You are on page 1of 10

Environmental Pollution 265 (2020) 114926

Contents lists available at ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Adsorption of chlorophenols on polyethylene terephthalate


microplastics from aqueous environments: Kinetics, mechanisms and
influencing factors*
Zheming Liu , Qingdong Qin , Zhixian Hu , Lu Yan , Un-Io Ieong , Yan Xu *
Department of Municipal Engineering, School of Civil Engineering, Southeast University, Nanjing, Jiangsu, 210096, China

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

Article history: Microplastics have received growing attention as carriers of organic pollutants in the water environment.
Received 14 February 2020 To better understand the contribution of hydrophobic interaction, hydrogen-bonding interaction, p-p
Received in revised form interaction and electrostatic interaction on the adsorption of hydrophilic compounds on microplastics
31 May 2020
and their adsorption behavior in natural waters, polyethylene terephthalate (PET, <150 mm) was used as
Accepted 31 May 2020
an adsorbent and 4-chlorophenol (MCP), 2,4-dichlorophenol (DCP) and 2,4,6-trichlorophenol (TCP) were
Available online 6 June 2020
used as adsorbates. The results of batch adsorption experiments showed that chlorophenols (CPs)
reached adsorption sites of PET through film diffusion and intra-particle diffusion. pH greatly affected the
Keywords:
Microplastics
adsorption capacity. Hydrophobic interaction was the main adsorption mechanism of undissociated CPs
Adsorption on PET. Hydrogen-bonding interaction was also an adsorption mechanism between undissociated CPs
Chlorophenols and PET, and the contribution of hydrogen-bonding interaction to adsorption decreased with the in-
Diffusion crease of chlorine content. Meanwhile, the increase of chlorine content was favorable to the hydrophobic
NOM interaction between undissociated CPs and PET. However, higher chlorine content CPs with lower pKa
values tended to dissociate at neutral pH condition and resulted in stronger electrostatic repulsion with
PET. The increase of solution ionic strength and fulvic acid content negatively affected the adsorption of
DCP and TCP on PET, but did not show significant impacts on MCP adsorption. Similarly, the adsorption
capacity obtained using Taihu lake water and Bohai seawater as matrices was much lower than that using
laboratory water for both DCP and TCP, while the adsorption coefficient (Kd) of MCP remained at
approximately 10.6 L/kg to 11.4 L/kg in the three different solution matrices. The Kd values exhibited
using natural water matrices consistently followed the order of DCP > MCP > TCP. This study provides
insights into the fate of CPs in the presence of microplastics and suggests that the potential risks posed by
CPs and microplastics to aqueous ecosystems merit further investigation.
© 2020 Elsevier Ltd. All rights reserved.

1. Introduction are believed to be the direct introduction with runoff and weath-
ering breakdown of meso- and macroplastics debris (Andrady,
The threat of plastic debris to the marine environment has been 2011), resulting in a tremendous number of microplastic particles
recognized since the end of the 20th century (Derraik, 2002). In the in aquatic environments (Rezania et al., 2018), terrestrial soil sys-
past decade, microplastics, plastic particles smaller than 5 mm, tems (Wang et al., 2020) and atmosphere (Chen et al., 2020). The
have attracted increasing attention due to their evidenced potential uptake of microplastics by marine organisms has been observed
adverse effects on aquatic ecosystems (Arthur et al., 2009; (Rezania et al., 2018) and is proven to have negative physical im-
Thompson et al., 2004). The two primary sources of microplastics pacts on their bodies (Wright et al., 2013). Moreover, microplastics
may enter the human body through commonly consumed items
and the environment (Cox et al., 2019) and pose risks to human
*
This paper has been recommended for acceptance by Eddy Y. Zeng.
health (Lehner et al., 2019; Wright and Kelly, 2017). Given their
* Corresponding author. hydrophobic properties, microplastics can both carry organic pol-
E-mail addresses: seulzm@163.com (Z. Liu), qinqingdong@seu.edu.cn (Q. Qin), lutants (Mato et al., 2001; Van et al., 2012) and release them into
huzhixian97@163.com (Z. Hu), luluyan_seu@163.com (L. Yan), 1031680789@qq. the environment (Teuten et al., 2009). Also, the interaction of
com (U.-I. Ieong), xuxucalmm@seu.edu.cn (Y. Xu).

https://doi.org/10.1016/j.envpol.2020.114926
0269-7491/© 2020 Elsevier Ltd. All rights reserved.
2 Z. Liu et al. / Environmental Pollution 265 (2020) 114926

microplastics and organic contaminants can aggravate the toxicity approximately 11.6 billion microplastics and 3.1 billion nanoplastics
of some chemicals and/or enhance bioaccumulation (Deng et al., at brewing temperature (95  C) (Hernandez et al., 2019).
2018; Ma et al., 2016; Qiao et al., 2019). Hydrophobic interaction, hydrogen-bonding interaction and p-
Microplastics are capable of adsorbing a variety of anthropo- p interaction are the adsorption mechanisms of undissociated hy-
genic organic pollutants from the environment (Wang et al., 2018). drophilic compounds on microplastics (Tourinho et al., 2019; Yu
The adsorption behaviors of the target organic pollutants onto et al., 2019). However, the contribution of these mechanisms to
microplastics were determined by the properties of microplastics, adsorption remains unknown. To better understand these adsorp-
organic pollutants as well as water matrices (Tourinho et al., 2019; tion mechanisms and other possible adsorption mechanisms, PET
Yu et al., 2019). Generally, smaller microplastics, having a larger and three typical CPs (4-chlorophenol, 2,4-dichlorophenol and
specific surface area, tend to exhibit a higher adsorption capacity 2,4,6-trichlorophenol) were selected as adsorbent and adsorbate,
(Ma et al., 2019; Zhan et al., 2016). However, when the size reduces respectively. The diffusion process of CPs to PET was studied by
to the nanoscale, the decrease of adsorption capacity caused by the adsorption kinetics. Adsorption isotherms, the effects of solution
aggregation of nanoparticles cannot be ignored (Wang et al., 2019). pH and chlorine content of CPs were investigated to explore the
For the same polymer type, the decrease of crystallinity usually adsorption mechanisms. In addition, the effects of ionic strength
leads to an increase in the adsorption capacity of microplastics to and NOM on adsorption were also investigated. Furthermore, to
hydrophobic organic pollutants (Guo et al., 2012). But for distinct better understand the adsorption behaviors in complex aquatic
polymer types, the property of monomer is the main influencing environments, the adsorption of CPs by PET under natural water
factor. For instance, the benzene rings of polystyrene microplastics matrices were also examined and compared.
can produce p-p interaction with an aromatic organic compound to
favor adsorption (Hüffer and Hofmann, 2016; Liu et al., 2019b). 2. Materials and methods
Hydrophobic organic pollutants with higher octanol-water parti-
tion coefficients appear to be more readily adsorbed to micro- 2.1. Materials and characterization
plastics (Velzeboer et al., 2014; Wang et al., 2018). However, the
speciation of dissociable hydrophilic organic contaminants is 4-Chlorophenol (MCP) (purity > 99%), 2,4,6-trichlorophenol
commonly pH-dependent, which ultimately affects the interaction (TCP) (purity > 99%) and fulvic acid (purity > 90%) were pur-
between microplastics and pollutants (Ma et al., 2019; Zhang et al., chased from Aladdin Bio-Chem Technology Ltd (Shanghai, China).
2018). Under certain conditions, microplastics with oxygen- 2,4-dichlorophenol (DCP) (purity > 99.5%) was purchased from
containing functional groups can readily form hydrogen bonds Macklin Biochemical Ltd (Shanghai, China). The polyethylene
with some hydrophilic organic pollutants (Liu et al., 2019b; Zhang terephthalate (PET) powder (guaranteed reagent, < 150 mm) was
et al., 2018). In addition, ionic strength (Liu et al., 2019a; Ma provided by DuPont (USA). Acetonitrile was purchased from TEDIA
et al., 2019), the content of natural organic matter (NOM) (Xu (HPLC grade, USA). Other chemicals were obtained from Sinopharm
et al., 2018a; Zhang et al., 2018) as well as temperature (Liu et al., Group (Shanghai, China). Ultrapure water (>18.2 MU cm, Merck
2018) could affect the adsorption in varying degrees. Millipore, Germany) was used in the experiments unless otherwise
Chlorophenols (CPs) are a class of chlorinated aromatic com- specified.
pounds commonly used for industrial and agricultural production The microscopic surface structure of PET was characterized by
(Olaniran and Igbinosa, 2011) and killing schistosome intermediate scanning electron microscope (SEM, EVO 18, Carl Zeiss AG, Ger-
host snails (Zheng et al., 2012). CPs were found to exhibit chronic many). The surface area and pore size distribution were analyzed by
toxic effects on biota (Ge et al., 2017) and have the potential to N2 adsorption with a specific surface area and pore size analyzer
cause histopathological changes, mutagenic and carcinogenic ef- (ASAP2020M, Micromeritics Instrument Corp., USA) using the
fects (Igbinosa et al., 2013). Among the CPs, 2-chlorophenol, 2,4- Brunauer-Emmett-Teller (BET) method and the Barrett-Joyner-
dichlorophenol, 2,4,6-trichlorophenol, and pentachlorophenol are Halenda (BJH) technique, respectively. The surface functional
listed in the Priority Pollutant List of the US Environmental Pro- groups were determined by attenuated total reflection Fourier-
tection Agency (ATSDR, 2007). Apart from the direct release, other transform infrared spectroscopy (FTIR, Nicolet iS50, Thermo
sources of CPs include chlorination of humic acid, biodegradation of Fisher Scientific, USA). The zeta potentials with the change of pH
some pesticides and herbicides, disinfection of water by chlori- and the isoelectric point (IEP) of PET were measured using zeta
nating, and so on (Czaplicka, 2004). CPs are ubiquitous in the potential analyzer (Zetasizer Nano ZS, Malvern Panalytical, UK).
aquatic environment (Jensen, 1996). The concentrations of CPs in
surface water generally range from 0.005 to 20 mg/L (Gao et al., 2.2. Analytical method
2008; Ge et al., 2017). Notably, in some treated industrial waste-
water, the concentration can even reach 1.5e3.0 mg/L (Stepanova The CPs were analyzed by ultra-performance liquid chroma-
et al., 2000). In natural conditions, the degradation of these chlo- tography (UPLC, ACQUITY UPLC H-Class PLUS, Waters Corp., USA)
rinated compounds cannot proceed readily, and the degradation equipped with a UV detector and reversed-phase column (BEH C18
mechanisms are very complex (Ge et al., 2017; Lin et al., 2019). 100 mm  2.1 mm, 1.7 mm, Waters Corp., USA). The mobile phase
Moreover, under some circumstances, CPs may even become the consisted of ultrapure water containing 0.5% acetic acid and
precursor of polychlorinated biphenyls and dioxins (Czaplicka, acetonitrile with a ratio of 45:55 (v:v) for MCP and DCP; and 40:60
2004). (v:v) for TCP. The flow rate was 0.3 mL/min and the column tem-
Polyethylene terephthalate (PET), having excellent mechanical perature was 35  C. The UV detector wavelength was set at 285 nm
and thermal properties, is widely used in making textile fibers and for MCP and DCP, and 292 nm for TCP.
packaging with an annual worldwide production of over 50 million
tons (Bornscheuer, 2016; Taniguchi et al., 2019). PET is difficult to 2.3. Adsorption experiments
biodegrade, though its monomers are joined together via ester
bonds (Taniguchi et al., 2019). A previous study reported that more CPs were dissolved in ultrapure water for the preparation of
than 1900 fibers were released per garment wash and 78% of fibers 200 mg/L stock solution and kept in the dark at 4  C. All the ex-
found in marine sediments and sewage were PET fibers (Browne periments were conducted in 30 ml amber borosilicate vials with
et al., 2011). Even a single PET plastic teabag could release PTFE gaskets containing 0.5 g of PET and 20 mL of specific solutions
Z. Liu et al. / Environmental Pollution 265 (2020) 114926 3

unless otherwise specified. To inhibit microbial activity, NaN3 was organic carbon (TOC) determined by the TOC analyzer (Torch, Tel-
added to all solutions at a final concentration of 25 mg/L. The initial edyne Tekmar, USA). To examine the effect of NOM on adsorption,
pH was adjusted by adding 0.1 M of HCl or NaOH solutions and the 19.5 mL of different concentrations (1.03, 2.05, 5.13 and 10.3 mg/L)
ionic strength was adjusted by adding 1 M of NaCl solution. And of fulvic acid solution containing 0.5 g of PET had been completely
then, the vials were mixed thoroughly at 2500 rpm for 30 s and mixed for 72 h prior to the addition of 0.5 mL of CPs stock solution
shaken continuously at 25 ± 1  C and 150 rpm. All experiments to each vial. And the final total concentrations of fulvic acid in the
were conducted in duplicate and the blank controls without PET system were 1.0, 2.0, 5.0 and 10.0 mg/L, respectively.
were also prepared. All samples were filtered through 0.22 mm Taihu Lake water and the Bohai seawater were used as water
syringe filters (polyethersulfone, AB-GEEKS, China) and 1.5 mL matrices to estimate the adsorption capacity of CPs on PET in the
subsamples were collected to analyze the residual CPs. The first natural water systems. The lake water was collected at Wuxi
10 mL filtered solution was discarded to minimize the influence of (12010 E, 31270 N) and the seawater was sampled from Dalian
membrane adsorption (less than 1%). (12190 E, 38 480 N). All water samples were filtered with 0.45 mm
In the adsorption kinetic experiments, 20 mL of 5 mg/L of CPs membranes and stored in the dark at 4  C until further use. Similar
solutions containing 0.5 g of PET were used and subsamples were to the previous experiments, the filtered water samples and PET
collected destructively at different time intervals (0, 0.17, 0.33, 0.5, had been pre-mixed for 72 h prior to CPs addition. More details of
0.75, 1, 2, 4, 8, 12, 24, 36, 48 and 72 h). Adsorption isotherms were the initial conditions of adsorption were listed as annotations un-
determined using 20 mL different concentrations (2, 3.5, 5, 10 and der each figure.
20 mg/L) of CPs with 0.5 g of PET. Samples were collected and
analyzed until the adsorption equilibrium was reached. The equi- 2.4. Models
librium time was determined by the previous kinetic experiments.
To explore the effect of pH on adsorption, the initial pH of CPs The equilibrium and non-equilibrium adsorption quantity of CPs
solution was adjusted to 2.0, 3.0, 4.0, 6.0, 8.0 and 10.0 respectively. on PET were calculated as:
Similarly, CPs with different background ionic strengths (z 0, 0.01,
0.05 and 0.1 M) were prepared to investigate the influence of ionic VðC0  CeðtÞ Þ
strength on adsorption. Fulvic acid was pre-dissolved and passed qeðtÞ ¼ (1)
m
through a 0.45 mm membrane (nitrocellulose, Whatman, UK) to
prepare a 440 mg/L fulvic acid stock solution based on the total where qe (mg/g) and qt (mg/g) are the amounts of adsorbate uptake
per mass of adsorbent at equilibrium and at any time t (h),

Fig. 1. (a) SEM images, (b) N2 adsorption-desorption isotherms with inset the corresponding pore distribution curves and (c) FTIR spectrum of PET microplastics.
4 Z. Liu et al. / Environmental Pollution 265 (2020) 114926

respectively; C0 (mg/L) is the adsorbate concentration of blank Table 1


sample; Ce (mg/L) and Ct (mg/L) are the concentrations of adsorbate Part of adsorption kinetic and isotherm parameters of chlorophenols on PET.

at equilibrium and at any time t (h), respectively; m (g) is the mass Model Parameters Chlorophenols
of adsorbent; V (mL) is the volume of the adsorbate solution. MCP DCP TCP
Pseudo-first-order (PFO) model (Eq. (2)) (Lagergren, 1898),
Kinetic
pseudo-second-order (PSO) model (Eq. (3)) (Blanchard et al., 1984)
Pseudo-first-order qea 39.6 76.8 31.4
and intra-particle diffusion model (Eq. (4)) (Weber and Morris, k1 0.139 0.278 1.24
1963) were used to fit the adsorption kinetic data. R2 0.954 0.818 0.945
Pseudo-second-order qe 41.6 70.3 29.8
qt ¼ qe ð1  ek1 t Þ (2) k2  103 5.31 9.78 125
R2 0.985 0.973 0.996
Isotherm
q2e k2 t Langmuir R2 0.997 0.992 0.989
qt ¼ (3) RMSE 0.086 0.071 0.059
1 þ k2 qe t Freundlich R2 0.997 0.978 0.982
RMSE 0.072 0.126 0.096
pffiffi Linear R2 0.998 0.948 0.928
qt ¼ kp t þ C (4)
RMSE 0.089 0.242 0.365

where k1 (1/h), k2 [g/(mg h)] and kp [mg/(g h1/2)] are the rate con- Adsorption kinetic data after 1 h were used to fix the pseudo-first-order and
pseudo-second-order model. The units of parameters are the same as described
stants of the PFO, PSO and intra-particle diffusion model, respec-
before. R2: regression coefficient; RMSE: root mean squared error. Other fitting
tively; C (mg/g) is a constant associated with the thickness of parameters were provided in Table S1.
boundary layer. a
qe values were determined by the maximum measured values (Hai Nguyen
Langmuir model (Eq. (5)) (Langmuir, 1918), Freundlich model et al., 2017).
(Freundlich, 1906) (Eq. (6)) and linear model (Eq. (7)) were used to
fit the adsorption equilibria data.
macropores. The specific surface area and total pore volume (p/
KL Ce p0 ¼ 0.990) were 0.864 m2/g and 1.83 mm3/g, respectively. The N2
qe ¼ qmax (5) adsorption-desorption isotherm belonged to Type II and the hys-
1 þ KL Ce
teresis loop belonged to Type H3 (Fig. 1b). The types of curves
indicated that the plate-like particles gave rise to slit-shape pores
qe ¼ KF Cen (6)
form, although the pore structure of PET particles was not devel-
oped (Sing, 1982). According to the FTIR spectra (Fig. 1c), hydroxyl
qe ¼ Kd Ce (7) group [3431 cm1 (stretching vibration of OeH)], methylene group
[2966 and 2907 cm1 (asymmetrical and symmetrical stretching
where KL (L/mg), KF [(mg/g)/(mg/L)n] and Kd (L/kg) are the constants
vibration of CeH)], carbonyl group [1713 cm1 (stretching vibration
of Langmuir, Freundlich and linear model, respectively; qmax (mg/g)
of C]O)] and carbon-oxygen single bond [1241 cm1 (stretching
is the maximum saturated monolayer adsorption capacity of the
vibration of CeO)] could be identified (Tobin, 1957). The IEP was
adsorbent; n (dimensionless) is the Freundlich intensity parameter.
estimated to be around 3.6 (Fig. S1).
The zeta potentials and IEP are important surface electro-
3. Results and discussion chemical properties of microplastics and have significant effects on
the adsorption of dissociable organic pollutants and metals
3.1. Characteristics of PET microplastics (Tourinho et al., 2019; Yu et al., 2019). Therefore, it is necessary to
further elucidate the electrification mechanism of microplastics.
The SEM images (Fig. 1a) show that the PET microplastic parti- Generally, the interface between the untreated microplastics and
cles have blocky and layered structure with some cracks and water is inert (Liu et al., 2019b). Unexpectedly, for the polymer

Fig. 2. Adsorption kinetics of chlorophenols on PET fitted by (a) pseudo-first-order, pseudo-second-order and (b) intra-particle diffusion models. Adsorption conditions: initial
chlorophenol concentration 5.0 mg/L; adsorbent dose 25 g/L; initial pH 8.0; ionic strength 0.01 mol/L. Error bars represent the deviation of replicates.
Z. Liu et al. / Environmental Pollution 265 (2020) 114926 5

Table 2
Physical and chemical properties of chlorophenols.

Compound MV (cm3/mol) MD (Å) Sw (g/L) logKow pKa

4-Chlorophenol 99.8 6.82 2.70 2.42 9.1


2,4-Dichlorophenol 111.7 7.08 0.45 3.17 7.8
2,4,6-Trichlorophenol 123.7 7.32 0.08 3.66 6.1

MV: molar volume, inquired from SciFinder at 293 K and 1 atm; MD: molecular
diameter, calculated from MV, assuming that molecules are spherical; Sw: aqueous
solubility from (Davis and Huang, 1990); Kow: octanol-water partition coefficient
from (Davis and Huang, 1990); pKa: dissociation constant from (Davis and Huang,
1990).

surfaces, the IEPs are around 4 and independent of the composition


or hydrophobicity (Zimmermann et al., 2010) which were similar to
the measurements in this study and other literature (Xu et al.,
2018a; Xu et al., 2018b). This phenomenon was likely to be attrib-
uted to the adsorption of hydroxide ions driven by the electrical
potential gradient produced by the first two molecular layers of
water away from the hydrophobic surface (Zangi and Engberts,
2005). The effect of hydrophobicity and multivalent ions on the
Fig. 4. Effect of solution pH on chlorophenols adsorption by PET: experimental data
zeta potentials of the inert interface between polymer and water
and theoretical values. Adsorption conditions: initial chlorophenol concentration
were comprehensively discussed in a previous study (Zimmermann 5.0 mg/L; adsorbent dose 25 g/L; ionic strength 0.01 mol/L; equilibrium time 96 h.
et al., 2010). Error bars represent the deviation of replicates.

3.2. Adsorption kinetics sites of PET, which was likely due to the differences in hydropho-
bicity (Kow, Table 2). The second stage of adsorption could be
The adsorption kinetics of CPs on PET are shown in Fig. 2 and the attributed to the surface and intra-particle diffusion (Hai Nguyen
parameters calculated with PFO and PSO model are listed in Table 1. et al., 2017) of CPs. At this stage, chlorophenols with less chlorine
The adsorption equilibrium was achieved within 72 h (Fig. 2a). The content were more inclined to diffuse into the pores of PET, which
PSO model was more appropriate to describe the adsorption pro- might arise from the differences in the molecule size of CPs (MD,
cess by comparing the R2 values for PFO and PSO models. The intra- Table 2) (Qin et al., 2009). The third stage could be attributed to the
particle diffusion model was also used to fit the data. Apparently, dynamic equilibrium of adsorption. In summary, the film diffusion
the curves did not pass through the origin (C s 0) which meant the controlled the adsorption rate at the early stage, while the intra-
adsorption was not solely governed by a single intra-particle particle diffusion at the later stage. The same phenomenon has
diffusion. However, the data can be fitted well by multiple linear been observed in the literature (Guo et al., 2018; Ma et al., 2019).
curves (Fig. 2b) which suggested that the adsorption process was
controlled by a multistep mechanism. 3.3. Isotherms and effect of pH on chlorophenols adsorption
Generally, the transport process of adsorbate to the adsorption
site of adsorbent can be divided into 4 steps (Weber, 1984): (1) bulk The adsorption isotherms of CPs on PET fitted by different
transport (occurs quickly), (2) film diffusion (occurs slowly), (3) models were provided in Fig. 3. The goodness of fit for the isotherm
intra-particle diffusion (occurs slowly) and (4) adsorption attach- of DCP and TCP on PET followed the order of Langmuir
ment (occurs quickly). Seen in Fig. 2b, the first stage of the model > Freundlich model > linear model by comparing R2 and
adsorption could be attributed to the film diffusion driven by RMSE values (Table 1) (Hüffer and Hofmann, 2016). However, for
physicochemical interactions (such as hydrophobic interaction, MCP, these three models could all fit the isotherm well (Table 1).
covalent forces and so on) (Guo et al., 2015). At this stage, about 22% Fig. 4 shows the effect of solution pH on the adsorption. Kd
of total adsorbed MCP, 36% of total adsorbed DCP and 70% of total (adsorption coefficient, Kd ¼ qe/Ce) was used to evaluate the effect
adsorbed TCP had successfully occupied the exterior adsorption of water hydrochemical properties on CPs adsorption by PET.

Fig. 3. Adsorption isotherms of chlorophenols on PET fitted by (a) Langmuir, (b) Freundlich and (c) linear models. Adsorption conditions: initial pH 8.0; ionic strength 0.01 mol/L;
equilibrium time 96 h.
6 Z. Liu et al. / Environmental Pollution 265 (2020) 114926

As pH increased from 4 to 10, the Kd values decreased sharply P < 0.05) (Fig. S2a). The significant linear correlation indicated that
from 12.2 L/kg to 3.9 L/kg for MCP, from 32.7 L/kg to 0.4 L/kg for DCP the hydrophobic interaction was the main adsorption mechanism
and from 43.3 L/kg to 0.2 L/kg for TCP, respectively. The fraction of at pH 2.
undissociated CPs a can be calculated by Eq. (8). Cation-p, hydrogen-p, polar-p, n-p EDA and p-p interactions are
often used to explain the interactions between adsorbent and
1 adsorbate with aromatic p-system (Keiluweit and Kleber, 2009).
a¼ (8)
1 þ 10pHpKa The adsorption mechanisms involving aromatic p-system of CPs on
When the solution pH was 10, more than 99% of DCP and TCP PET might also be attributed to the interactions between (1) the
were dissociated and the Kd values were close to 0. This indicated hydrogen atom attached to the oxygen of CPs and the p-electron of
that the negatively charged dissociated CPs could form strong aromatic ring of PET (hydrogen-p interaction), (2) the oxygen atom
electrostatic repulsion with negatively charged PET and could not of dissociated CPs and the p-electron of aromatic ring of PET (polar-
be adsorbed. Meanwhile, the dissociation of CPs would increase p interaction) and (3) the p-electron of aromatic ring of CPs and PET
their hydrophilicity. When the pH was 4, more than 99% of CPs (p-p interaction). The existence of the ester group of PET decreased
molecules were undissociated and the Kd reached the maximum the electron density from p-electron of the aromatic ring. This
value. Assuming that the pH value remained the same in the charge transfer made p-electron an electron acceptor (Fig. S2b)
adsorption process and the adsorption isotherm of CPs satisfied the which is difficult to form hydrogen-p interaction with the hydrogen
Langmuir model, the Kd theoretical value could be estimated by Eq. atom attached to the oxygen of CPs (Keiluweit and Kleber, 2009).
(9) derivated from Eq. (1), (5) and (8). The derivation process of Eq. The polar-p interaction was negligible compared with the electro-
(9) is provided in the Supporting Information. static repulsion between dissociated CPs and PET inferred from the
experimental results in the previous section. The increase of chlo-
1 KLð4Þ KLð4Þ rine content resulted in a decrease of the p-electron density of the
¼ qmaxð4Þ  (9) aromatic ring of CPs which made it unfavorable to produce p-p
a Kd 1þmV Kd
interaction between CPs and PET (Fig. S2b). Although low-
where KL(4) (L/mg) and qmax(4) (mg/g) are parameters of Langmuir chlorinated CPs are more prone to form p-p interaction with PET,
model at pH 4 and are listed in Table S2. As is shown in Fig. 4, the p-p interaction may not be the main adsorption mechanism
theoretical values were consistent with the experimental data as compared with hydrophobic interaction and hydrogen-bonding
the pH increased from 4 to 10. In summary, within the range of pH interaction.
that may occur in natural waters, a decrease in pH was favorable for
the adsorption of CPs on PET, especially for the CPs with high 3.5. Effect of ionic strength and fulvic acid on chlorophenols
chlorine content. adsorption
As the pH decrease from 4 to 2, the Kd values of CPs on PET
decreased apparently (Fig. 4). This implied the hydrogen bond As is shown in Fig. 5a, at pH 8, when the ionic strength increased
could form between CPs and PET. With the decrease of solution pH, from 0 to 0.10 mol/L, the Kd values of DCP and TCP decreased from
more and more H3Oþ could compete for the hydrogen-bonding 25.9 L/kg to 23.8 L/kg and 8.5 L/kg to 6.1 L/kg, respectively. This
adsorption sites with CPs (Qin et al., 2012). Compared with CPs, indicated Naþ with better migration ability might compete with the
the mobility of H3Oþ is greater due to its smaller size. Therefore, undissociated CPs on the adsorption sites of PET and form steric
H3Oþ could reach the hydrogen-bonding adsorption sites much hindrance (cation competition). Moreover, the increase of ionic
easier and made the surface of PET positively charged (Fig. S1). The strength led to a decrease in the solubility of organic pollutants (so-
formation of hydrogen bond between CPs and PET could be called salting-out effect) which has been reported to enhance the
attributed to the interaction between the hydroxyl group of CPs and adsorption of organic pollutants on microplastics (Ma et al., 2019;
oxygen-containing functional groups of PET including the carbonyl Zhan et al., 2016). To further elucidate the contribution of the
group and the terminal hydroxyl group. The interaction between salting-out effect on the adsorption of CPs by PET, the effect of ionic
the hydrogen atom and p-electron was also commonly classified as strength on adsorption was also evaluated at pH 4. When the so-
a kind of hydrogen bonds (Davis and Huang, 1990) and this inter- lution pH was 4, the zeta potential of PET was approximately zero
action will be discussed in the next section. If the decrease in (Fig. S1) which could minimize the competition of cations to a great
adsorption when pH reduced from 4 to 2 was only caused by the extent and almost all CPs were undissociated. Fig. S3 showed that
H3Oþ competition, the contribution of hydrogen bonds to the the Kd values did not change significantly with the increase of ionic
adsorption at pH 4 was 37%, 29% and 11% for MCP, DCP and TCP, strength at pH 4. Although the increase in ionic strength decreased
respectively. In other words, the contribution of hydrogen-bonding the solubility of CPs, the reduction was less than 5% (Xie et al.,
interaction to the adsorption of undissociated CPs decreased with 1994). Meanwhile, the salting-out effect became weaker with the
the increase of chlorine content. decrease of CPs concentrations (Kusmierek and Swiatkowski, 2015).
When the pH was 8, more DCP and TCP were dissociated and the
salting effect was negligible. Therefore, the cation competition was
3.4. Effect of chlorine content on chlorophenols adsorption regarded as the main mechanism for low concentrations of DCP and
TCP at pH 8. Notably, in our study, the ionic strength had no sig-
When the solution was at pH 2, the electrostatic interaction and nificant effect on the adsorption of MCP (Fig. 5a). This might result
hydrogen-bonding interaction between CPs and PET could be from both the weak salting-out effect and/or cation competition. A
ignored and the Kd values for MCP, DCP and TCP were 6.9 L/kg, similar phenomenon was also reported in the literature about the
18.9 L/kg and 34.2 L/kg, respectively. This result indicated that there adsorption of MCP on multi-walled carbon nanotubes (Kusmierek
were other interactions between CPs and PET besides electrostatic and Swiatkowski, 2015).
interaction and hydrogen-bonding interaction. Considering the As is shown in Fig. 5b, the existence of fulvic acid negatively
presence of hydrophobic groups (benzene ring, chlorine atom and affected the adsorption of DCP and TCP by PET. When the con-
so on) on CPs and PET, the hydrophobic interaction could not be centration of fulvic acid increased from 0 to 2 mg/L, the Kd values of
excluded. When the pH was 2, the regression of log Kd against log CPs decreased to some extent. When increasing from 2 mg/L to
Kow was log Kd ¼ (0.57 ± 0.10)  log Kow  (0.35 ± 0.29) (R2 ¼ 0.99, 10 mg/L, the Kd values remained almost unchanged compared with
Z. Liu et al. / Environmental Pollution 265 (2020) 114926 7

Fig. 5. Effect of solution (a) ionic strength and (b) fulvic acid content on chlorophenols adsorption by PET. Adsorption conditions: initial chlorophenol concentration 5.0 mg/L;
adsorbent dose 25 g/L; initial pH 8.0; fulvic acid concentration 0 mg/L when evaluating the effect of ionic strength; ionic strength 0.01 mol/L when evaluating the effect of solution
fulvic acid concentration; equilibrium time 96 h. Error bars represent the deviation of replicates. Different superscript letters within a column indicate significant differences
(P < 0.05) among treatments according to one-way ANOVA test.

the Kd at 2 mg/L fulvic acid concentration. Xu and colleagues had experiment was negligible. As is shown in Fig. 6a, there was no
reported a similar trend in the adsorption of 2,20 ,4,40 -tetra- significant difference among the Kd values of MCP in different water
bromodiphenyl ether on polyethylene, polypropylene, polyamide matrices at 0.05 level. The Kd values of DCP and TCP were signifi-
and polystyrene affected by humic acid and fulvic acid content (Xu cantly lower in natural waters than in laboratory water. These re-
et al., 2019). To better explain the effect of fulvic acid, the adsorp- sults were similar to the previous study involving the effect of
tion capacity of fulvic acid on PET was also examed. As is shown in solution ionic strength and NOM content on adsorption. Regardless
Fig. S4, the adsorption capacity of fulvic acid on PET is much larger of the water matrix we used in this study, the adsorption capacity of
than that of CPs observably at pH 8. The qe of fulvic acid increased DCP was much higher than that of MCP and TCP. Compared with
sharply as C0 increased from 0 mg/L to 5 mg/L and the adsorption TCP, more DCP existed in the undissociated form at pH 8, therefore,
tended to saturate gradually when C0 > 5 mg/L. This indicated that DCP is easier to be adsorbed onto PET. However, the hydrophobicity
most of the adsorption sites were occupied rapidly by fulvic acid played a crucial role in the adsorption of MCP and DCP. Although
prior to the addition of CPs, even though the concentration of fulvic the pKa of CPs decreases with the increase of chlorine content
acid might not be high. However, a considerable amount of CPs was (Garba et al., 2019), the contribution of hydrophobicity to adsorp-
still adsorbed to the microplastics. This was very likely owing to the tion cannot be overlooked, especially under acidic conditions.
fact that undissociated CPs could be associated with fulvic acid by Multiple CPs often coexist in the environment, therefore, the
hydrophobic interaction and p-p interaction (Smejkalov 
a and adsorption of CPs mixture in laboratory water was also investi-

Piccolo, 2008; Smejkalov 
a et al., 2009), fulvic acid might act as a gated. The mixture of three CPs was added to the vials right before
bridge with the surface of PET and CPs. The existence of fulvic acid reaction occurred and the equilibrium data for each CP was fitted
or humic acid even enhanced the adsorption of oxytetracycline on using linear model, respectively (Fig. S5). The fitting parameter Kd
polystyrene foam particles mainly attributed to the bridging effect was 10.0 L/kg for MCP (R2 ¼ 0.998), 27.5 L/kg for DCP (R2 ¼ 0.996)
(Zhang et al., 2018). Therefore, the effect of fulvic acid on the and 9.8 L/kg for TCP (R2 ¼ 0.995). The Kd value of each component
adsorption of DCP and TCP on PET was relatively limited but not in the mixed CPs reduced by less than 10% compared with that in a
negligible. solution with the sole component. This indicated that the adsorp-
tion of each CP by PET was less affected by the presence of other CPs
when the concentration of CPs was low.
3.6. Adsorption of chlorophenols on PET microplastics in natural
water matrices
4. Conclusion
More complex interactions were expected between PET micro-
plastics and CPs in natural water. The Kd value of each equilibria In this study, we examined the adsorption behaviors of three
data was analyzed for the statistical difference of the adsorption in typical CPs with distinct chlorine contents on PET under different
different water matrices (Fig. 6) and the linear model was used to pHs, ionic strengths and the concentrations of fulvic acid. The
evaluate the adsorption capacity (Table 3). The properties of the adsorption isotherms were well fitted by Langmuir model for DCP
natural water matrices were also provided in Table 3. The con- and TCP. For the lowest chlorine content MCP, Langmuir model,
centrations of CPs were below the detection limit (10 mg/L) in the Freundlich model and linear model could all fit the adsorption
natural water matrices. It was also reported that the concentration isotherm well. Multiple adsorption mechanisms were identified for
of each CP was less than 1 mg/L in the sampling area (Gao et al., the adsorption of CPs on PET. Generally, hydrophobic interaction
2008). Also, the filtration before the experiment could remove was the main adsorption mechanism of undissociated CPs on PET.
most of the microplastics in natural waters. Thus, the effect of Hydrogen bonds formed between CPs and PET enhanced the
preexisting CPs and microplastics in natural water on the adsorption of the undissociated CPs and the contribution of
8 Z. Liu et al. / Environmental Pollution 265 (2020) 114926

Fig. 6. Adsorption isotherms of chlorophenols on PET in different water matrices fitted by linear model with inset the distribution of adsorption coefficient (Kd) of each point in
different water matrices. To compare with the natural waters, laboratory water was the ultrapure water and the initial pH has been adjusted to 8.0 before experiments. Natural
waters have not been adjusted before experiments.

Table 3
The basic properties of different water matrices and the isotherm parameters of chlorophenols on PET in different water matrices fitted by linear model.

pH Conductivity TOC MCP DCP TCP


(mS/cm) (mg/L)
Kd R2 Kd R2 Kd R2

Laboratory water 8.0 <0.01 Not detected 10.6 0.992 29.4 0.995 10.5 0.983
Bohai seawater 7.9 53.2 4.7 11.4 0.993 21.7 0.995 5.7 0.997
Taihu lake water 8.0 0.55 5.2 10.7 0.990 18.0 0.993 4.6 0.992

TOC: total organic carbon; The unit of Kd: L/kg.

hydrogen-bonding interaction to adsorption increased with the Acknowledgments


decrease of chlorine content. The increase of chlorine content
facilitated hydrophobic interaction between undissociated CPs and The work was funded by the National Natural Science Founda-
PET. Meanwhile, higher chlorine content CPs having lower pKa tion of China (41671468, 41301546), the Natural Science Foundation
values tended to dissociate at neutral pH condition and resulted in of Jiangsu Province (BK20171356), the National Key Research and
stronger electrostatic repulsion with PET. Remarkably, the adsorp- Development Program of China (2018YFA0901200), the Under-
tion of CPs on PET was greatly influenced by pH, with adsorption graduate Innovation and Entrepreneurship Training Program of
capacity increasing rapidly from pH 2 to 4, then decreasing grad- Jiangsu Province (201810286044X), Qing Lan Project of Jiangsu
ually from pH 4 to 10. At pH 8, both ionic strength and the con- Province, Science and Technology Service Network Initiative (KFJ-
centration of fulvic acid did not affect the adsorption of MCP STS-QYZX-051), the Priority Academic Program Development of
significantly, while the adsorption of DCP and TCP decreased with Jiangsu Higher Education Institutions and the Fundamental
the increase of ionic strength or concentration of fulvic acid. These Research Funds for the Central Universities. The authors are
findings were further confirmed in natural water matrices. The grateful to collaborators at the Analytical Center of NIGLAS for their
adsorption capacity of MCP on PET in both Taihu lake water and in-kind support in laboratory facilities.
Bohai seawater was similar to that observed in laboratory water,
while the adsorption of DCP and TCP greatly reduced in Taihu Lake Appendix A. Supplementary data
water and Bohai seawater. This work contributed to the under-
standing of the adsorption mechanisms of hydrophilic compounds Supplementary data to this article can be found online at
on microplastics and adsorption behavior in natural waters. https://doi.org/10.1016/j.envpol.2020.114926.

Declaration of competing interest


References
The authors declare that they have no known competing
Andrady, A.L., 2011. Microplastics in the marine environment. Mar. Pollut. Bull. 62,
financial interests or personal relationships that could have 1596e1605.
appeared to influence the work reported in this paper. Arthur, C., Baker, J., Bamford, H., 2009. Proceedings of the International Research
Workshop on the Occurrence, Effects and Fate of Microplastic Marine Debris.
Sept 9-11, 2008. NOAA Technical Memorandum NOS-OR&R-30.
CRediT authorship contribution statement ATSDR, 2007. Comprehensive Environmental Response, Compensation, and Liability
Act (Cercla) Priority List of Hazardous Substances.
Blanchard, G., Maunaye, M., Martin, G., 1984. Removal of heavy metals from waters
Zheming Liu: Writing - original draft, Data curation, Visuali-
by means of natural zeolites. Water Res. 18, 1501e1507.
zation. Qingdong Qin: Conceptualization, Writing - review & Bornscheuer, U.T., 2016. Feeding on plastic. Science 351, 1154e1155.
editing, Supervision. Zhixian Hu: Investigation, Writing - original Browne, M.A., Crump, P., Niven, S.J., Teuten, E., Tonkin, A., Galloway, T.,
draft. Lu Yan: Investigation. Un-Io Ieong: Investigation. Yan Xu: Thompson, R., 2011. Accumulation of microplastic on shorelines woldwide:
sources and sinks. Environ. Sci. Technol. 45, 9175e9179.
Conceptualization, Methodology, Writing - review & editing, Chen, G.L., Feng, Q.Y., Wang, J., 2020. Mini-review of microplastics in the atmo-
Supervision. sphere and their risks to humans. Sci. Total Environ. 703, 135504.
Z. Liu et al. / Environmental Pollution 265 (2020) 114926 9

Cox, K.D., Covernton, G.A., Davies, H.L., Dower, J.F., Juanes, F., Dudas, S.E., 2019. Olaniran, A.O., Igbinosa, E.O., 2011. Chlorophenols and other related derivatives of
Human consumption of microplastics. Environ. Sci. Technol. 53, 7068e7074. environmental concern: properties, distribution and microbial degradation
Czaplicka, M., 2004. Sources and transformations of chlorophenols in the natural processes. Chemosphere 83, 1297e1306.
environment. Sci. Total Environ. 322, 21e39. Qiao, R.X., Lu, K., Deng, Y.F., Ren, H.Q., Zhang, Y., 2019. Combined effects of poly-
Davis, A.P., Huang, C.P., 1990. Adsorption of some substituted phenols onto hydrous styrene microplastics and natural organic matter on the accumulation and
CdS(s). Langmuir 6, 857e862. toxicity of copper in zebrafish. Sci. Total Environ. 682, 128e137.
Deng, Y.F., Zhang, Y., Qiao, R.X., Bonilla, M.M., Yang, X.L., Ren, H.Q., Lemos, B., 2018. Qin, Q.D., Liu, K., Fu, D.F., Gao, H.Y., 2012. Effect of chlorine content of chlorophenols
Evidence that microplastics aggravate the toxicity of organophosphorus flame on their adsorption by mesoporous SBA-15. J. Environ. Sci. 24, 1411e1417.
retardants in mice (Mus musculus). J. Hazard Mater. 357, 348e354. Qin, Q.D., Ma, J., Liu, K., 2009. Adsorption of anionic dyes on ammonium-
Derraik, J.G.B., 2002. The pollution of the marine environment by plastic debris: a functionalized MCM-41. J. Hazard Mater. 162, 133e139.
review. Mar. Pollut. Bull. 44, 842e852. Rezania, S., Park, J., Md Din, M.F., Mat Taib, S., Talaiekhozani, A., Kumar Yadav, K.,
Freundlich, H.F.M., 1906. Über die adsorption in lo €sungen. Z. Phys. Chem. 57, Kamyab, H., 2018. Microplastics pollution in different aquatic environments and
115e124. biota: a review of recent studies. Mar. Pollut. Bull. 133, 191e208.
Gao, J.J., Liu, L.H., Liu, X.R., Zhou, H.D., Huang, S.B., Wang, Z.J., 2008. Levels and Sing, K.S.W., 1982. Reporting physisorption data for gas solid systems d with
spatial distribution of chlorophenols d 2,4-dichlorophenol, 2,4,6- special reference to the determination of surface-area and porosity. Pure Appl.
trichlorophenol, and pentachlorophenol in surface water of China. Chemo- Chem. 54, 2201e2218.
sphere 71, 1181e1187. 
Smejkalov , D., Piccolo, A., 2008. Host-guest interactions between 2,4-
a
Garba, Z.N., Zhou, W.M., Lawan, I., Xiao, W., Zhang, M.X., Wang, L.W., Chen, L.H., dichlorophenol and humic substances as evaluated by 1H NMR relaxation and
Yuan, Z.H., 2019. An overview of chlorophenols as contaminants and their diffusion ordered spectroscopy. Environ. Sci. Technol. 42, 8440e8445.
removal from wastewater by adsorption: a review. J. Environ. Manag. 241, 
Smejkalov , D., Spaccini, R., Fontaine, B., Piccolo, A., 2009. Binding of phenol and
a
59e75. differently halogenated phenols to dissolved humic matter as measured by
Ge, T.T., Han, J.Y., Qi, Y.M., Gu, X.Y., Ma, L., Zhang, C., Naeem, S., Huang, D.J., 2017. The NMR spectroscopy. Environ. Sci. Technol. 43, 5377e5382.
toxic effects of chlorophenols and associated mechanisms in fish. Aquat. Tox- Stepanova, L.I., Lindstro €m-Seppa €, P., H€
anninen, O.O.P., Kotelevtsev, S.V., Glaser, V.M.,
icol. 184, 78e93. Novikov, C.N., Beim, A.M., 2000. Lake Baikal: biomonitoring of pulp and paper
Guo, X.T., Ge, J.H., Yang, C., Wu, R.R., Dang, Z., Liu, S.M., 2015. Sorption behavior of mill waste water. Aquat. Ecosys. Health Manag. 3, 259e269.
tylosin and sulfamethazine on humic acid: kinetic and thermodynamic studies. Taniguchi, I., Yoshida, S., Hiraga, K., Miyamoto, K., Kimura, Y., Oda, K., 2019.
RSC Adv. 5, 58865e58872. Biodegradation of PET: current status and application aspects. ACS Catal. 9,
Guo, X.T., Pang, J.W., Chen, S.Y., Jia, H.Z., 2018. Sorption properties of tylosin on four 4089e4105.
different microplastics. Chemosphere 209, 240e245. Teuten, E.L., Saquing, J.M., Knappe, D.R.U., Barlaz, M.A., Jonsson, S., Bjorn, A.,
Guo, X.Y., Wang, X.L., Zhou, X.Z., Kong, X.Z., Tao, S., Xing, B.S., 2012. Sorption of four Rowland, S.J., Thompson, R.C., Galloway, T.S., Yamashita, R., Ochi, D.,
hydrophobic organic compounds by three chemically distinct polymers: role of Watanuki, Y., Moore, C., Pham, H.V., Tana, T.S., Prudente, M.,
chemical and physical composition. Environ. Sci. Technol. 46, 7252e7259. Boonyatumanond, R., Zakaria, M.P., Akkhavong, K., Ogata, Y., Hirai, H., Iwasa, S.,
Hai Nguyen, T., You, S.-J., Hosseini-Bandegharaei, A., Chao, H.-P., 2017. Mistakes and Mizukawa, K., Hagino, Y., Imamura, A., Saha, M., Takada, H., 2009. Transport and
inconsistencies regarding adsorption of contaminants from aqueous solutions: release of chemicals from plastics to the environment and to wildlife. Phil.
a critical review. Water Res. 120, 88e116. Trans. Roy. Soc. B 364, 2027e2045.
Hernandez, L.M., Xu, E.G., Larsson, H.C.E., Tahara, R., Maisuria, V.B., Tufenkji, N., Thompson, R.C., Olsen, Y., Mitchell, R.P., Davis, A., Rowland, S.J., John, A.W.G.,
2019. Plastic teabags release billions of microparticles and nanoparticles into McGonigle, D., Russell, A.E., 2004. Lost at sea: where is all the plastic? Science
tea. Environ. Sci. Technol. 304, 838-838.
Hüffer, T., Hofmann, T., 2016. Sorption of non-polar organic compounds by micro- Tobin, M.C., 1957. The infrared spectra of polymers. II. the infrared spectra of
sized plastic particles in aqueous solution. Environ. Pollut. 214, 194e201. polyethylene terephthalate. J. Phys. Chem. A 61, 1392e1400.
Igbinosa, E.O., Odjadjare, E.E., Chigor, V.N., Igbinosa, I.H., Emoghene, A.O., Tourinho, P.S., Koci, V., Loureiro, S., van Gestel, C.A.M., 2019. Partitioning of chemical
Ekhaise, F.O., Igiehon, N.O., Idemudia, O.G., 2013. Toxicological profile of contaminants to microplastics: sorption mechanisms, environmental distribu-
chlorophenols and their derivatives in the environment: the public health tion and effects on toxicity and bioaccumulation. Environ. Pollut. 252,
perspective. Sci. World J. 11. 1246e1256.
Jensen, J., 1996. Chlorophenols in the terrestrial environment. Rev. Environ. Contam. Van, A., Rochman, C.M., Flores, E.M., Hill, K.L., Vargas, E., Vargas, S.A., Hoh, E., 2012.
Toxicol. 146, 25e51. Persistent organic pollutants in plastic marine debris found on beaches in San
Keiluweit, M., Kleber, M., 2009. Molecular-level interactions in soils and sediments: Diego, California. Chemosphere 86, 258e263.
the role of aromatic p-systems. Environ. Sci. Technol. 43, 3421e3429. Velzeboer, I., Kwadijk, C.J.A.F., Koelmans, A.A., 2014. Strong sorption of PCBs to
Kusmierek, K., Swiatkowski, A., 2015. The influence of an electrolyte on the nanoplastics, microplastics, carbon nanotubes, and fullerenes. Environ. Sci.
adsorption of 4-chlorophenol onto activated carbon and multi-walled carbon Technol. 48, 4869e4876.
nanotubes. Desalination Water Treat. 56, 2807e2816. Wang, F., Wong, C.S., Chen, D., Lu, X.W., Wang, F., Zeng, E.Y., 2018. Interaction of
Lagergren, S., 1898. About the theory of so-called adsorption of solution substances. toxic chemicals with microplastics: a critical review. Water Res. 139, 208e219.
K. Sven. Vetenskapsakad. Handl. 24, 1e39. Wang, J., Liu, X.H., Liu, G.N., Zhang, Z.X., Wu, H., Cui, B.S., Bai, J.H., Zhang, W., 2019.
Langmuir, I., 1918. The adsorption of gases on plane surfaces of glass, mica and Size effect of polystyrene microplastics on sorption of phenanthrene and
platinum. J. Am. Chem. Soc. 40, 1361e1403. nitrobenzene. Ecotoxicol. Environ. Saf. 173, 331e338.
Lehner, R., Weder, C., Petri-Fink, A., Rothen-Rutishauser, B., 2019. Emergence of Wang, W.F., Ge, J., Yu, X.Y., Li, H., 2020. Environmental fate and impacts of micro-
nanoplastic in the environment and possible impact on human health. Environ. plastics in soil ecosystems: progress and perspective. Sci. Total Environ. 708,
Sci. Technol. 53, 1748e1765. 134841.
Lin, X.-Q., Li, Z.-L., Liang, B., Zhai, H.-L., Cai, W.-W., Nan, J., Wang, A.-J., 2019. Weber, W.J., 1984. Evolution of a technology. J. Environ. Eng. ASCE 110, 899e917.
Accelerated microbial reductive dechlorination of 2,4,6-trichlorophenol by Weber, W.J., Morris, J.C., 1963. Kinetics of adsorption on carbon from solution.
weak electrical stimulation. Water Res. 162, 236e245. J. Sanit. Eng. Div. 1, 1e2.
Liu, G.Z., Zhu, Z.L., Yang, Y.X., Sun, Y.R., Yu, F., Ma, J., 2019a. Sorption behavior and Wright, S.L., Kelly, F.J., 2017. Plastic and human health: a micro issue? Environ. Sci.
mechanism of hydrophilic organic chemicals to virgin and aged microplastics in Technol. 51, 6634e6647.
freshwater and seawater. Environ. Pollut. 246, 26e33. Wright, S.L., Thompson, R.C., Galloway, T.S., 2013. The physical impacts of micro-
Liu, X.M., Shi, H.H., Xie, B., Dionysiou, D.D., Zhao, Y.P., 2019b. Microplastics as both a plastics on marine organisms: a review. Environ. Pollut. 178, 483e492.
sink and a source of bisphenol A in the marine environment. Environ. Sci. Xie, W.H., Zheng, Z.Q., Tang, M., Li, D., Shiu, W.Y., Mackay, D., 1994. Solubilities and
Technol. 53, 10188e10196. activity coefficients of chlorobenzenes and chlorophenols in aqueous salt so-
Liu, X.W., Zheng, M.G., Wang, L., Ke, R.H., Lou, Y.H., Zhang, X.J., Dong, X.F., Zhang, Y., lutions. J. Chem. Eng. Data 39, 568e571.
2018. Sorption behaviors of tris-(2,3-dibromopropyl) isocyanurate and hex- Xu, B., Liu, F., Brookes, P.C., Xu, J., 2018a. Microplastics play a minor role in tetra-
abromocyclododecanes on polypropylene microplastics. Mar. Pollut. Bull. 135, cycline sorption in the presence of dissolved organic matter. Environ. Pollut.
581e586. 240, 87e94.
Ma, J., Zhao, J.H., Zhu, Z.L., Li, L.Q., Yu, F., 2019. Effect of microplastic size on the Xu, B.L., Liu, F., Brookes, P.C., Xu, J.M., 2018b. The sorption kinetics and isotherms of
adsorption behavior and mechanism of triclosan on polyvinyl chloride. Environ. sulfamethoxazole with polyethylene microplastics. Mar. Pollut. Bull. 131,
Pollut. 254, 10. 191e196.
Ma, Y.N., Huang, A.N., Cao, S.Q., Sun, F.F., Wang, L.H., Guo, H.Y., Ji, R., 2016. Effects of Xu, P.C., Ge, W., Chai, C., Zhang, Y., Jiang, T., Xia, B., 2019. Sorption of polybrominated
nanoplastics and microplastics on toxicity, bioaccumulation, and environmental diphenyl ethers by microplastics. Mar. Pollut. Bull. 145, 260e269.
fate of phenanthrene in fresh water. Environ. Pollut. 219, 166e173. Yu, F., Yang, C.F., Zhu, Z.L., Bai, X.T., Ma, J., 2019. Adsorption behavior of organic
Mato, Y., Isobe, T., Takada, H., Kanehiro, H., Ohtake, C., Kaminuma, T., 2001. Plastic pollutants and metals on micro/nanoplastics in the aquatic environment. Sci.
resin pellets as a transport medium for toxic chemicals in the marine envi- Total Environ. 694.
ronment. Environ. Sci. Technol. 35, 318e324. Zangi, R., Engberts, J.B.F.N., 2005. Physisorption of hydroxide ions from aqueous
10 Z. Liu et al. / Environmental Pollution 265 (2020) 114926

solution to a hydrophobic surface. J. Am. Chem. Soc. 127, 2272e2276. Zheng, W.W., Yu, H., Wang, X., Qu, W.D., 2012. Systematic review of pentachloro-
Zhan, Z.W., Wang, J.D., Peng, J.P., Xie, Q.L., Huang, Y., Gao, Y.F., 2016. Sorption of phenol occurrence in the environment and in humans in China: not a negligible
3,30 ,4,40 -tetrachlorobiphenyl by microplastics: a case study of polypropylene. health risk due to the re-emergence of schistosomiasis. Environ. Int. 42,
Mar. Pollut. Bull. 110, 559e563. 105e116.
Zhang, H.B., Wang, J.Q., Zhou, B.Y., Zhou, Y., Dai, Z.F., Zhou, Q., Christie, P., Luo, Y.M., Zimmermann, R., Freudenberg, U., Schweiß, R., Küttner, D., Werner, C., 2010. Hy-
2018. Enhanced adsorption of oxytetracycline to weathered microplastic poly- droxide and hydronium ion adsorption d a survey. Curr. Opin. Colloid Interface
styrene: kinetics, isotherms and influencing factors. Environ. Pollut. 243, Sci. 15, 196e202.
1550e1557.

You might also like