds in higher trophic level aquatic organisms and humans
from the United States. Chemosphere 61, 693–700.
The work was funded by a Singapore Ministry of Education (MOE) Li, Y., Zhu, G., Ng, W.J., Tan, S.K., 2014. A review on removing pharmaceutical contami-
Academic Research Fund (AcRF) Tier 1 grant to B.C. Kelly. Thanks are al- nants from wastewater by constructed wetlands: design, performance and mecha-
nism. Sci. Total Environ. 468-469, 908–932.
so given to Singapore-Delft Water Alliance and NUS Environmental Re- Matamoros, V., Caselles-Osorio, A., Garcia, J., Bayona, J.M., 2008. Behaviour of pharmaceu-
search Institute (NERI) for facilities support. tical products and biodegradation intermediates in horizontal subsurface flow con-
structed wetland. A microcosm experiment. Sci. Total Environ. 394, 171–176.
Matamoros, V., Nguyen, L.X., Arias, C.A., Salvado, V., Brix, H., 2012. Evaluation of aquatic
Appendix A. Supplementary data
plants for removing polar microcontaminants: a microcosm experiment.
Chemosphere 88, 1257–1264.
Supplementary data to this article can be found online at http://dx. Metcalfe, C.D., Chu, S.G., Judt, C., Li, H.X., Oakes, K.D., Servos, M.R., et al., 2010. Antidepres-
sants and their metabolites in municipal wastewater, and downstream exposure in
doi.org/10.1016/j.scitotenv.2017.05.137.
an urban watershed. Environ. Toxicol. Chem. 29, 79–89.
Monteiro, S.C., Boxall, A.B.A., 2009. Factors affecting the degradation of pharmaceuticals in
References agricultural soils. Environ. Toxicol. Chem. 28, 2546–2554.
Morley, N.J., 2009. Environmental risk and toxicology of human and veterinary waste
Armitage, J.M., Arnot, J.A., Wania, F., Mackay, D., 2013. Development and evaluation of a pharmaceutical exposure to wild aquatic host-parasite relationships. Environ.
mechanistic bioconcentration model for ionogenic organic chemicals in fish. Environ. Toxicol. Pharmacol. 27, 161–175.
Toxicol. Chem. 32, 115–128. Nakata, H., 2005. Occurrence of synthetic musk fragrances in marine mammals and
Bayen, S., Zhang, H., Desai, M.M., Ooi, S.K., Kelly, B.C., 2013. Occurrence and distribution of sharks from Japanese coastal waters. Environ. Sci. Technol. 39, 3430–3434.
pharmaceutically active and endocrine disrupting compounds in Singapore's marine Paz, A., Tadmor, G., Malchi, T., Blotevogel, J., Borch, T., Polubesova, T., et al., 2016. Fate of
environment: influence of hydrodynamics and physical-chemical properties. Envi- carbamazepine, its metabolites, and lamotrigine in soils irrigated with reclaimed
ron. Pollut. 182, 1–8. wastewater: sorption, leaching and plant uptake. Chemosphere 160, 22–29.
Breton, R., Boxall, A., 2003. Pharmaceuticals and personal care products in the environ- Petrie, B., McAdam, E.J., Scrimshaw, M.D., Lester, J.N., Cartmell, E., 2013. Fate of drugs dur-
ment: regulatory drivers and research needs. QSAR Comb. Sci. 22, 399–409. ing wastewater treatment. Trac-Trend Anal. Chem. 49, 145–159.
Carlsson, C., Johansson, A.K., Alvan, G., Bergman, K., Kuhler, T., 2006. Are pharmaceuticals Pi, N., Tam, N.F., Wong, M.H., 2011. Formation of iron plaque on mangrove roots receiving
potent environmental pollutants? Part I: environmental risk assessments of selected wastewater and its role in immobilization of wastewater-borne pollutants. Mar.
active pharmaceutical ingredients. Sci. Total Environ. 364, 67–87. Pollut. Bull. 63, 402–411.
Corcoran, J., Winter, M.J., Tyler, C.R., 2010. Pharmaceuticals in the aquatic environment: a Pi, N., Ng, J.Z., Kelly, B.C., 2017. Uptake and elimination kinetics of perfluoroalkyl sub-
critical review of the evidence for health effects in fish. Crit. Rev. Toxicol. 40, 287–304. stances in submerged and free-floating aquatic macrophytes: results of mesocosm
Crane, M., Watts, C., Boucard, T., 2006. Chronic aquatic environmental risks from exposure experiments with Echinodorus horemanii and Eichhornia crassipes. Water Res. 117,
to human pharmaceuticals. Sci. Total Environ. 367, 23–41. 167–174.
Daughton, C.G., Ternes, T.A., 1999. Pharmaceuticals and personal care products in the en- Pilon-Smits, E., 2005. Phytoremediation. Annu. Rev. Plant Biol. 56, 15–39.
vironment: agents of subtle change? Environ. Health Perspect. 107 (Suppl. 6), Prosser, R.S., Trapp, S., Sibley, P.K., 2014. Modeling uptake of selected pharmaceuticals and
907–938. personal care products into food crops from biosolids-amended soil. Environ. Sci.
Dettenmaier, E.M., Doucette, W.J., Bugbee, B., 2009. Chemical hydrophobicity and uptake Technol. 48, 11397–11404.
by plant roots. Environ. Sci. Technol. 43, 324–329. Shenker, M., Harush, D., Ben-Ari, J., Chefetz, B., 2011. Uptake of carbamazepine by cucum-
Dietz, A.C., Schnoor, J.L., 2001. Advances in phytoremediation. Environ. Health Perspect. ber plants–a case study related to irrigation with reclaimed wastewater.
109 (Suppl. 1), 163–168. Chemosphere 82, 905–910.
Dordio, A.V., Carvalho, A.J., 2013. Organic xenobiotics removal in constructed wetlands, Togola, A., Budzinski, H., 2008. Multi-residue analysis of pharmaceutical compounds in
with emphasis on the importance of the support matrix. J. Hazard. Mater. 252–253, aqueous samples. J. Chromatogr. A 1177, 150–158.
272–292. Tsao, D.T., 2003. Advances in Biochemical Engineering Biotechnology. Springer, Berlin.
Dordio, A., Ferro, R., Teixeira, D., Palace, A.J., Pinto, A.P., Dias, C.M.B., 2011. Study on the use USEPA, 2007. Method 1694: Pharmaceuticals and Personal Care Products in Water, Soil,
of Typha spp. for the phytotreatment of water contaminated with ibuprofen. J Envi- Sediment, and Biosolids by HPLC/MS/MS. 2015.
ron Anal Chem]–>Int. J. Environ. Anal. Chem. 91, 654–667. USEPA, 2014. Contaminants of Emerging Concern. 2015.
EC, 2011. Proposal for a Directive of the European Parliament and of the Council Vanderford, B.J., Snyder, S.A., 2006. Analysis of pharmaceuticals in water by isotope dilu-
amending Directives 2000/60/EC and 2008/105/EC as Regards Priority Substances tion liquid chromatography/tandem mass spectrometry. Environ. Sci. Technol. 40,
in the Field of Water Policy. 2015. 7312–7320.
la Farre, M., Perez, S., Kantiani, L., Barcelo, D., 2008. Fate and toxicity of emerging pollut- WHO, 2012. Pharmaceuticals in Drinking-water. 2015.
ants, their metabolites and transformation products in the aquatic environment. Wu, C., Spongberg, A.L., Witter, J.D., Fang, M., Czajkowski, K.P., 2010a. Uptake of pharma-
Trac-Trend Anal. Chem. 27, 991–1007. ceutical and personal care products by soybean plants from soils applied with bio-
Felizeter, S., McLachlan, M.S., de Voogt, P., 2012. Uptake of perfluorinated alkyl acids by solids and irrigated with contaminated water. Environ. Sci. Technol. 44, 6157–6161.
hydroponically grown lettuce (Lactuca sativa). Environ. Sci. Technol. 46, Wu, J., Qian, X., Yang, Z., Zhang, L., 2010b. Study on the matrix effect in the determination
11735–11743. of selected pharmaceutical residues in seawater by solid-phase extraction and ultra-
Garcia-Rodriguez, A., Matamoros, V., Fontas, C., Salvado, V., 2014. The ability of biological- high-performance liquid chromatography-electrospray ionization low-energy
ly based wastewater treatment systems to remove emerging organic contaminants–a collision-induced dissociation tandem mass spectrometry. J. Chromatogr. A 1217,
review. Environ. Sci. Pollut. Res. Int. 21, 11708–11728. 1471–1475.
Halling-Sorensen, B., Nors Nielsen, S., Lanzky, P.F., Ingerslev, F., Holten Lutzhoft, H.C., Zarate Jr., F.M., Schulwitz, S.E., Stevens, K.J., Venables, B.J., 2012. Bioconcentration of triclo-
Jorgensen, S.E., 1998. Occurrence, fate and effects of pharmaceutical substances in san, methyl-triclosan, and triclocarban in the plants and sediments of a constructed
the environment–a review. Chemosphere 36, 357–393. wetland. Chemosphere 88, 323–329.
Hijosa-Valsero, M., Reyes-Contreras, C., Dominguez, C., Becares, E., Bayona, J.M., 2016. Be- Zhang, D.Q., Gersberg, R.M., Hua, T., Zhu, J.F., Tuan, N.A., Tan, S.K., 2012a. Pharmaceutical
haviour of pharmaceuticals and personal care products in constructed wetland com- removal in tropical subsurface flow constructed wetlands at varying hydraulic load-
partments: influent, effluent, pore water, substrate and plant roots. Chemosphere ing rates. Chemosphere 87, 273–277.
145, 508–517. Zhang, D.Q., Gersberg, R.M., Zhu, J., Hua, T., Jinadasa, K.B., Tan, S.K., 2012b. Batch versus
Holling, C.S., Bailey, J.L., Vanden Heuvel, B., Kinney, C.A., 2012. Uptake of human pharma- continuous feeding strategies for pharmaceutical removal by subsurface flow con-
ceuticals and personal care products by cabbage (Brassica campestris) from fortified structed wetland. Environ. Pollut. 167, 124–131.
and biosolids-amended soils. J. Environ. Monit. 14, 3029–3036. Zhang, D.Q., Hua, T., Gersberg, R.M., Zhu, J., Ng, W.J., Tan, S.K., 2013a. Carbamazepine and
Huerta, B., Rodriguez-Mozaz, S., Nannou, C., Nakis, L., Ruhi, A., Acuna, V., et al., 2016. Deter- naproxen: fate in wetland mesocosms planted with Scirpus validus. Chemosphere 91,
mination of a broad spectrum of pharmaceuticals and endocrine disruptors in biofilm 14–21.
from a waste water treatment plant-impacted river. Sci. Total Environ. 540, 241–249. Zhang, D.Q., Hua, T., Gersberg, R.M., Zhu, J.F., Ng, W.J., Tan, S.K., 2013b. Fate of caffeine in
Hurtado, C., Dominguez, C., Perez-Babace, L., Canameras, N., Comas, J., Bayona, J.M., 2016. mesocosms wetland planted with Scirpus validus. Chemosphere 90, 1568–1572.
Estimate of uptake and translocation of emerging organic contaminants from irriga- Zhang, D., Gersberg, R.M., Ng, W.J., Tan, S.K., 2014. Removal of pharmaceuticals and per-
tion water concentration in lettuce grown under controlled conditions. J. Hazard. sonal care products in aquatic plant-based systems: a review. Environ. Pollut. 184,
Mater. 305, 139–148. 620–639.
Jayasiri, H.B., Purushothaman, C.S., Vennila, A., 2013. Pharmaceutically active compounds
(PhACs): a threat for aquatic environment? J. Mar. Sci. Res. Dev. 4, e122.

Potrebbero piacerti anche