Chemical Engineering Journal: Yan Laam Cheng, Jong-Gook Kim, Hye-Bin Kim, Jeong Hwan Choi, Yiu Fai Tsang, Kitae Baek

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Chemical Engineering Journal 410 (2021) 128381

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Occurrence and removal of microplastics in wastewater treatment plants


and drinking water purification facilities: A review
Yan Laam Cheng a, 1, Jong-Gook Kim b, 1, Hye-Bin Kim b, Jeong Hwan Choi b, Yiu Fai Tsang a, c, *,
Kitae Baek b, *
a
Department of Science and Environmental Studies, The Education University of Hong Kong, Tai Po, New Territories 999077, Hong Kong
b
Department of Environment & Energy and Soil Environment Research Center, Jeonbuk National University, Jeonju, Jeollabukdo 57896, Republic of Korea
c
State Key Laboratory of Marine Pollution, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China

A R T I C L E I N F O A B S T R A C T

Keywords: Wastewater treatment plants (WWTPs) are important sources of microplastics (MPs), having a high daily
Microplastics discharge of MPs despite a relatively high removal efficiency of MPs. Recently, MPs have been detected in
Removal drinking water purification facilities (DWPFs), indicating a direct threat to humans. In this review, we summarize
Treatment processes
published papers on the occurrence and removal efficiency of MPs, focusing on the status of removal mechanisms
Treated effluent
through different unit operations and processes. For WWTPs, the mass concentrations of MPs in influent ranged
Drinking water
Microfibers from 61 to 5,600 µg/L, and 0.5–170 µg/L of MPs were found in treated effluent, representing an overall removal
efficiency of 93.8%–99.8%. Due to limited publications using mass as MP concentration, the comparison of MP
abundance among studies was based on the number concentration. MPs have been found in influents at levels of
1.01–31,400 items/L, and in treated effluents at 0.004–447 items/L, with a high variation of MP removal from
10.2% to 99.9%. Preliminary and primary treatments contribute to MP removal, while the efficiency of sec­
ondary and tertiary treatments is highly dependent on the applied techniques. Disinfection has a relatively high
potential to breakdown MPs into smaller sizes. For DWPFs, MPs in raw water ranged from 1 to 6,614 items/L,
and 1–930 items/L of MPs were found in treated water, representing an overall removal efficiency of 66.9%–
100% regardless of the treatment types. Yet, the understanding about the role of DWPFs on MP removal is
inadequate. Although many studies on WWTPs have been published, many issues have yet to be resolved, and
further research is warranted on standardization of analytical methods and the removal mechanisms behinds to
develop effective remediation of MPs in WWTPs and DWPFs.

1. Introduction macroplastics and mesoplastics. MPs are either plastic particles between
1 µm and 5 mm in diameter or fragmented macroplastics not exceeding
Plastic production has significantly increased to 288 million tonnes 5 mm [5–7]. MPs can also be categorized by shape, color, and polymer
annually over the last 60 years and is expected to reach 33 billion tons by type. However, no standardized category for MP classification has been
2050 [1]. The problems posed by plastic debris have raised global established [8].
concerns because of widespread distribution and the associated envi­ MPs are classified as either primary or secondary MPs [9]. Primary
ronmental and ecological consequences [2–4]. MPs are a subject of MPs are intentionally manufactured in microscopic sizes for different
mainstream research and have received more attention than applications, whereas secondary MPs are the degradation of plastic

Abbreviations: A2O, Anaerobic-anoxic-aerobic; ASP, Activated sludge process; BPA, Bisphenol A; DBP, Dibutyl phthalate; dw, Dry weight; DWPFs, Drinking water
purification facilities; FTIR, Fourier-transform infrared spectroscopy; GAC, Granular activated carbon; HDPE, High-density polyethylene; MBR, Membrane biore­
actor; MPs, Microplastics; PA, Polyamide; PAM, Polyacrylamide; PE, Polyethylene; PES, Polyester; PET, Polyethylene terephthalate; PP, Polypropylene; PS, Poly­
styrene; PVC, Polyvinyl chloride; RO, Reverse osmosis; ROS, Reactive oxygen species; SBR, Sequencing batch reactor; UV, Ultraviolet; ww, Wet weight; WWTPs,
Wastewater treatment plants.
* Corresponding authors.
E-mail addresses: tsangyf@eduhk.hk (Y. Fai Tsang), kbaek@jbnu.ac.kr (K. Baek).
1
Co-first authors.

https://doi.org/10.1016/j.cej.2020.128381

Available online 4 January 2021


1385-8947/© 2021 Elsevier B.V. All rights reserved.
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

waste by external forces or the breakdown of large particles [6,10]. adsorbent of persistent organic pollutants due to their high specific
Fig. 1 summarizes the pathways of MPs from their points of origin to surface area and strong hydrophobicity [10,22,23]. The adsorption
terrestrial and aquatic environments. Boucher and Friot reported that process occurs through pore-filling or hydrophobic partitioning in­
the major land-based source of primary MPs in marine was road runoff teractions in the interlayer structure of MPs [24,25]. Chemicals,
(66%), followed by treated effluent from WWTPs (25%). Fibers shed particularly hydrophobic organic chemicals, adsorbed onto the surface
from synthetic clothes and microbeads from personal care products are can cause a synergistic toxic effect [26–28]. MPs can also adsorb heavy
discharged into domestic wastewater and then accumulate in aquatic metals from the surrounding environment [29,30]. The small size and
environments if treatment process do not remove MPs from the influent high stability of MPs correspond with a high potential to damage or­
[11,12]. Municipal wastewater typically contains MPs from domestic ganisms. With the high affinity for toxic contaminants, such as organ­
activities (e.g., washing machines and body cleaning), agriculture, and ochlorine pesticides, polycyclic aromatic hydrocarbons, and
industry, all of which are a major source of MPs. Wastewater treatment polychlorinated biphenyls, MPs can be fatal to human beings when
can effectively remove MPs, but the treated effluent still contains a absorbed and inhaled [31–33]. Concentrations of adsorbed pollutants on
considerable amount of MPs, particularly the smallest MPs which are the MPs can reach up to 1 million times that of ambient concentrations.
most difficult to detect and remove [13]. Massive use of plastic products According to the Marine Strategy Frame Instruction Report, the MP-
and poor disposal management has resulted in the ubiquitous presence pollutant association has toxic effects on organisms in the forms of
of MPs in aquatic environments. Plastic waste from landfills or unin­ enzyme production, nutrient dilution, reduced growth rates, and
tentional disposal is another source of MPs in aquatic environments. reproductive failure and death [34].
Although plastics are durable, they are still susceptible to degradation Quantitative data has been used to report the occurrences of MPs and
through mechanical degradation, including oxidation, photo­ calculate the MP removal in WWTPs and DWPFs. For wastewater and
degradation, hydrolysis, and biodegradation at the early stages and drinking water samples, the commonly reported unit of MP abundance
fragmentation afterward [14]. MPs produced by abrasion of larger was items per liter, whereas a few reported as items per cubic meter. For
plastic products can directly enter aquatic environments through surface sludge samples, the commonly reported unit of MP abundance was items
runoff and deposit or accumulate on roads and soil surfaces [15,16]. per gram (dw), whereas a few reported as items per kilogram (dw) or
Tires and coating materials are examples of MPs that erode from the use items per gram (ww). The quantification of MPs is directly affected by
of plastic products. MPs can be ingested by an aquatic organism, float the analytical methods, and most studies used visual identification
from one aquatic system to another, or deposit and accumulate on the through different microscopies and spectroscopies (Tables S1-S3). These
seabed or riverbed [17]. MPs are also light and can enter the environ­ methods provided valuable information in terms of size distribution,
ment through atmospheric fallout. Dris et al. reported that daily atmo­ shape composition, color composition, and composition of polymer
spheric fallout contributes between 2 and 355 items/m2 [18]. Wind types [35]. Due to MP fragmentation, it has been suspected the over­
transfer is another possible pathway for transporting MPs from one estimation of the MP presence in the samples [36]. The mass concen­
location to another [19,20]. However, the sources and release of MPs are tration of MPs has been reported as a conserved base quantity which the
difficult to trace precisely, and no standardized protocol has been physical and chemical characteristics of MPs do not affect the quanti­
established. fication of MPs [36]. Among considered papers in this review, two
Concerns about health risks and ecotoxicological impacts associated published papers reported mass concentration of MPs, which cannot
with MPs have been raised in recent decades [21]. MPs are an efficient consider for comparison in this review.

Fig. 1. Sources and pathways of MPs (modified from [5,14,15,18,20,84]).

2
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Recently, different reviews have been published to summarize the 3. Occurrences and removal of MPs in treatment plants
occurrences and fates of MPs in WWTPs and DWPFs. For WWTPs, re­
views have extensively focused on the methodologies adopted in sample 3.1. Occurrence of MPs in WWTPs
collection, MP extraction, and MP analysis [37–39]. Meanwhile, most
review papers have covered the overall occurrence of MPs in influent 3.1.1. Wastewater
and treated effluent, in terms of MP abundance and MP characterization Municipal wastewater is a major source of MPs in WWTPs [46,47]. A
(i.e., size, shape, and polymer type), and MP discharged to the receiving significant amount of MPs is distributed in WWTPs because of washing
environment [13,37–41]. Some reviews have reported the removal ef­ machines, tire wires, and human activities [11,48–50]. The composi­
ficiency of MPs in different treatment levels or by conventional unit tions of MP sources are also determined by the complexity of the
operations and processes [39–42]. Besides, a few papers have consid­ wastewater collection system and the design flow of the WWTP. The
ered the MP retention in sludge [39,41]. However, these reviews presence of MPs was greatly affected by the treatment levels (pre­
considered the removal of MPs based on the change in MP abundance liminary, primary, secondary, or tertiary treatment as shown in Table 1).
without considering the changes in size distributions, shape composi­ Table 2 summarizes the MP concentrations in influent and treated
tions, and compositions of polymer types. The physiochemical charac­ effluent from WWTPs. Based on number concentration, MPs in the
teristics of MPs might have influences on the removal of MPs by different influent ranged from 1.01 to 31,400 items/L, while 0.004–447 items/L
unit operations and processes, which should be considered when of MPs were found in treated effluent, contributing a high variation of
examining the removal mechanisms. Due to the limited publications in MP removal from 10.2% to 99.9%. The large variations were influenced
the occurrence of MPs in DWPFs, reviews have extensively focused on by MP characterization (i.e., size cut-off and analytical methods) and
the occurrence of MPs in freshwater system and drinking water [43–45]. unit operations and processes of WWTPs (Tables S1 and S4) [55,68].
The discussions about the removal of MPs by water treatment are limited Despite a relatively low MP concentration in treated effluents, the daily
when the results in published papers were mainly the occurrence of MPs discharge of MPs was still relatively high because of the high treatment
in raw water and treated water in DWPFs [43–45]. Besides, the physi­ capacity of WWTPs [39,64]. The average of the total daily discharged
ochemical characteristics of MPs should consider for the removal effi­ MPs was 1.6 × 109 items/L (estimate based on data in Table 2). The
ciency of MPs in DWPFs. estimate showed >1 × 1010 items/L of MPs being discharged from some
This review aims to examine the removal mechanisms of MPs be­ WWTPs, indicating that the total discharged MPs were affected by
hinds the applied unit operations and processes in WWTPs and DWPFs. characteristics of WWTPs.
For WWTPs, it is to explore the occurrence and removal efficiency of Based on mass concentration, MPs in the influent ranged from 61 to
MPs with respect to both wastewater and sludge in terms of size, shape, 5,600 µg/L, while 0.5–170 µg/L of MPs were found in treated effluent,
and polymer type and to evaluate the MP removal performance of each representing an overall removal efficiency of 93.8%–99.8% (Table 2).
unit operation and process in different treatment levels. For DWPFs, it is Recently, MP quantification based on number concentration has been
to assess the occurrence and removal efficiency of MPs in drinking water criticized for exaggerating MP abundance because this involves count­
from the perspectives of size, shape, and polymer type and to determine ing the numbers of suspected particles [36]. For example, one polymer
the effectiveness of MP removal by each unit operation and process in type was PP (12%) based on number concentration, but they possessed
both physical and chemical treatments. 39% in mass concentration [36]. Furthermore, due to the polymer
density and size of MPs, 27%, 14%, and 10% of acrylates, PES, and PE in
2. Methodology number concentration contributed 12%, 8%, and 7% in mass concen­
tration, respectively [36]. The number concentration and mass con­
Since 2009, several studies have conducted field investigations of centration showed opposite results. Lv et al. found that MP
MPs in WWTPs. Recent studies have considered the removal efficiency concentration based on number unit dynamically increased cross-
of MPs by each unit operation and process, while the early studies have treatment systems due to the breakdown of larger MPs [56]. The over­
considered only the abundance of MPs in influent and treated effluent all removal efficiencies of MPs were 53.6% and 82.1% based on number
and the overall removal efficiency of MPs. This review extensively fo­ concentration, but these values varied when reported on a mass basis (i.
cuses on SCI(E), which was published from 2014 to 2020. Thirty papers e., 97% and 99.5%) [56]. Therefore, determining particle number and
on WWTPs, including fourteen on sludge treatment, and three papers on mass simultaneously is more reasonable. MP concentration based on
DWPFs have been published. The keywords included “microplastics,” mass offers a more reliable and representative approach compared with
“wastewater treatment plants,” “drinking water treatment plants,” and counting the numbers of suspected MPs [36,75]. Identification and
combinations of the terms. simultaneous application of criteria for mass and particle number are
For the evaluation of the effectiveness and efficiency of the removal
performance, the major concern was the actual removal rate of MPs
Table 1
reported in published papers. For WWTPs, thirteen published papers
Unit operations and processes of wastewater treatment in typical WWTPs.
were considered because they reported the MP abundance and the
changes in the characteristics of MPs in each unit operation and process. Treatment levels Unit operations and processes
For DWPFs, only one published paper was considered and compared Preliminary Grease chamber; grid (coarse grid and fine grid); grit chamber
with the results from laboratory. Since the MP concentration was re­ treatment (aerated grit chamber and rotary grit chamber); grit lift; oil
chamber; screen (bar screen, coarse screen, fine screen, step
ported based on the number concentration, the changes in size distri­
screen); sieve; skimming tank
butions, shape compositions, and compositions of polymer types were Primary treatment Coagulation-flocculation; sedimentation
also considered to examine the removal efficiency in WWTPs and Secondary Anoxic selector; A2O process; ASP; biofilter; filter bed; media
DWPFs. Furthermore, with the available discussion from laboratory treatment filter; oxidation ditch; sequence batch reactor process; trickling
results, the potential effects of MPs on the performance of unit opera­ filter
Tertiary treatment Biological active filter; decarbonation; disc filter; dissolved air
tions and processes were considered to further examine the
flotation; gravity sand filter; membrane filtration
effectiveness. (ultrafiltration, RO); maturation pond; MBR; nitrification
Disinfection Chlorination; sodium hypochlorite disinfection; ozonation; UV
disinfection

Notes: A2O process: Anaerobic-anoxic-aerobic process, ASP: Activated sludge


process, MBR: Membrane bioreactor, RO: Reverse osmosis, UV disinfection:
Ultraviolet disinfection

3
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Table 2
Overall MP removal efficiency and daily MP discharge in the selected WWTPs.
Number Treatment type Unit operations and processes Country MP concentration MP concentration Removal Discharge References
of in influent in effluent efficiency (items/day)
WWTPs (items/L) (items/L) (%)

1 Secondary Grit and grease chamber, primary Spain 3.20 0.31 90.3 1.09 × 107 [51]
sedimentation, ASP
1 Secondary Screen, grit and oil chamber, France 290 32 88.1 8.40 × 109 [52]
primary sedimentation, biofilter
1 Secondary Bar screen, primary sedimentation, Canada 31.1 0.5 98.4 3.29 × 104 [53]
trickling filter, secondary
sedimentation, chlorination
1 Secondary Step screen, sieve, grit chamber, Finland 57.6 1 98.3 1.00 × 107 [54]
primary sedimentation, ASP,
disinfection*
Pilot-scale Step screen, sieve, grit chamber, 0.4 99.3 –
membrane primary sedimentation, ASP, MBR,
bioreactor disinfection*
1 Secondary Coarse grid, fine grid, aerated grit China 79.9 28.4 64.4 5.68 × 108 [55]
chamber, primary sedimentation,
A2O process, secondary
sedimentation, chlorination
1 Secondary Aerated grit chamber, oxidation China 0.28 (5,600 µg/L) 0.14 (170 µg/L) 50.0 (97.0) 6.5 × 106 [56]
(System A) ditch, secondary sedimentation, UV
disinfection
Secondary Rotary grit chambers, A2O process, 0.06 (30 µg/L) 78.6 (99.5) 3.5 × 106
(System B) MBR, UV disinfection
1 Secondary – Sweden 15.1 0.008 99.9 4.25 × 104 [57]
1 Secondary Coarse screen, fine screen, grit and Scotland 15.7 0.25 98.4 6.52 × 107 [58]
grease chamber, primary
sedimentation, ASP
1 Secondary Screen, primary sedimentation, United 147.0 3.7 97.5 3.1 × 108 [59]
ASP, Sodium hypochlorite States
disinfection
2 Screen, anoxic selectors, ASP, 126.0–147.0 17.3–17.6 86.0–88.2 2.0–3.3 ×
Sodium hypochlorite disinfection 108
1 Secondary Coarse screen, grit lift, fine screen, Korea 29.9 0.4 98.5 1.9 × 104 [60]
primary sedimentation, A2O
process, UV disinfection
1 Coarse screen, grit lift, fine screen, 16.5 0.1 99.1 7.9 × 104
primary sedimentation, SBR, filter
bed, UV disinfection
1 Coarse screen, grit lift, fine screen, 13.9 0.3 98.0 1.1 × 105
primary sedimentation, media filter,
secondary sedimentation, UV
disinfection
7 Secondary – The 68.0–910.0 39.0–65.3 10.2–95.7 – [61]
Netherlands
7
1 Tertiary Coarse screen, grit chamber, Scotland 1–13 1–3 96 2.2 × 10 [62]
primary sedimentation, ASP,
nitrification tank
1 Tertiary Screen, grit and grease chamber, Italy 2.5 0.4 84 1.6 × 108 [63]
ASP, sand filtration, disinfection*
1 Tertiary Bar screen, grit chamber, pre- Finland 610 13.5 97.8 – [64]
aeration, primary sedimentation,
ASP, biological filter
1 Tertiary Coarse screen, grit chamber, Finland 567.8 2.5 99.6 – [65]
primary sedimentation, ASP,
biological active filter
1 Tertiary Bar screen, grit chamber, pre- Finland – 0.03–0.3 – – [66]
aeration, primary sedimentation,
ASP, biological filter, disc filter***
1 Step screen, sieve, grit chamber, – 0.005 – –
primary sedimentation, ASP,
MBR***, disinfection*
1 Preliminary treatment****, primary – 0.1 – –
sedimentation, ASP, dissolved air
flotation
1 Coarse screen, grit chamber, fine – 0.02 – –
screen, primary sedimentation, ASP,
gravity sand filter, UV disinfection
1 Secondary Bar screen, grit chamber, primary United 133.0 5.9 95.6 – [67]
sedimentation, ASP, chlorination States
1 Tertiary Bar screen, grit chamber, primary 83.3 2.6 96.9 –
sedimentation, trickling filter,
secondary sedimentation, ASP,
granular sand filter, chlorination
(continued on next page)

4
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Table 2 (continued )
Number Treatment type Unit operations and processes Country MP concentration MP concentration Removal Discharge References
of in influent in effluent efficiency (items/day)
WWTPs (items/L) (items/L) (%)

Pilot-scale Bar screen, grit chamber, anaerobic 0.53 99.4 –


anaerobic MBR, chlorination
membrane
bioreactor
1 Primary Screen, coagulation-flocculation, Hong Kong 1.01 0.4 60.4 7.32 × 108 [68]
sedimentation, chlorination
1 Secondary Screen, grit chamber, primary 2.06 0.27 86.9 2.24 × 107
sedimentation, ASP, UV disinfection
1 Primary Screen, grit chamber, sedimentation Australia – 0.48 – 4.60 × 108 [69]
1 Secondary Screen, grit chamber, primary – 0.48 – 8.16 × 108
sedimentation, ASP, UV disinfection
1 Tertiary Screen, primary sedimentation, – 0.21–0.28 – 1.00 × 107
biological treatment**, flocculation, –3.60 × 108
dechlorination, ultrafiltration,
membrane filtration (ultrafiltration,
reverse osmosis), decarbonation
1 Secondary Grit removal, centrate thickening United 1 8.8 × 10-4 99.9 9.30 × 105 [70]
system, ASP, chlorination States
-6
7 Tertiary Primary sedimentation, ASP, – 0–2.43 × 10 99.9 7.43–2.08
gravity filter, chlorination × 102
5 Secondary**** Biological treatment** Germany – 0.08–7.52 – 4.19 × 104 [71]
3 Secondary Skimming tank, primary, – – –1.24 × 107
sedimentation, biological
treatment**
1 Tertiary Skimming tank, primary, – –
sedimentation, biological
treatment**, filtration*****
1 Biological treatment**, maturation – –
pond
2 Skimming tank, primary, – –
sedimentation, biological
treatment**, maturation pond
9 Secondary – Denmark 2,223–18,285 19–447 94.9–99.8 – [36]
(61–1,189 µg/L) (0.5–11.9 µg/L) (93.8–99.8)
1 Tertiary – 8,149 19 99.8 –
11 – – China 79.3–342.7 3.6–13.6 89.2–97.1 2.72 × 107 [72]
–1.02 × 108
11 Secondary Coarse screen, primary United – 0.004–0.195 – 5.2 × 104 [73]
sedimentation, ASP, disinfection* States –1.2 × 107
1 Tertiary Coarse screen, primary – 0.009 – 1.0 × 105
sedimentation, ASP, granular
filtration (sand, anthracite coal),
disinfection*
1 Coarse screen, primary – 0.019 – 4.1 × 106
sedimentation, ASP, biological
aerated filter, disinfection*
1 Coarse screen, primary – 0.047 – 1.5 × 107
sedimentation, ASP, granular
filtration (gravel, sand, anthracite
coal), disinfection*
1 Coarse screen, primary – 0.064 – 2.0 × 106
sedimentation, ASP, granular
filtration (gravel, anthracite coal),
disinfection*
1 Coarse screen, primary – 0.092 – 4.1 × 106
sedimentation, ASP, granular
filtration (sand), disinfection*
1 Coarse screen, primary – 0.127 – 9.6 × 106
sedimentation, ASP, granular
filtration (dual media), disinfection
*
1 Tertiary Grit chamber, primary Korea 4,200 33 99.2 8.8 × 108 [74]
sedimentation, ASP, coagulation,
ozonation
1 Tertiary Grit chamber, primary 31,400 297 99.1 1.39 × 1011
sedimentation, ASP, coagulation,
membrane filter (10 µm)
1 Tertiary Grit chamber, primary 5,840 66 98.9 1.37 × 109
sedimentation, ASP, coagulation,
rapid sand filtration

Notes: The value in blanket is the removal rate based on mass concentration. A2O process: Anaerobic-anoxic-aerobic process, ASP: Activated sludge process, MBR:
Membrane bioreactor, RO: Reverse osmosis, UV disinfection: Ultraviolet disinfection
*
The author did not state the details of disinfection adopted in the selected WWTPs.
**
The author did not state the details of biological treatment adopted in the selected WWTPs.

5
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

***
Pilot-scale system was adopted in the selected WWTPs.
****
No primary treatment was adopted before biological treatment.
*****
The author did not state the details of filtration adopted in the selected WWTPs.

required. aquatic environments; and surface runoff is therefore believed to be


another major pathway. Eroded tires and fragmented road paint enter
3.1.2. Sludge freshwater bodies through runoff to roadside drainage [84]. Both
Sludge is another pathway for MPs to enter terrestrial environments, intentional and unintentional disposal of MP-containing products or
such as landfills and soil [10,76]. Sludge treatment consists of thick­ plastic debris contribute secondary MPs to freshwater bodies. Water,
ening, stabilization, conditioning, and dewatering (Table 3). Table 4 wind, or UV radiation can break down plastic debris into secondary MPs
summarizes the MP concentration in all phases of sewage sludge from [14], making MP concentration in raw water a reliable indicator of the
WWTPs. MPs were measured at 18.5–21.2 items/g (dw), 4.4–113 items/ intensity of MP pollution in water supply sources.
g (dw), and 4.0–27.3 items/g (dw) in primary, secondary, and MBR
sludge, respectively. A high effectiveness of sedimentation corresponded 3.3. MP characteristics in WWTPs and DWPFs
to a high MP concentration in raw sludge but low MP abundance in the
treated effluent. The MP concentrations in stabilization were 14.2–15.4 3.3.1. Sizes
items/g (dw), 2.7–170.9 items/g (dw), and 10.0–14.1 items/g (dw) in The term “microplastics” refers to particles of a wide range of sizes
thermal dried sludge, digested sludge, and lime stabilized sludge. No up to a few microns in diameter. The classification of MPs is greatly
comprehensive study is available showing that the sludge treatment influenced by size because indiscriminate definitions affect the strate­
process will cause either MP fragmentation or MP degradation. Instead, gies used in sampling (i.e., sieves employed in field studies) [13]. Filella
all published reports considered the occurrence of MPs at different et al. argued that it is impossible to compare results among multiple
stages of sludge. The MP concentration in dewatered sludge was much studies as neither clear criteria have been adopted to define the size of
higher than in raw sludge, at 9.7–186.7 items/g (dw). Mechanical MPs nor have standardized sampling and extraction protocols been
dewatering may make a high contribution to MP fragmentation, established [85]. Common mesh sizes of sieves range from 0.7 to 500
resulting in a significant increase in MP concentration in dewatered µm, while the detection limit of sizes generally range from 10 to 250 µm
sludge. (Tables S4-S6). The selection of sieves has an influence on MP removal
rates; for example, small MPs may be lost when a sieve with a large mesh
3.2. Occurrence of MPs in DWPFs size is used [54]. Size is an influential factor affecting MP removal by
different unit operations and processes, which one large MP can be
As the list of items used for regular monitoring of water quality does fragmented to several small MPs and therefore a negative removal is
not typically include MPs, and the effects of MP contamination of reported. The differences in size fractions and unit operations and pro­
drinking water on humans is unknown. Few published studies have cesses in WWTPs are associated with diverse removal performances
conducted field investigations of MP concentrations in raw water and among studies, and standardization of size fractions is compulsory for
treated water from DWPFs. Only one field investigation focused on MP the future studies.
occurrence with respect to the unit operations and processes in DWPFs, Most MPs fall in the size range of 10 to 5,000 µm (Fig. S1). As influent
while a few laboratory studies have investigated the effectiveness of passes through each treatment level, larger MPs gradually reduce
coagulation and ultrafiltration on MP removal [79–81]. The presence of whereas smaller MPs gradually increase (Fig. S1). In general, the relative
MPs is greatly affected by the physical and chemical treatments in abundance of MPs > 500 µm in influent is 35.5% and reduces to 33.1%,
DWPFs (Table 5). Table 6 summarizes MP abundance in raw and treated 28.7%, and 28.1% after preliminary, primary, and secondary treat­
water and the overall removal efficiency in different DWPFs. The MP ments, respectively, but tertiary effluent consists of 36.8% of MPs > 500
concentration in raw water varied from <1 to 6,614 items/L, and the µm. Meanwhile, the relative abundance of MPs between 300 and 500 µm
concentration in treated water varied from <1 to 930 items/L, repre­ in influent generally reduces from 30.7% to 2% after tertiary treatment.
senting an overall removal rate of 69.9%–100% [80,83]. Despite a MPs > 500 µm are either removed in different treatment levels or
relatively high rate of MP removal in DWPFs, the relatively high con­ fragmented into smaller MPs. MPs < 300 µm commonly occupy in each
centration of MPs in treated water implies that the treatment methods treatment level at range from 43.5% to 55.7% (Fig. S1). Although
employed in DWPFs are less efficient than those employed in WWTPs. influent only consists of 44.8% of MPs between 20 and 100 µm, this size
MPs have a low mass and are easily transported and deposited on range remains a high level above 70.0% after preliminary treatment.
either land or aquatic environments by wind and rain. Apart from This implies that MPs are more prone to be fragmented into smaller MPs,
contamination from dense population centers or industrial activity, the instead of being removed.
small MPs tend to be introduced through atmospheric fallout or the Several studies showed different size distributions and removal ef­
fragmentation of plastic debris in catchment areas or freshwater sys­ ficiencies of MPs that passed though unit operations and processes
tems. Approximately 3 to 10 tonne of pure polymer fibers are deposited applied [54,55,63,65,69]. Lv et al. compared the preliminary effluent
in every 2,500 km2 annually [18]. Atmospheric fallout and wind transfer from aerated grit chamber and rotary grit chamber, which the former
are considered major contributors to MP occurrence in raw water. No effectively removed all MPs between 250 and 500 µm from influent
treatment of stormwater is typically required before it is discharged to [56]. MPs between 250 and 500 µm increased by 27.3% after rotary grit
removal, while MPs between 125 and 250 and MPs between 25 and 62.5
Table 3 µm were observed in preliminary effluent at 5.3% and 75.8%, respec­
Unit operations and processes of sludge treatment in typical WWTPs. tively [56]. Since no MPs in influent fall in these size ranges, the vortex
water flow in rotary grit chamber contributes to MP fragmentation. The
Treatment Unit operations and processes
types addition of mechanical stirring in grit chamber might also contribute to
MP fragmentation [56]. Therefore, preliminary treatment is likely to
Thickening Dissolved air flotation; gravity belt filter; rotary drum;
sedimentation facilitate the initial fragmentation of MPs or enhance the frangibility of
Stabilization Anaerobic digestion; aerobic digestion; composting MPs.
Conditioning Chemical conditioning (ferric chloride, lime, aluminum sulphate, Primary effluent contains 81.0% and 64.0% of MPs between 20 and
organic polymer); thermal conditioning 250 µm and MPs between 100 and 500 µm, respectively (Fig. S1). Pri­
Dewatering Belt filter press; centrifugation; drying beds; plate filter press
mary treatment influenced the size distribution of MPs mostly among

6
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Table 5

References
Unit operations and processes of water treatment in typical DWPFs.
Treatment types Unit operations and processes

[53]

[54]

[55]
[56]

[57]
[58]
[60]

[61]

[77]
[63]
[65]
[70]
[71]
[78]
Physical Clarification; flotation; GAC filtration; membrane filtration;
treatment rapid sand filtration

14.9 (A2O process); 9.7 (SBR); 13.2


Chemical Coagulation-flocculation
MP concentration in dewatered

treatment
Disinfection Chlorination; ozonation; UV disinfection

Notes: GAC filtration: Granular activated carbon filtration, UV disinfection:


sludge (items/g) (dw)

Ultraviolet disinfection

other treatment levels, which can efficiently remove MPs > 300 µm and
(Media) MPs between 100 and 300 µm at 97.9% and 99.7%, respectively [65].
240.3

186.7
Liu et al. also reported the reduction in MPs > 1,000 µm, but the portion
16.7

22.7
113
<2

47


of MPs between 20 and 300 µm gradually increased [55]. MP frag­
mentation by mechanical erosion might contribute to the reduction of
large MPs [55]. Although larger MPs are more prone to settle in primary
sedimentation, the removal performance of primary treatment might
stabilization (items/g) (dw)
MP concentration in lime

require further investigation.


In general, the removal performance of secondary treatment relies on
the presence of secondary sedimentation. Bayo et al. showed that the
portion of MPs between 200 and 600 µm significantly reduced from 61%
in aeration tank to 38% in sedimentation tank [51]. Lv et al. also indi­
cated the portion of MPs between 62.5 and 125 µm substantially reduced
12.1

from 43.7% in oxidation ditch to 0% in secondary effluent [56]. The size









distribution in secondary sludge reflects the removal performance of


secondary treatment, which is efficient settlement of large MPs.
Generally, large MPs (>250 µm) dominate in secondary sludge (Fig. S2).
MP concentration in anaerobic

Magni et al. reported that three-fourths of MPs in secondary sludge were


>100 µm [63]. Lv et al. indicated that all MPs in secondary sludge were
digestion (items/g) (dw)

>500 µm [56]. Leslie et al. found that 50%–66% of MPs in secondary


Notes: A2O process: Anaerobic-anoxic-aerobic process, MBR: Membrane bioreactor, SBR: Sequencing batch reactor

sludge from three WWTPs were between 10 and 300 µm [61]. Lares et al.
found that 11% of MPs were smaller than 250 µm in secondary sludge
[54]. Small MPs might be entrapped by biological flocs or coated with
biofilm, which unintentionally settled to secondary sludge. However,
170.9

4.0








the removal mechanism of secondary treatment remains unclear.


The removal efficiency of tertiary treatment depends on the pore size
of filter. When the pore size of filter is smaller than MPs, MPs are less
MP concentration in thermal

prone to pass through unit operations and processes applied. Ziajahromi


et al. reported that RO completely removed MPs with a size of 25–100,
MP concentrations in sludge collected from various processes in the selected WWTPs.

190–500, and > 500 µm, and those 100–190 µm were reduced by 20.8%
drying(items/g) (dw)

after RO [69]. Lv et al. showed that the portion of MPs between 62.5 and
125 µm greatly reduced from 62.5% to 16.7% after MBR, which one-
fourth of MPs in MBR sludge were between 62.5 and 125 µm [56].
Lares et al. found that 15% of MPs were smaller than 250 µm in MBR
9.0

sludge [54]. Despite the insignificant difference in the proportions of









small MPs in MBR sludge, MBR is a direct removal method for small
MP concentration in raw sludge

MPs, depending on the pore size of membrane.


76.3 (ww) (secondary sludge)
23.0 (secondary sludge); 27.3

0.7 (ww) (secondary sludge)


0.72 (secondary sludge); 4.0

Large MPs (>1 mm) were commonly found in stabilized sludge


1 (ww) (secondary sludge)
14.9 (primary sludge); 4.4

regardless of the stabilization methods (i.e., anaerobic digestion, lime


stabilization, and thermal drying), while small MPs (<250 µm) were
(secondary sludge)

dominant in dewatered sludge (Fig. S2). Mahon et al. indicated that


(items/g) (dw)

(MBR sludge)

(MBR sludge)

97.7%, 99.2%, and 76.0% of MPs were >1 mm in thermal dried,


digested, and lime stabilized sludge, respectively [78]. Although Lares
et al. found only one-third of MPs in digested sludge were large MPs,
62% of MPs were between 250 and 1,000 µm [54]. These results imply





that sludge stabilization may not contribute to MP fragmentation as


small MPs are rarely identified in stabilized sludge. However, limitations
Number of

on detectable sizes may have contributed to reports of low occurrence of


WWTPs

small MPs (<300 µm). For dewatering, small MPs predominated in


28
1

1
1

1
1
3

1
1
1
6
7

dewatered sludge by 76.0%–89.1% [55,61]. Mechanical dewatering


equipment, such as belt filter presses, plate filter presses, and drying
Netherlands

United States

beds, has a high potential to facilitate MP fragmentation, which can


Germany
Scotland

cause a significant increase in the fraction of MPs < 300 µm. These might
Country

Sweden
Finland

Finland
Canada

Ireland
Table 4

China
China

China
Korea

require further investigation with respect to the fate of MPs during


Italy
The

7
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Table 6
Overall MP removal efficiency and daily MP discharge in the selected DWPFs.
Country Number of Unit operations and processes MP concentration in raw MP concentration in Removal Discharge References
DWPFs water (items/L) treated water (items/L) efficiency (%) (items/day)

China 1 Coagulation-flocculation, 6,614.0 930.0 85.9 1.1 × 1014 [80]


clarification, sand filtration, GAC
filtration, ozonation
Czech 1 Coagulation-flocculation, sand 1,473.0 443.0 69.9 1.4 × 1011 [82]
Republic filtration, chlorination
1 Coagulation-flocculation, 1,812.0 338.0 81.3 2.9 × 109
clarification, sand filtration, GAC
filtration
1 Coagulation-flocculation, flotation, 3,605.0 628.0 82.6 4.8 × 109
sand and GAC filtration
Germany 5 Aeration, filtration, filtration, <1.0 <1.0 100 – [83]
filtration

Notes: GAC filtration: Granular activated carbon filtration

sludge treatment. each unit design facilitate MP fragmentation by mechanical breakdown.


In general, the smallest (<10 µm) MPs predominated by 79.9%– Influent contained 21.5 and 8.54 items/L of fibers and fragments,
90.0% in raw water while MPs>50 µm accounted for <10% in raw water respectively, which was reduced by 92.7% and 89.1% in fibers and
(Figs. S3a and S3b). Mintenig et al. found that all MPs in raw water were fragments, respectively, after primary treatment [53]. Fibers and frag­
between 50 and 150 µm (Fig. S3a) [83]. DWPFs can completely remove ments reportedly accumulate in primary sludge to fractions of 65.1%
large MPs, and no MPs > 50 µm were observed in treated water and 34.2%, respectively, while a few foams and pellets can be found in
[80,82,83]. Complete removal of small MPs was achieved when the primary sludge, at 0.5% and 0.2%, respectively [53]. Since primary
DWPFs only applied triple filtrations without coagulation-flocculation sedimentation removes MPs by gravity, the removal of particular shape
[82]. In general, DWPFs achieved a relatively high removal efficiency corresponds to the polymer type. Without chemical or biological pro­
of 69.9%–81.9% for MPs < 10 µm, but this size still predominated in cesses, the removal efficiency of primary sedimentation is solely
treated water by 75%–99.9% (Figs. S3a and S3b) [80,82]. Among dependent on the density of MPs, which the polymer types of fiber are
physical treatment, GAC filtration is associated with superior perfor­ usually high in density. Besides, fiber might be adhered on the surface of
mance in removing small MPs compared with other techniques suspended solids and unintentionally removed in primary treatment.
(Fig. S3b). It can remove MPs < 10 µm and 10–50 µm at a rate of 56.4%– Chemically enhanced primary treatment can achieve the removal of fi­
61.0% and 50%–100%, respectively [80,82]. Although sand filtration is bers and fragments by 88.9% and 47.7%, respectively [68]. Fibers might
a relatively ineffective method to remove MPs < 10 µm by 31.8%, it can be entrapped by flocs or adhered on flocs, which warrant further
completely remove MPs > 50 µm [80]. After coagulation-flocculation, investigation on the removal mechanism.
dissolved air flotation generally had a better removal performance ASP removed 57.6% and 56.9% of fibers and fragments from primary
than clarification, regardless the sizes of MPs [80,82]. Dissolved air effluent [51,53,54,64,65,68,74]. Particularly, secondary sedimentation
flotation can achieve complete removal of MPs > 50 µm and remove contributed to the removal of 82.0% and 81.1% of fibers and fragments,
MPs < 10 µm and 10–50 µm by 79.3% and 91.3%, respectively [82]. respectively [51]. Similar to primary treatment, the biological flocs
Clarification can achieve removal of MPs > 50 µm and 10–50 µm by collide with all shapes of MPs, enhancing the accumulation rate of fibers
97.0% and 78%, respectively, but it was ineffective to remove MPs < 10 and pellets in secondary sludge. In secondary sludge, fiber is still the
µm at a rate of 28.3%–47.5% [80]. This implies that it is difficult to settle predominant shape by 50%–100%, followed by fragments (5.7%–25%)
small MPs, even coagulant and flocculant are added to facilitate clari­ [53,54,56,61]. Neither foams nor pellets were identified in secondary
fication. Except for coagulation-flocculation, size of MPs is not a factor and MBR sludge [53,54,56,63]. These findings imply that secondary
affecting the removal performance of disinfection. Ozonation and sedimentation has a greater potential to settle fibers and pellets, while
chlorination can remove MPs between 10 and 50 µm by 41% and 50%, primary sedimentation is more effective for settling fragments.
respectively [80]. Yet, ozonation increased MPs < 5 µm by 8.4%, while The removal efficiency of tertiary treatment depends on the unit
chlorination increased MPs between 5 and 10 µm by 4.7% [80]. The operations and process applied. Fibers and fragments were increased by
reduction in MPs > 10 µm might not be due to the removal by either 4.7% and 210.9%, respectively, after biological active filter, because
ozonation or chlorination, but both contribute to MP fragmentation. fragmentation occurred at the fiber-made filter [65]. Meanwhile, fiber is
more prone to longitudinally pass though rapid sand filtration [69].
3.3.2. Shapes Lares et al. found that MPs were composed of 88.3% fibers and 11.7%
Shape is another characteristic to reflect the removal performance by fragments in MBR sludge, whereas Lv et al. found no MP fibers in MBR
different unit operations and processes. The morphologies of MPs sludge, which instead consisted of fragments (75%) and films (25%)
include fiber, fragment, pellet, film, and foam. Other studies classified [54,56]. The variations in shape composition in MBR sludge shows that
MPs into lines, granules, sheets, and spheres [21,86–88]. Some consid­ MBRs may not be able to remove certain shapes, whereas the removal
ered flakes and chips as shapes of MPs [39,57,73]. The surface rough­ efficiency of MBR is a function of size. Despite a relatively high removal
ness of MPs had effects on removal, which fibers and pellets were less rate for fibers and fragments at different treatment levels, fibers and
prone to capture by mechanical methods [8]. Fragments and granules fragments are still dominant in treated effluent at 26.9% and 32.5%,
exhibit angular and twisted morphologies and curved surfaces that are respectively (Fig. S4), reinforcing the need for a better understanding of
prone to be removed [89]. removal mechanisms.
Fiber and fragment shapes generally occupy a large proportion of The composition of MP shapes is similar among different stages of
MPs in influent at 61.7% and 47.8%, respectively (Fig. S4). Lv et al. sludge, except for dewatered sludge (Fig. S5). Fiber is still the dominant
showed that the influent contained no fragment, but the preliminary shape, appearing as 82.4%, 87.0%, and 62.1% of MPs in digested,
effluent from aerated grit chamber and rotary grit chamber contained thermal dried, and lime stabilized sludge, respectively [54,78]. With the
40.9% and 84.1% of fragment, respectively [56]. Although both used exception of a relatively high proportion of fragments (37.0%) in lime
different mechanisms to remove solids, the turbulence of water flow in stabilized sludge, fragments represented 7.5% and 8.4% of MPs in

8
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

digested and thermal dried sludge, respectively [54,78]. Meanwhile, MPs in samples. Gies et al. only verified part of the suspected MPs in
sludges stabilized by different methods contained<3.1% and 0.4% of influent and primary effluent, and no information about polymer types
films and pellets, respectively [54,78]. The insignificant difference in was provided for secondary effluent [53]. Complicating matters further
the shape compositions between raw and stabilized sludge implies that are misleading results from non-MPs. Other than MPs, non-plastics such
sludge stabilization has no effect on removal of certain shapes or MP as glass and silica have been identified because of confusion arising from
fragmentation. Despite relatively high variation in fiber constituents similar colors and shapes [55]. Blair et al. reported that only 39% of
(15%–72.5%) in dewatered sludge, the fragment was the major shape in 1,308 suspected particles were confirmed to be MPs [62]. Bayo et al.
dewatered sludge, appearing as 22.2%–75.8% of MPs, followed by film found that MPs comprised 46.6% of suspected particles, with the bal­
(12.9%–14.0%) [55,57,60,90]. The proportion of pellets increased in ance composed of silicon dioxide, silicates, and cellulose [51]. Failure to
dewatered sludge, with Liu et al. and Li et al. reporting 17.1% and 1.3% consider the suspected MPs particles at each treatment level can lead to
pellet content in dewatered sludge, respectively [55,90]. Mechanical confusion regarding MP removal efficiency. Without standardized
dewatering may therefore contribute to increases in fragment, film, and criteria for MP characterization and identification, it is difficult to
pellet by MP fragmentation. determine which treatment methods have a greater potential to remove
In general, DWPFs are more capable to remove pellets and frag­ low density of MPs.
ments, regardless the unit operations and processes applied (Fig. S6a). More than 23 polymer types of MPs, including 11 co-polymers, have
DWPFs can effectively remove pellets and fragments at rate of 75.7%– been detected in wastewater to date (Tables S4) [51,58,63]. Except PVC,
91.0% and 73.7%–85.5%, respectively, but the effectiveness on fiber PA, PE, PES, PET, PP, and PS are commonly detected in wastewater and
removal varied from 25.0% to 89.2% [80,82]. Fiber removal (25.0%) sludge samples. Generally, polymers of higher density are more common
was the worst when DWPFs adopted only coagulation-flocculation and in raw sludge, while polymers of lower density are more common in
sand filtration (Fig. S6a) [82]. The unit operation and process adopted treated effluent. Lares et al. reported that the portion of PES gradually
after coagulation-flocculation is critical to determine MP removal, with reduced from 90% to 49% between influent and secondary effluent,
clarification tending to remove more fibers and pellets, and flotation while 95% of MPs were PES in secondary sludge [54]. Meanwhile, they
more capable of removing fragments [82]. Although clarification also found that the portion of PE gradually increased from 5% to 36%
removed 54.6% of fibers from raw water, the removal efficiency of between influent and secondary effluent, while only 5% of MPs were PE
fragments and pellets was relatively low, at 36.3% and 29.8%, respec­ in secondary sludge [54]. Although Magni et al. identified a relatively
tively [80]. Since PET particles are mostly in the shape of fiber and PET high proportion of PE (18%) in secondary sludge, two-thirds of the MPs
has 1.38 g/cm3 of density, fibers are more prone to be settled during in secondary sludge contained high-density polymers, including PES,
clarification. In addition, fibers might be entrapped to flocs or adhered polyacrylates (1.05 g/cm3), polyetrafluorethylene (2.2 g/cm3), and
on the surface of flocs, which enhanced the removal of fibers during polyurethane (0.96–1.26 g/cm3) [63]. These results imply that polymer
clarification. type is a crucial factor affecting the retention of MPs in sludge or the
Among physical treatments, GAC filtration had better removal per­ maintenance of MPs in wastewater.
formance than sand filtration, regardless the shapes of MPs (Fig. S6b) Unlike MP fragmentation into small size or angular shape, polymer
[80]. It is simply because the efficiency of sand filtration is corresponded type of MPs does not change to another type because of fragmentation.
to sizes of MPs, while the effectiveness of GAC filtration is related to the This can be an indirect tool to indicate the potential fate of MPs in
polarity of MPs. Since the clarified water consisted of 5.4% of MPs > 10 different treatment level. Despite a relatively high retention of PES in
µm, this limited the removal performance of sand filtration with the secondary sludge, secondary effluent still contained 23% of PES fibers
efficiency no >40% regardless of shapes [80]. Besides, the porous spaces [63]. Simultaneously, fibers increased by 68.0% after secondary treat­
between filter media allowed the penetration of small MPs, which the ment [63]. These imply that PES fibers might be fragmented during
treated water from sand filtration contained 52.6% and 29.5% of fibers treatment process; therefore, the composition of PES and fibers
and fragments, respectively [80]. Meanwhile, PE and PP particles are increased in secondary effluent. Murphy et al. found that the composi­
mostly in the shape of pellets and fragments, which are more prone to be tions of PA, PES, and PP gradually increased from influent to treated
adsorbed during GAC filtration. This achieved removal of pellets and effluent, but no information about the changes in shape composition was
fragments by 81.5% and 66.0%, respectively, which was the most provided [58]. The relationship between shape and polymer type of MPs
effective methods among other techniques [80]. Except from should be carefully considered in the future research on the fates of MPs
coagulation-flocculation, shape of MPs is not a factor affecting the in wastewater treatment.
removal performance of disinfection. Despite the insignificant removal The proportions of polymer types in digested sludge are similar to
of pellets by 1.8%, ozonation increased fibers and fragments by 6.0% those of raw sludge. PES is the major polymer type in digested sludge at
and 7.9%, respectively [80]. Wang et al. explained that the shearing 86%, followed by PA (6%) and PE (4%) [54]. Due to limited results and a
force of the flow induced the breaking of MPs into fibers and fragments lack of standardized criteria, it is difficult to determine whether diges­
[80]. Meanwhile, although chlorination reduced fibers and fragments by tion can degrade polymers. Significant changes among different studies
4.4% and 12.5%, respectively, pellets were increased by 16.9% after were found in the proportions of polymer types in dewatered sludge. PE
chlorination [80]. The increase in particular shape might be due to the and PP were the major polymer types in dewatered sludge, at 40% and
chemical degradation and fragmentation of the corresponding polymer 34%, respectively [55,58,71]. Meanwhile, Liu et al. and Mintenig et al.
type. identified 58% and 0%–20% of PA in dewatered sludge, respectively
[55,71]. A small amount of PES (16.7%) was identified in dewatered
3.3.3. Polymer types sludge [58]. PS particles (7%) were identified in dewatered sludge, but
The term polymer refers to the chemical structures and properties of no PS was found in raw sludge or digested sludge [58,71]. Neither PVC
MPs, which are useful indicators of their fate and the pathways of MPs in nor PET were identified in the dewatered sludge [55,58,71]. Sudden
WWTPs [75]. The composition of polymer types corresponds to shape changes in the compositions of polymer types many warrant further
composition. PES (60%–96.3%) and PA (3.1%–20%) particles are usu­ investigations of the fate of MPs during sludge treatment.
ally fibrous MPs, while PE (63.9%–78%), PP (20%–100%), and PS Since coagulation-flocculation is a common unit operation and pro­
(12%–80%) particles are commonly found in the shapes of fragments cess, the presence of PAM in treated water is significantly relevant to
and pellets [54,56,63]. The reduction of particular polymer type might water treatment. PAM is added as flocculant to facilitate the floc for­
indicate the removal of the corresponding shape. Yet, the composition of mation, but no unit operation and process can completely remove PAM
polymer types is rather difficult to compare, because studies have from water [80,82]. GAC filtration is associated with superior perfor­
adopted random sampling on identification and validation of suspected mance in removing PAM particles compared with other techniques [80].

9
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Besides, GAC filtration had the highest removal rates of PET, PE, and PP 3.4. Removal of MPs in WWTPs
particles compared with other techniques [80]. Unlike other physical
treatments, the removal mechanism of GAC filtration is depended on the Preliminary, primary, secondary, and sometimes tertiary treatments
polarity of solids, in other words, the polymer type is a significant factor are used to purify wastewater with physical, chemical, and biological
affecting the removal efficiency of GAC filtration. Since PE and PP treatments and to meet discharge standards for biochemical oxygen
particles are non-polar MPs, GAC filtration achieved 77.6% and 62.1% demand and total suspended solids (Table 1). Drawing on data from 13
of removal on these MPs [80]. Meanwhile, the efficiency of clarification studies, the removal efficiency was evaluated, with influent MP quan­
is also depended on the polymer types, which density of MPs is associ­ tities varying from 1.01 to 31,400 items/L. The removal efficiency of
ated with the settlement. Clarification was responsible to settle 52.7% of MPs by preliminary treatment reached 6.0%–58.6%. Primary treatment
PET particles, which shared equal contribution with GAC filtration [80]. contributed an extra 19.1%–99.0% to removal, whereas MP elimination
Although PE and PP particles have<1 g/cm3 of density, clarification in secondary treatment varied from − 66.7% to 92.6% (Fig. 2). Mean­
achieved 48.8% and 47.4% of removal, respectively [80]. This implies while, the removal efficiency by tertiary treatment ranged from –72.7%
that coagulation-flocculation plays a significant role in agglomeration of to 99.9% (Fig. 2). Unit operations and processes in each treatment level
MPs to flocs, which facilitate the removal of light-density MPs during and the characteristics of MPs contributed to different removal effi­
clarification. ciencies. This heterogeneous value can be attributed to diverse unit
Polymer type of MPs is also crucial to examine the effects of chemical operations and processes associated with different treatment levels
treatment on MPs. Undoubtedly, both ozonation and chlorination have (Fig. 2). The removal efficiency differed according to treatment level.
no effects on MP removal when the reduction of each polymer type is
<7% [80]. Additionally, these might induce chemical degradation and 3.4.1. Preliminary treatment
thereby accelerate MP fragmentation. PET and PE particles increased MP removal in preliminary treatment has received relatively little
after ozonation by 10.2% and 5.4%, respectively, while chlorination attention, and few studies have considered the removal efficiency of MPs
increased PP particles by 4.1% [80]. Yet, the effects of ozonation and by preliminary treatment [56,58,62,67]. Unit operations and processes
chlorination on MPs might require further investigation. Despite a in preliminary treatment are not designed to remove MPs from influent,
relatively high removal of PET particles by 44.3%, the effectiveness of but 6.0%–58.6% of MPs are unintentionally removed (Table S7). The
sand filtration is not corresponded to the polymer type of MPs. Since the typical screen size is >6 mm; therefore, MPs may be removed when they
removal of fibers was slightly higher than the removal of other shapes, are attached to large debris or entrapped in large debris [67]. Yet, it is
the removal of PET particles was correspondingly higher than the still possible for direct removal of MPs if fine screen or microscreen are
removal of other polymer types. employed in WWTPs.
In general, preliminary treatment achieved a relatively high removal

Fig. 2. Removal efficiency of MPs in different unit operations and processes in the selected WWTPs.

10
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

of MPs. Michielssen et al. reported that the removal efficiencies of pre­ 76% [62]. Besides, Murphy et al. indicated a moderately high removal
liminary treatment in two WWTPs were 58.6% and 35.1%, which both efficiency of MPs by sedimentation (i.e., 60.9%) [58]. The absence of
WWTPs have adopted bar screen and grit chamber for preliminary PVC particles and a few PS particles (5.9%) were observed in primary
treatment [67]. Despite no detail information about the changes on size effluent, whereas it contained 14.7% of PE particles [58]. Since the
distribution, size composition, and composition of polymer type, MPs removal mechanism of sedimentation is the settlement of suspended
might be adhered on the surface of large debris or unintentionally solids by gravity, high-density of MPs are settled by sedimentation,
captured along with large debris during preliminary treatment. Murphy while the light-density of MPs pass through sedimentation. This implies
et al. also found that screening and grit/grease removal contributed to a polymer type is a major factor affecting the removal efficiency of sedi­
removal rate of MPs at 44.6%, while grit and grease contained <2 and mentation. Both PES and PA particles have density higher than 1 g/cm3,
7.9 items/g of MPs, respectively [58]. PE particles were substantially but abundant of PES fibers (29.4%) and PA fibers (14.7%) were observed
removed at a rate of 85.2%, followed by PP particles (74.4%) and PA in primary effluent [58]. Additionally, Bayo et al. reported a poor
particles (71.7%) [58]. Besides, abundant PE microbeads from a facial removal performance of MPs by sedimentation (i.e., 19.0%), which fi­
cleanser were observed in the grease at a rate of 32.2% [58]. Although bers increased by 21.4% comparing with influent [51]. Even the density
the microscopic sizes of MPs are more prone to pass through preliminary of MPs is higher than that of wastewater, shape, especially fibrous MPs,
treatment, the light-density of MPs float on the surface and water and might be a significant factor when considering the removal performance
are ready to be skimmed during grease removal [58]. The removal of of sedimentation. Furthermore, Lares et al. reported an almost complete
MPs by preliminary treatment is therefore unintentional results of removal of MPs by sedimentation with a rate of 99.0% [54]. Over 98% of
screening and grit/grease removal. removal was found in each size range (i.e., 1–5 mm, 0.5–1 mm, 0.25–0.5
Considering the design of preliminary treatment, the performance of mm, and < 0.25 mm); thereby sizes of MPs did not affect the removal
MP removal in preliminary treatment varies. Blair et al. reported that the efficiency by sedimentation [54]. To conclude, sedimentation is a sig­
removal efficiency was only 6% in preliminary treatment, but they did nificant unit operation and process in removing MPs, which shape and
not provide a detailed explanation [62]. Since the coarse screen of polymer type of MPs might influence the removal performance by
selected WWTP has targeted to remove solids >12 mm, MPs are prone to sedimentation.
pass through screening. In the study of Murphy et al., the targeted sizes Another method employed in primary treatment is coagulation-
of solids are >19 mm and 6 mm, which might be possible to remove flocculation, which is a chemically enhanced sedimentation [93,94].
larger MPs. Furthermore, the removal performance of MPs by grit/ This can achieve 50.0%–97.9% of MP removal (Fig. S2). The addition of
grease chamber depends on the designs of unit operation and process. coagulants such as alum or ferric chloride and gentle stirring facilitate
MP concentration in aerated grit chambers was reduced by 21.4%, floc formation, enhancing sedimentation by increasing particle sizes
whereas that in a rotary grit chamber increased by 371.4% [56]. In the [42]. Although Talvitie et al. considered the removal efficiency of
aerated grit chamber, the addition of air causes a spiral water flow, and coagulation-flocculation combined with the preliminary treatment, the
both heavier grit and lighter organic particles are thrown out and settled removal efficiency of MPs varied from 50% to 97.9% [64,65]. Ruan et al.
after hitting the side wall. The turbulence in water flow might facilitate found that the chemically enhanced primary sedimentation had MP
MP removal, when the light-density of MPs is pull in the turbulence and removal by 78.2%, whereas the traditional sedimentation only reached
entrapped with heavier solids. Rotary grit chamber creates vortex water 41.7% [68]. Based on the changes in shape composition comparing with
flow, which solids with a specific gravity of 2.4 to 2.65 are settled by influent, coagulation-flocculation can achieve the removal of fibers and
gravity. The vortex water flow might pull down light-density of MPs, the films by 93.2% and 86.3%, respectively, while the removal of fragment
friction force from mechanical stirring causes MP fragmentation. Rotary varied from 32.4% to 96.4%. Fibers might be entrapped during floc
grit chamber removed 35.3% of MPs > 0.5 mm, the dominant portion of formation or adhered on the surface of flocs, which accidentally
small MPs (25–62.5 µm) in the preliminary effluent implied a break­ removed by coagulation-flocculation. This requires further investigation
down of MPs from mechanical stirring in the rotary grit chamber [56]. in laboratory to verify whether coagulation-flocculation achieve higher
Preliminary treatment may therefore facilitate the initial fragmentation removal of MPs in terms of shapes.
of MPs in wastewater.
3.4.3. Secondary treatment
3.4.2. Primary treatment Secondary treatment is typically used to remove suspended or dis­
Studies have generally considered the MP removal in primary solved organic matter and nutrients through different biological unit
treatment combining with preliminary treatment, which achieved a operations and processes. Sun et al. summarized that secondary treat­
relatively high removal of MPs at a rate of 40.7%–97.9% ment achieved 0.2%–14%, while Gatudou et al. indicated secondary
[53,55,64,65,74]. As shown in Fig. 2 and Table S7, 19.1%–99.0% of MPs treatment contributed to additional removal by 7%–20% [37,39].
are removed in primary treatment, which is the major level to remove However, the removal efficiency of MPs varied from –66.7% to 92.6% in
MPs from wastewater. The common method adopted in primary treat­ secondary treatment (Fig. 2 and Table S7). During biological treatment,
ment is sedimentation. Circular and rectangular tanks are common de­ bacteria can attach to and form a biofilm layer on MP surfaces, causing
signs for sedimentation, which circular designs are associated with the bacteria-associated MPs to settle in a final sedimentation tank as
superior sedimentation performance [91]. No study has examined MP secondary sludge [51,95]. This suggests that MPs could be redistributed
removal rate according to the designs of the sedimentation tank, but MPs from the liquid to solid phase. However, few studies have investigated
are prone to remove with suspended solids [92]. After primary treat­ MP-microorganism interactions (such as biofilm formation, aggregation,
ment, low density of MPs remains in the primary effluent and can be and inhibition) or the mechanisms behind MP removal by biological
released to the environment or secondary treatment, while settled MPs treatment. MPs may leach additives to wastewater or adsorbed organic
are transferred to disposal or sludge treatment. Future research should pollutants and heavy metals from wastewater. This leachate may inhibit
not only consider the MP removal in preliminary and primary treat­ or suspend the growth of microorganisms or the formation of biological
ments separately, but also estimate the removal efficiency of each unit floc. However, the fates of MPs in biological treatment remains unclear.
operation and process in preliminary treatment. ASP has a superior removal performance of MPs in secondary
Sedimentation in primary treatment is the major contributor of MP treatment. The removal efficiency of MPs by ASP varied from –66.7% to
removal along WWTPs, which achieves 19.1%–99.0% of removal 92.6% (Fig. 2). During ASP, MPs are entrapped in biological flocs and
(Fig. S2). Michielssen et al. reported a relatively high MP removal by then be separated from primary effluent [51,65,74]. Unlike primary
sedimentation in two WWTPs (i.e., 61.6% and 82.1%) [67]. Blair et al. sedimentation, secondary sedimentation achieved an average of 71.3%
also reported a high MP removal by sedimentation with a rate of 60%– fiber removal [51,64,65,68,74]. The sedimentation after aeration allows

11
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

the settlement of biological flocs, which suspended fibrous MPs might be During high-pressure backwash, some MPs can pass through the filter,
further entrapped by the moving down flocs or adhered on the surface of resulting in lower removal efficiency [74]. The removal efficiency of
biological flocs. Bayo et al. showed that the MP removal by aeration MPs is highly dependent on the size of the filter pores, which MBR has
achieved only 17.8%, whereas secondary sedimentation enhanced MP the best MP removal among the techniques of physical filtration. The
removal by 85.4% [51]. This implies that sedimentation might be an porosity of filter media in rapid sand filtration is much larger than in
important unit operation and process in secondary treatment, which other techniques, MPs are more prone to passing through this unit
allow an additional MP removal during settlement of biological flocs. operation and process.
However, Lares et al. reported a negative removal of MPs by ASP at a In contrast, tertiary treatment does not guarantee superior perfor­
rate of –66.7% [54]. With the availability of sedimentation, secondary mance in MP removal compared with other treatment levels. Mason
treatment serves to remove the remaining MPs from primary effluent, et al. found no clear correlation between tertiary treatment and MP
which the sizes of MPs are smaller, and the density of MPs are lighter. removal after examining 17 different WWTPs in the United States [73].
Apart from ASP, a few studies have considered the removal efficiency The pore size of RO systems is the smallest among the membrane
of MPs by other unit operations and processes in secondary treatment. techniques and should exhibit the best performance of MP removal.
Oxidation ditch is a relatively effective unit operation and process to However, Ziajahromi et al. found that RO was associated with an un­
remove MPs, which achieved 95.6% [56]. Most MPs, especially 92.3% of expectedly low removal efficiency of 2.8% but provided no detailed
fibers, were entrapped by activated sludge [56]. However, based on explanation [69]. Talvitie et al. reported that a biological active filter
mass concentration, the removal efficiency by oxidation ditch only was ineffective at removing MPs, contributing to a negative removal of
reached 16.5% [56]. In general, A2O process is less effective in removing 72.7% [65]. Furthermore, Carr et al. showed that MP removal perfor­
MPs with a variation from –21.2% to 28.1% (Fig. 2). To note, A2O mance by gravity filtration in bench-scale research was superior to
process achieved 15% of MP removal based on mass concentration [56]. practices in WWTPs, and no MPs were detected in bench-scale backwash
The recirculated sludge in A2O process might contribute to the relatively [70]. Tertiary treatment therefore contributes both positive and nega­
low MP removal by A2O process, which MPs were accumulated within tive effects to MP removal.
A2O process [56]. Both studies showed an increased in fibers from
22.1% to 257.1%, while MPs between 0.1 and 1 mm increased by 3.4.5. Disinfection
115.8% [55,56]. During A2O process, stirring in anaerobic and anoxic Chlorination, ozonation, and UV radiation are disinfection methods
tanks and aeration in aerobic tanks might enhance the friction force in employed in WWTPs before discharging the treated effluent [42,96].
the water, which cause MP fragmentation. Yet, this requires further These disinfectants might degrade MPs or break them into smaller sizes,
investigation. and even nanoplastics, before they degrade in the environment. How­
ever, no study has focused on the fate of MPs before and after disin­
3.4.4. Tertiary treatment fection and our understanding of the effect of disinfection on MP
Tertiary treatment is the final cleaning process to further improve the removal is limited.
quality of treated effluent before it is reused or discharged to the envi­ Ozone is a strong oxidant that can be converted into radicals causing
ronment [71]. Either advanced techniques (membrane filtration, MBR, degradation of polymers [97]. Ozonation can reportedly contribute to a
dissolved air flotation, and activated carbon adsorption) or conventional removal rate of MPs at 89.9%, and a decrease in MP concentrations to 33
techniques (sand filtration and disc filters) are employed, depending on items/L [74]. Meanwhile, ozonation can contribute to decomposition of
site-specific needs and opportunities to use the resulting effluent MPs, but no detail explanation was provided [98]. To the best of our
(Table 1). Tertiary treatment provides an additional MP removal op­ knowledge, ozonation can degrade MPs within a brief contact time, but
portunity for WWTPs, with a removal efficiency from –72.7% to 96.3% the fate of MPs before and after chlorination has not been yet investi­
(Fig. 2). Filtration (MBR, gravity bed filter, sand filtration, and disc fil­ gated in detail.
ters) is a common strategy used in tertiary treatment. MBR is the most Due to large surface-to-volume ratios, MPs are more prone to re­
effective technique for MP removal when the pores of the membrane are actions in chlorination. Michielssen et al. found that chlorination
smaller than 0.1 µm [56]. Talvitie et al. found that MBR achieved nearly contributed to a removal rate of MPs at 7%, but no explanation was
complete (99.9%) removal of MPs, with the treated effluent containing provided [67]. Liu et al. found that MP concentrations were reduced by
only 0.005 items/L of MPs [66]. Lv et al. investigated a parallel system 7.1% after chlorination, and the treated effluent contained 28.4 items/L
in secondary WWTPs, reporting that one of the systems that adopted of MPs [81]. Kelkar et al. stated that PP, HDPE, and PS are resistant to
membrane filtration to replace traditional sedimentation achieved chlorination and their chemical and physical structure does not change
80.5% MP removal in treated effluent (0.028 items/L of MPs) [56]. [99]. However, Ruan et al. reported that MP abundance increased by
During MBR, MPs is directly captured by the membranes, which the 81.8% after chlorination, because intense redox reactions contributed to
permeate contains a few amount of MPs. However, the presence of MPs the breakdown of MPs [68]. Low density of MPs is prone to breakdown
might induce frequent clogging in MBR, which require backwash to during chlorination [68]. The fate of MPs during chlorination is uncer­
clear the sludge cake layer. During backwash, MPs might accidentally tain, but the chemical interactions between MPs and chlorination leads
release and pass through tertiary treatment. Thus, the removal perfor­ to the formation of new bonds and structures and alters the adsorption
mance of MBR varies, depending on the frequency of backwash and the mechanisms between MPs and other organic pollutants.
pore size of membrane. The fate of MPs before and after UV disinfection has yet to be
Apart from MBR, the removal efficiency of rapid sand filtration and investigated in detail [56,68]. Lv et al. reported that UV disinfection
disk filters varies among published reports. Since the porosity of filter contributes to a removal efficiency of 1%, breaking MPs into smaller
media is larger than MPs, MPs are more prone to pass though rapid sand sizes by oxidation [56]. Photochemical reactions during UV disinfection
filtration. However, the removal efficiency of MPs by rapid sand filtra­ enhances the frangibility of MPs by reducing elasticity and absorbing
tion ranged from 55.6% to 95.0% (Fig. 2). During rapid sand filtration, light [16]. In addition, Naik et al. reported that UV radiation can alter
MPs are entrapped by filter media (i.e., sand grains) or adhered on the the surface of HDPE, transforming smooth MPs into fibrous particles
surface of filter media [66]. Meanwhile, the removal efficiency of MPs [100]. Contrary to HDPE, PS and PP polymer types did not physically
by disc filters varied from 40.0% to 98.5% (Fig. 2). The removal per­ degrade [100]. Due to rough surfaces and the presence of oxygen-
formance by disc filters depends on the pore size of filters. Remarkably, containing functional groups, UV irradiation accelerated the aging of
the MP removal by 10 µm sized disc filter (i.e., 40.0%–79.6%) was worse PS, and enhanced the adsorption capacity of MPs after aging [101]. The
than 20 µm sized disc filter (i.e., 98.5%) [66,74]. The presence of MPs effects of UV disinfection, especially on molecular structure, are influ­
induced clogging in disc filters, requiring frequent backwash [74]. enced by the polymer types of MPs.

12
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

3.4.6. Factors affecting MP removal in WWTPs been found: a high MP concentration corresponds to a high concentra­
Overall, the removal efficiency of MPs varies among different tion of BPA, and less methane is produced [102]. Wei et al. reported
treatment levels. First, the operating conditions of unit operations and finding DBP leachate during alkaline anaerobic fermentation in the
processes affect overall MP removal. For example, high MP concentra­ presence of PET particles [103]. Almost 2.4 µg/L of DBP was released
tions accelerate clogging in membrane techniques, which requires when the concentration of PET particles was 60 items/g [103]. The
frequent backflow and reduces removal efficiency of the membrane presence of MPs can also affect hydrogen production in alkaline
techniques [74]. The same technique with different operating conditions anaerobic fermentation. Wei et al. reported that a high level of PET in
(e.g., hydraulic retention time, flow rate, and treatment capacity) can anaerobic digestion limited the availability of protein, polysaccharides,
alter MP removal. Lv et al. indicated that the variations of MP removal glutamate, glucose, and butyrate, which restrained hydrolysis, acido­
between two preliminary systems were caused by different hydraulic genesis, and acetogenesis, and therefore reduced hydrogen production
retention times [56]. Second, the characteristics of the influent deter­ [102]. In the presence of PET, microbial populations in hydrogen pro­
mine the MP concentrations and characteristics of MPs in the influent ducers and hydrolytic microbes were limited, and the hydrogen yield
and affect the overall removal efficiency of MPs. Gündoğdu et al. re­ was limited to 70.7% [103].
ported that influent is highly dependent on population, flow, seasonal
variations, and sources (i.e., domestic wastewater or industrial waste­ 3.6. Removal of MPs in DWPFs
water) [46]. The presence of MPs in industrial influent was 1.8 times
greater than that from domestic influent [8]. Bayo et al. described a Although DWPFs can achieve an overall MP removal rate of 69.9%–
relationship between seasonal variability and MP abundance [51]. MP 100%, a substantial amount of MPs has been found in treated water
concentrations among seasons showed significant statistical differences, [80,82,83]. A DWPF employing coagulation-flocculation and sand
with summer higher than other seasons [51]. In general, composition of filtration achieved a removal efficiency of MPs at 69.9% [82]. Mean­
polymer type detected in a WWTP reflects daily life of the people the while, there was no significant difference in rates of MP removal when
plant serves [13,57,70,72]. Influent from urban WWTPs may contain DWPFs applied either clarification or flotation after coagulation-
primarily PE, PP, PA, and PES, because of the use of personal care flocculation. DWPF with clarification removed 81.3% of MPs while a
products and toothpaste, and the washing of synthetic clothes similar facility with flotation removed 82.6% [82]. To the best of our
[58,69,70,73]. In contrast, industrial wastewater contains high amounts knowledge, only Wang et al. investigated MP removal in each treatment
of PA, PE, and PP [55,72]. Third, different unit operations and processes method used in DWPFs, and only Ma et al. investigated MP removal in
contribute to different removal mechanisms of MPs among treatment water through coagulation-flocculation and ultrafiltration in the labo­
levels. Table 1 shows that various unit operations and processes are ratory [79–81]. Wang et al. found the overall MP removal in advanced
employed in WWTPs, depending on site specifics. The removal mecha­ DWPFs reached 82.1%–88.6% [80]. Despite multiple reports of signifi­
nism in preliminary treatment is mainly unintentional removal by cant MP abundance in raw water, the removal efficiency of advanced
screening and grit/grease removal through entrapment between MPs DWPFs is slightly higher than that of conventional WWTPs [82].
and solids or adherence to the surface of solids, whereas MP removal in Generally, chemical treatment contributes between 40.5% and 54.5% of
primary treatment (i.e., sedimentation) is determined by the polymer MP removal, and physical treatment contributed another 29.0% to
type and shape. Secondary treatment (i.e., biological treatment) 60.9%, but disinfection resulted in MP fragmentation at a rate between
removes MPs from wastewater by exploiting interactions between mi­ –9.3% and 6.8% (Fig. 3). Fig. 3 provides overall performances and
croorganisms and MPs (e.g., biofilm formation) and the presence of removal efficiencies of each treatment, based on data from Wang et al.
secondary sedimentation. The effectiveness of tertiary treatment varies [80].
among unit operations and processes being employed and cannot
guarantee significant MP removal. In particular, the effectiveness of 3.6.1. Physical treatment
filtration depends on the pore sizes of the filter media and the frequency Physical treatment is the major method adopted in DWPFs, which
of clogging. In general, different factors, including site-specific WWTPs achieves 29.0%–60.9% of removal (Fig. 3). To settle flocs from
(design of treatment levels, treatment capacity, and flow rate), MP coagulation-flocculation, conventional DWPFs employ clarification. MP
characteristics (size, shape, and polymer type), and operating conditions removal reached 40.5%–54.5% when coagulation-flocculation was
of unit operations and processes contribute to performance of MP combined with clarification (Fig. 3). Complete settlement of MPs > 10
removal. Considerable research is needed on other parameters affecting µm and a relatively high settlement of MPs between 5 and 10 µm
MP removal. (44.9%–75.0%) were observed, especially fibrous MPs [80]. This
implied that coagulation-flocculation enhances the aggregation of larger
3.5. Sludge treatment MPs with flocs, which increases the density and facilitates clarification
[80]. However, lab-scale research showed that clarification after
Despite the lack of comprehensive studies of the occurrence of MPs coagulation-flocculation had poor removal performance of PE particles
from raw sludge to dewatered sludge, the existence of MPs in sludge is with the efficiency of 15%–40% [79,81]. The removal of MPs by clari­
acknowledged to pose a threat to the performance of anaerobic diges­ fication is downward movement, and it highly depends on the aggre­
tion. The presence of MPs can affect anaerobic digestion in terms of gation of MPs and flocs. If MPs are not adsorbed or trapped by flocs
methane production, induction of ROS, and hydrogen production under during coagulation-flocculation, MPs are more prone to escape from
alkaline anaerobic fermentation [102,103]. With the presence of PVC clarification. To note that, pristine MPs were used in laboratory exper­
particles in anaerobic digestion, the microbial populations in metha­ iment, while MPs found in DWPFs were mostly weathered. This differ­
nogens, hydrolytic microbes, and acidogens changed slightly, with ence might affect the interactions among MPs, coagulant, and
methane production eventually reduced to 75.8% in the presence of 60 flocculant, and thereby affecting the removal performance by
PVC items/g [102]. However, the impacts of the presence of MPs may clarification.
depend on the type of MP polymer. With the presence of PE particles at Apart from clarification, advanced DWPFs use dissolved air flotation
the same concentration levels as PVC MPs, a significant reduction in to replace traditional settlement of flocs from coagulation-flocculation.
methane production was found [104]. Leachate from MPs such as BPA Although Pivokonsky et al. found no significant difference in overall
and DBP also inhibited performance in anaerobic digestion. A total of removal efficiency between clarification and flotation, the overall MP
3.6 µg/L of BPA was released when the concentration of PVC particles removal was slightly higher (82.6%) in advanced DWPFs adopted
was 60 items/g [102]. Although no clear explanation is available on the flotation [82]. Unlike clarification, dissolved air flotation is an upward
effects of BPA leachate on methane production, a strong relationship has movement, when fine air bubbles facilitate the flotation of MPs. Even

13
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Fig. 3. Removal efficiency of MPs in different unit operations and processes in the selected DWPFs.

MPs are not adsorbed or trapped by flocs during coagulation- factor affecting the removal performance by GAC filtration. Non-polar
flocculation, MPs might be removed by dissolved air flotation. MPs MPs are more prone to be adsorbed by granular activated carbon. The
are more prone to float on the surface and skimmed during dissolved air adsorption of PE and PP particles favors the removal of pellets and
flotation, because of the density of common polymer types (0.83–1.58 g/ fragments by GAC filtration.
cm3). Dissolved air flotation is believed to result in higher MP removal
efficiency when compared with clarification. 3.6.2. Chemical treatment
Membrane filtration has been proposed to employ after coagulation- Coagulation-flocculation is the primary method of chemical treat­
flocculation, and lab-scale research has shown complete MP removal by ment in DWPFs. Coagulation is the process of destabilizing colloidal
ultrafiltration [79,81]. Similar to the membrane filtration in WWTPs, particles with coagulants, while flocculation is the process of colliding
the removal of MPs by ultrafiltration is the direct entrapment of MPs and aggregating destabilized particles into large flocs. Wang et al. found
without the limitation on shapes or polymer types. The only concern of that 40.5%–54.5% of MPs were removed by coagulation-flocculation
the removal performance is the sizes of MPs. If the sizes of MPs are larger combined with clarification in DWPF, whereas lab-scale research
than the pore size of membrane, this technique can achieve complete showed MP removal by coagulation-flocculation reached < 40%
removal of MPs. Membranes with a pore size of 30 nm retained almost [79–81]. Either alum-based coagulants or iron-based coagulants are
all PE particles, regardless the types of coagulants being dosed and the typically added during coagulation. A high dose of coagulant generally
availability of coagulation-flocculation [79,81]. Yet, the dynamics corresponds to a high removal efficiency in solids. The MP removal ef­
among cake layer formation, membrane fouling, the presence of MPs, ficiency of alum-based salts (<10%) was higher than that of iron-based
and the flocs were relatively complex, which affected the effectiveness salts (<5%) for each diameter of PE particles (500 µm to 5 mm) when a
of membrane filtration on MP removal [79,81]. The presence of small typical dose of coagulants in DWPFs (0.74 mM alum and 0.36 mM iron)
MPs and high dosage of coagulants induced frequent membrane fouling was added [81]. The effectiveness of coagulation-flocculation is also
and disrupted the uniformity of cake layer formed, so the specific based on the pH of water. Superior removal performance for PE particles
membrane flux reduced to 0.17 after operating for 300 s [79,81]. was found at a pH lower than 6.0, with alum-based salts forming smaller
Meanwhile, the presence of large MPs enhanced the porosity of the cake flocs with larger specific surface areas. The iron-based coagulant
layer and even caused breakage of cake layer, which mitigated mem­ removed PE particles more effectively (>15%) at a pH below 8.0 [79].
brane fouling [79,81]. Apart from the replacement of clarification after The removal rate improved with the use of anionic PAM and high doses
coagulation-flocculation, membrane techniques can also be employed of alum-based coagulants [79,81]. Increased settlement of PE particles
for filtration to replace traditional techniques. It is believed to have a was observed when an interaction occurred between anion PAM and
higher MP removal efficiency compared with typical sand filtration. The alum-based floc, resulting in the capture of small PE particles. A similar
pore size of membrane is typically 0.1 nm to 1.0 µm, and a small pore situation occurred when anion PAM was added to an iron-based coag­
size corresponds to a high removal efficiency of MPs. However, mem­ ulant, with the removal efficiency approaching 91% [79]. This implies
brane filtration is not a cost-effective method due to frequent membrane that the addition of PAM in coagulation-flocculation facilitates removal
fouling and the high operating cost associated with high pressures. of MPs, particularly small MPs. Since the removal performance of
In filtration, conventional DWPFs employ sand filtration to remove coagulation-flocculation is affected by the types of coagulant, pH of
additional particles before disinfection. DWPFs reportedly can achieve water, and the availability of PAM while the operating conditions of
complete removal on MP through the application of triple filtration DWPFs are fixed, the removal mechanisms might depend on the inter­
[83]. Low levels of MPs in raw water lead to low concentrations in action with MPs.
treated water, which do not reflect the removal efficiency of DWPFs on Regarding the impact of MP size on coagulation-flocculation, lab-
MPs [83]. The presence of MPs in treated water is likely due to the scale research has produced differing results. The highest removal effi­
abrasion of DWPF pipes and tanks [83]. Wang et al. found that sand ciency of small PE particles (500 µm) was only 8.3% and it gradually fell
filtration can remove 29.0%–44.4% of MPs, which was effective to to 2.7%, 1.7%, and 1.0% for MPs 0.5–1 mm, 1–2 mm, and 2–5 mm in
remove MPs > 10 µm [80]. Remarkably, it can remove 30.9%–49.3% size, respectively [81]. The surface charge of MPs is reportedly a sig­
and 23.5%–50.9% of fibers and pellets, respectively [80]. Yet, this nificant factor affecting removal efficiency [79,81]. Lab-scale research
technique is not a promising method for removing MPs from water [80]. revealed that MPs purchased from a manufacturer tend to be clear. MPs
The effectiveness of sand filtration depends on particle size, which the may be stable particles that rarely collide with coagulants, leading to a
typical effectiveness size is 0.45 to 0.70 mm in rapid sand filtration. The low removal efficiency. In addition, pure MP-flocs cannot settle because
potential removals of MPs by sand filtration are either entrapment be­ of their low density. However, MPs in DWPFs may adsorb a certain
tween the filter media or adherence to the surface of filter media [80]. amount of chemicals or heavy metals from the environment, producing a
Apart from sand filtration, advanced DWPFs use GAC filtration to slightly different surface charge than is associated with MPs in lab-scale
replace traditional filtration of impurities in the water. Wang et al. re­ research, and a greater potential to form flocs with coagulants.
ported that it can remove 56.8%–60.9% of MPs, especially removing
73.7%–98.5% of MPs < 5 µm [80]. Besides, it had a relatively good 3.6.3. Disinfection
performance on removing pellets (76.8%–86.3%), followed by frag­ Disinfection has no contribution to MP removal, instead it might
ments (60.3–69.1%) [80]. This might be corresponded to the high induce MP fragmentation. Wang et al. showed that ozonation had a
removal of PE and PP particles at the rate of 72.9–86.2% and negative removal of MP ranged from –9.3% to –0.3%, while the removal
59.4–66.8%, respectively [80]. GAC filtration depends on the polarity of efficiency of chlorination ranged from –0.7% to 6.8% [80]. Based on the
the solids through adsorption, which polymer type becomes the crucial changes in polymer type, PET and PE particles were increased by 10.2%

14
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

and 5.4% after ozonation, respectively, while only PP particles were 5. Conclusion
increased by 4.1% after chlorination [80]. Meanwhile, ozonation
induced the fragmentation on fibers and fragments, while chlorination WWTPs and DWPFs have two opposing roles: major removal
broke down pellet [80]. Wang et al. explained that MPs might encounter equipment and important sources of MPs. This paper comprehensively
great shearing force in ozonation, but it was more likely the destruction reviewed the occurrences and removal of MPs in WWTPs and DWPFs.
of residual organic matter by ozonation to remove biofilm-coated MPs Standards and protocols for MPs in WWTPs are lacking. Size, polymer
[80]. The impacts of disinfection on the fates of MPs might require type, sampling method, treatment process, and analysis differ by study,
further investigation. making comparisons difficult. MP abundance in terms of mass should be
considered to replace the number unit. Mass units destroy other in­
4. Recommendations for future research fluences by thermo-analytical methods, thereby offering more accurate
MP quantification than number units.
Physical treatment in WWTPs plays a major role in MP removal [58]. For WWTPs, the mass concentrations of MPs in influent ranged from
Despite a relatively high removal rate of MPs in WWTPs, a considerable 61 to 5,600 µg/L, and 0.5–170 µg/L of MPs were found in treated
amount of MPs is released to the environment through the discharge of effluent, representing an overall removal efficiency of 93.8%–99.8%.
treated effluent or sludge application. Our current understanding of the Since only two published papers reported the mass concentrations of
correlation between WWTPs and MP removal is limited to field in­ MPs, the comparison of MP abundance in this review was based on the
vestigations, and the removal mechanisms are not yet clear. While the number concentration. The number concentration of MPs in influent and
presence of small MPs in treated effluent provides evidence of MP treated effluents ranges from 1.01 to 31,400 items/L and 0.004 to 447
fragmentation during treatment, the actual fate of MPs within unit op­ items/L, respectively. Fragments/fibers are the dominant shapes, and
erations and processes has not been examined in the field. The removal PE/PA/PS is the major polymer type of MPs. The relatively low con­
mechanisms of each unit operation and process (ASP, sedimentation, centration of MPs in effluent remains a concern due to large daily dis­
grit removal, and MBR) are not supported by laboratory studies. Several charges. MPs were removed during preliminary, primary, secondary,
important factors, including the characteristics of populations (e.g., in­ and tertiary treatments at efficiencies of 6.0%–58.6%, 19.1%–99.0%,
come level, domestic activities, and living standards), characteristics of –66.7%–92.6%, and –72.7%–99.9%, respectively. Preliminary and pri­
influent (e.g., types of wastewater and combined stormwater), and the mary treatments make a higher removal contribution than do secondary
operating conditions of unit operations and processes, have not been and tertiary treatments. Girt/grease removal exhibits the highest
consider. These factors may limit the effectiveness of management removal efficiency through mechanical and physical potential, but it
strategies to eliminate discharges of MPs from WWTPs. might facilitate the initial MP fragmentation. The effectiveness of ter­
The diverse methodologies involved in designing sample collection tiary treatment is highly dependent on the applied unit process. Disin­
procedures, MP extraction, and MP analysis further undermine the fection has the greatest potential to induce MP fragmentation or
usefulness of comparisons among studies (Tables S1-S3). Future chemical degradation of MPs. The concentration of MPs in dewatered
research should consider several strategies. First, standardized protocols sludge was much higher than in raw sludge, at 9.7–186.7 items/g (dw).
for sampling and MP extraction should be established. Sampling Mechanical dewatering may contribute to MP fragmentation, resulting
methods, consideration of the diurnal and temporal nature of waste­ in a significant increase in MP concentration in dewatered sludge.
water, and the representativeness of volume are major parameters that During sludge treatment, no unit processes are available to remove MPs
determine MP occurrence rates in WWTPs [90]. Sampling details, from raw sludge. The presence of MPs in sludge can affect anaerobic
including location, weather, unit operations and processes, flow rate, digestion in terms of methane production, induction of ROS, and
and types of influents, should be disclosed. Due to the high organic hydrogen production. Although the latest studies consider the perfor­
content of wastewater and sludge, digestion is an unavoidable step in mance of MP removal in each unit operation and process, in-depth
MP extraction. Researchers have adopted various methods, including discussions of the responsible removal mechanisms are rare. Instead of
oxidation and chemical degradation, to model digestion, but no com­ MP removal, the fate of MPs is primarily the function of the separation
parisons have been made among the methods. A vast amount of infor­ from liquid to solid in the treatment plant. Future research should
mation is required to incorporate all parameters and statistical data consider the possibly of mineralization and transition of MPs within
relevant to WWTPs and MPs. Reliable and rapid analytical techniques WWTPs.
should be developed to eliminate the subjective judgment inherent in For DWPFs, the concentration of MPs in raw water and treated water
visual observation of samples [105]. FTIR and Raman spectroscopy are of DWPFs ranges from 1 to 6,614 items/L and 1 to 930 items/L,
commonly used for the chemical characterization of MPs [54]. Mass respectively. Fragment shapes are dominant, and PET is a major polymer
concentration through FTIR and thermoanalytical techniques are a more type. The removal efficiency of MPs varies from 69.9% to 100%
robust measures that are less affected by size [106,107]. These methods regardless of the types of treatment. A relatively high overall removal
are effective for chemical analysis, but novel validation methods should rate of MPs is available in the literature, but the reported variance in MP
be developed for rapid, reliable, and cost-effective analysis. concentration in treated water indicates a human health risk.
Both our understanding of removal mechanisms within WWTPs and Coagulation-flocculation may cut MP abundance in water by half, but
effective methods for MP removal are limited in the field [70]. Instead of the removal performance depends on the operating conditions (e.g., pH,
being removed, some MPs are separated or fragmented into smaller coagulant, and the addition of PAM) and surface charge of MPs. Mem­
pieces in WWTPs. Future research should consider a paradigm shift from brane filtration is the preferred process to remove MPs from water,
MP removal to separation, due to the inherent nature of biological and followed by GAC filtration and dissolved air flotation. Disinfection
chemical resistance of MPs. Ideally, MPs could be decomposed into the might contribute MP fragmentation by chemical degradation.
carbon cycle and completely mineralized by biodegradation, photo­
degradation, thermo-oxidation, and hydrolysis. The possible fates of Declaration of Competing Interest
MPs within WWTPs should include complete mineralization of MPs into
water and carbon dioxide and the transition of MPs to wastewater and The authors declare that they have no known competing financial
sludge. To conclude, future research is needed with respect to the sep­ interests or personal relationships that could have appeared to influence
aration of MPs in WWTP. the work reported in this paper.

15
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

Acknowledgments [19] A. Isobe, K. Kubo, Y. Tamura, S. Kako, E. Nakashima, N. Fujii, Selective transport
of microplastics and mesoplastics by drifting in coastal waters, Mar. Pollut. Bull.
89 (1-2) (2014) 324–330, https://doi.org/10.1016/j.marpolbul.2014.09.041.
This work was supported by the Research Grants Council of the Hong [20] H. Zhang, Transport of microplastics in coastal seas, Estuar. Coast. Shelf S 199
Kong SAR, China (No. 18202116), the Departmental Collaborative Fund (2017) 74–86, https://doi.org/10.1016/j.ecss.2017.09.032.
(No. 04488), and Internal Research Grant (No. RG38/2019-2020R, SFG/ [21] K. Zhang, W. Gong, J.Z. Lv, X. Xiong, C.X. Wu, Accumulation of floating
microplastics behind the Three Gorges Dam, Environ. Pollut. 204 (2015)
TFG) of The Education University of Hong Kong, Basic Science Research 117–123, https://doi.org/10.1016/j.envpol.2015.04.023.
Program through the National Research Foundation of Korea (NRF) [22] H. Lee, W.J. Shim, J.-H. Kwon, Sorption capacity of plastic debris for hydrophobic
funded by the Ministry of Education (No. 2019R1A4A1027795), and the organic chemicals, Sci. Total Environ. 470 (2014) 1545–1552, https://doi.org/
10.1016/j.scitotenv.2013.08.023.
Korea Ministry of Environment (MOE) as Waste to Energy-Recycling [23] J. Li, K. Zhang, H. Zhang, Adsorption of antibiotics on microplastics, Environ.
Human Resource Development Project (YL-WE-18-002). Pollut. 237 (2018) 460–467, https://doi.org/10.1016/j.envpol.2018.02.050.
[24] X. Guo, J. Pang, S. Chen, H. Jia, Sorption properties of tylosin on four different
microplastics, Chemosphere 209 (2018) 240–245, https://doi.org/10.1016/j.
Appendix A. Supplementary data chemosphere.2018.06.100.
[25] W. Wang, J. Wang, Comparative evaluation of sorption kinetics and isotherms of
Supplementary data to this article can be found online at https://doi. pyrene onto microplastics, Chemosphere 193 (2018) 567–573, https://doi.org/
10.1016/j.chemosphere.2017.11.078.
org/10.1016/j.cej.2020.128381. [26] M. Jang, W.J. Shim, G.M. Han, M. Rani, Y.K. Song, S.H. Hong, Widespread
detection of a brominated flame retardant, hexabromocyclododecane, in
References expanded polystyrene marine debris and microplastics from South Korea and the
Asia-Pacific coastal region, Environ. Pollut. 231 (2017) 785–794, https://doi.org/
10.1016/j.envpol.2017.08.066.
[1] M.A. Browne, P. Crump, S.J. Niven, E. Teuten, A. Tonkin, T. Galloway, T.
[27] I. Velzeboer, C.J.A.F. Kwadijk, A.A. Koelmans, Strong sorption of PCBs to
C. Richard, Microplastic moves pollutants and additives to worms, reducing
nanoplastics, microplastics, carbon nanotubes, and fullerenes, Environ. Sci.
functions linked to health and biodiversity, Curr. Biol. 23 (2013) 2388–2392,
Technol. 48 (9) (2014) 4869–4876, https://doi.org/10.1021/es405721v.
https://doi.org/10.1016/j.cub.2013.10.012.
[28] P. Wu, Z. Cai, H. Jin, Y. Tang, Adsorption mechanisms of five bisphenol analogues
[2] A. McCormick, T.J. Hoellein, S.A. Mason, J. Schluep, J.J. Kelly, Microplastic is an
on PVC microplastics, Sci. Total Environ. 650 (2019) 671–678, https://doi.org/
abundant and distinct microbial habitat in an urban river, Environ. Sci. Technol.
10.1016/j.scitotenv.2018.09.049.
48 (20) (2014) 11863–11871, https://doi.org/10.1021/es503610r.
[29] D. Brennecke, B. Duarte, F. Paiva, I. Caçador, J. Canning-Clode, Microplastics as
[3] R.W. Obbard, S. Sadri, Y.Q. Wong, A.A. Khitun, I. Baker, R.C. Thompson, Global
vector for heavy metal contamination from the marine environment, Estuar.
warming releases microplastic legacy frozen in Arctic Sea ice, Earths Future 2 (6)
Coast. Shelf S 178 (2016) 189–195, https://doi.org/10.1016/j.ecss.2015.12.003.
(2014) 315–320, https://doi.org/10.1002/eft2.2014.2.issue-610.1002/
[30] M.E. Hodson, C.A. Duffus-Hodson, A. Clark, M.T. Prendergast-Miller, K.L. Thorpe,
2014EF000240.
Plastic bag derived-microplastics as a vector for metal exposure in terrestrial
[4] L. Van Cauwenberghe, A. Vanreusel, J. Mees, C.R. Janssen, Microplastic pollution
invertebrates, Environ. Sci. Technol. 51 (8) (2017) 4714–4721, https://doi.org/
in deep-sea sediments, Environ. Pollut. 182 (2013) 495–499, https://doi.org/
10.1021/acs.est.7b0063510.1021/acs.est.7b00635.s001.
10.1016/j.envpol.2013.08.013.
[31] A. Bakir, S.J. Rowland, R.C. Thompson, Enhanced desorption of persistent
[5] M.A. Browne, P. Crump, S.J. Niven, E. Teuten, A. Tonkin, T. Galloway,
organic pollutants from microplastics under simulated physiological conditions,
R. Thompson, Accumulation of microplastic on shorelines woldwide: Sources and
Environ. Pollut. 185 (2014) 16–23, https://doi.org/10.1016/j.
sinks, Environ. Sci. Technol. 45 (21) (2011) 9175–9179, https://doi.org/
envpol.2013.10.007.
10.1021/es201811s.
[32] A.A. Koelmans, B. Meulman, T. Meijer, M.T.O. Jonker, Attenuation of
[6] V. Hidalgo-Ruz, L. Gutow, R.C. Thompson, M. Thiel, Microplastics in the marine
polychlorinated biphenyl sorption to charcoal by humic acids, Environ. Sci.
environment: A review of the methods used for identification and quantification,
Technol. 43 (3) (2009) 736–742, https://doi.org/10.1021/es802862b.
Environ. Sci. Technol. 46 (6) (2012) 3060–3075, https://doi.org/10.1021/
[33] L. Liu, R. Fokkink, A.A. Koelmans, Sorption of polycyclic aromatic hydrocarbons
es2031505.
to polystyrene nanoplastic, Environ. Toxicol. Chem. 35 (7) (2016) 1650–1655,
[7] R.R. Hurley, A.L. Lusher, M. Olsen, L. Nizzetto, Validation of a method for
https://doi.org/10.1002/etc.3311.
extracting microplastics from complex, organic-rich, environmental matrices,
[34] J. Gago, F. Galgani, T. Maes, R.C. Thompson, Microplastics in seawater:
Environ. Sci. Technol. 52 (13) (2018) 7409–7417, https://doi.org/10.1021/acs.
Recommendations from the marine strategy framework directive implementation
est.8b0151710.1021/acs.est.8b01517.s001.
process, Front. Mar. Sci. 3 (2016) 219, https://doi.org/10.3389/
[8] Z. Long, Z. Pan, W. Wang, J. Ren, X. Yu, L. Lin, H. Lin, H. Chen, X. Jin,
fmars.2016.00219.
Microplastic abundance, characteristics, and removal in wastewater treatment
[35] C.F. Araujo, M.M. Nolasco, A.M.P. Ribeiro, P.J.A. Ribeiro-Claro, Identification of
plants in a coastal city of China, Water Res. 155 (2019) 255–265, https://doi.org/
microplastics using Raman spectroscopy: Latest developments and future
10.1016/j.watres.2019.02.028.
prospects, Water Res. 142 (2018) 426–440, https://doi.org/10.1016/j.
[9] S. Estahbanati, N.L. Fahrenfeld, Influence of wastewater treatment plant
watres.2018.05.060.
discharges on microplastic concentrations in surface water, Chemosphere 162
[36] M. Simon, N. van Alst, J. Vollertsen, Quantification of microplastic mass and
(2016) 277–284, https://doi.org/10.1016/j.chemosphere.2016.07.083.
removal rates at wastewater treatment plants applying Focal Plane Array (FPA)-
[10] H.S. Auta, C.U. Emenike, S.H. Fauziah, Distribution and importance of
based Fourier Transform Infrared (FT-IR) imaging, Water Res. 142 (2018) 1–9,
microplastics in the marine environment: A review of the sources, fate, effects,
https://doi.org/10.1016/j.watres.2018.05.019.
and potential solutions, Environ. Int. 102 (2017) 165–176, https://doi.org/
[37] G. Gatidou, O.S. Arvaniti, A.S. Stasinakis, Review on the occurrence and fate of
10.1016/j.envint.2017.02.013.
microplastics in sewage treatment plants, J. Hazard. Mater. 367 (2019) 504–512,
[11] F.S. Cesa, A. Turra, J. Baruque-Ramos, Synthetic fibers as microplastics in the
https://doi.org/10.1016/j.jhazmat.2018.12.081.
marine environment: A review from textile perspective with a focus on domestic
[38] H.-J. Kang, H.-J. Park, O.-K. Kwon, W.-S. Lee, D.-H. Jeong, B.-K. Ju, J.-H. Kwon,
washings, Sci. Total Environ. 598 (2017) 1116–1129, https://doi.org/10.1016/j.
Occurrence of microplastics in municipal sewage treatment plants: A review,
scitotenv.2017.04.172.
Environ. Health Toxicol. 33 (3) (2018) e2018013, https://doi.org/10.5620/eht.
[12] K. Duis, A. Coors, Microplastics in the aquatic and terrestrial environment:
e2018013.
Sources (with a specific focus on personal care products), fate and effects,
[39] J. Sun, X. Dai, Q. Wang, M.C.M. van Loosdrecht, B.-J. Ni, Microplastics in
Environ. Sci. Eur. 28 (2016) 2, https://doi.org/10.1186/s12302-015-0069-y.
wastewater treatment plants: Detection, occurrence and removal, Water Res. 152
[13] S. Ziajahromi, P.A. Neale, F.D.L. Leusch, Wastewater treatment plant effluent as a
(2019) 21–37, https://doi.org/10.1016/j.watres.2018.12.050.
source of microplastics: Review of the fate, chemical interactions and potential
[40] P.L. Ngo, B.K. Pramanik, K. Shah, R. Roychand, Pathway, classification and
risks to aquatic organisms, Water Sci. Technol. 74 (2016) 2253–2269, https://
removal efficiency of microplastics in wastewater treatment plants, Environ.
doi.org/10.2166/wst.2016.414.
Pollut. 255 (2019) 113326, https://doi.org/10.1016/j.envpol.2019.113326.
[14] B. Gewert, M.M. Plassmann, M. MacLeod, Pathways for degradation of plastic
[41] S. Raju, M. Carbery, A. Kuttykattil, K. Senathirajah, S.R. Subashchandrabose,
polymers floating in the marine environment, Environ. Sci.-Proc. Imp. 17 (9)
G. Evans, P. Thavamani, Transport and fate of microplastics in wastewater
(2015) 1513–1521, https://doi.org/10.1039/C5EM00207A.
treatment plants: Implications to environmental health, Rev. Environ. Sci.
[15] J. Boucher, D. Friot, Primary microplastics in the oceans: A global evaluation of
Biotechnol. 17 (4) (2018) 637–653, https://doi.org/10.1007/s11157-018-9480-
sources, Gland, Switzerland: IUCN (2017), https://doi.org/10.2305/IUCN.
3.
CH.2017.01.en.
[42] M. Enfrin, L.F. Dumee, J. Lee, Nano/microplastics in water and wastewater
[16] Y.K. Song, S.H. Hong, M.i. Jang, G.M. Han, S.W. Jung, W.J. Shim, Combined
treatment processes - Origin, impact and potential solutions, Water Res. 161
effects of UV exposure duration and mechanical abrasion on microplastic
(2019) 621–638, https://doi.org/10.1016/j.watres.2019.06.049.
fragmentation by polymer type, Environ. Sci. Technol. 51 (8) (2017) 4368–4376,
[43] D. Eerkes-Medrano, H.A. Leslie, B. Quinn, Microplastics in drinking water: A
https://doi.org/10.1021/acs.est.6b0615510.1021/acs.est.6b06155.s001.
review and assessment, Curr. Opin. Environ. Sustain. 7 (2019) 69–75, https://doi.
[17] A.L. Andrady, Microplastics in the marine environment, Mar. Pollut. Bull. 62 (8)
org/10.1016/j.coesh.2018.12.001.
(2011) 1596–1605, https://doi.org/10.1016/j.marpolbul.2011.05.030.
[44] A.A. Koelmans, N.H.M. Nor, E. Hermsen, M. Kooi, S.M. Mintenig, J. De France,
[18] R. Dris, J. Gasperi, M. Saad, C. Mirande, B. Tassin, Synthetic fibers in atmospheric
Microplastics in freshwaters and drinking water: Critical review and assessment
fallout: A source of microplastics in the environment? Mar. Pollut. Bull. 104 (1-2)
(2016) 290–293, https://doi.org/10.1016/j.marpolbul.2016.01.006.

16
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

of data quality, Water Res. 155 (2019) 410–422, https://doi.org/10.1016/j. treatment technologies, Water Res. 123 (2017) 401–407, https://doi.org/
watres.2019.02.054. 10.1016/j.watres.2017.07.005.
[45] K. Novotna, L. Cermakova, L. Pivokonska, T. Cajthaml, M. Pivokonsky, [67] M.R. Michielssen, E.R. Michielssen, J. Ni, M.B. Duhaime, Fate of microplastics
Microplastics in drinking water treatment - Current knowledge and research and other small anthropogenic litter (SAL) in wastewater treatment plants
needs, Sci. Total Environ. 667 (2019) 730–740, https://doi.org/10.1016/j. depends on unit processes employed, Environ. Sci. Water Res. Technol. 2 (6)
scitotenv.2019.02.431. (2016) 1064–1073, https://doi.org/10.1039/C6EW00207B.
[46] S. Gündoğdu, C. Çevik, E. Güzel, S. Kilercioğlu, Microplastics in municipal [68] Y. Ruan, K. Zhang, C. Wu, R. Wu, P.K.S. Lam, A preliminary screening of HBCD
wastewater treatment plants in Turkey: A comparison of the influent and enantiomers transported by microplastics in wastewater treatment plants, Sci.
secondary effluent concentrations, Environ. Monit. Assess. 190 (2018) 626, Total Environ. 674 (2019) 171–178, https://doi.org/10.1016/j.
https://doi.org/10.1007/s10661-018-7010-y. scitotenv.2019.04.007.
[47] L. Lin, L.-Z. Zuo, J.-P. Peng, L.-Q. Cai, L. Fok, Y. Yan, H.-X. Li, X.-R. Xu, [69] S. Ziajahromi, P.A. Neale, L. Rintoul, F.D.L. Leusch, Wastewater treatment plants
Occurrence and distribution of microplastics in an urban river: A case study in the as a pathway for microplastics: Development of a new approach to sample
Pearl River along Guangzhou City, China, Sci. Total Environ. 644 (2018) wastewater-based microplastics, Water Res. 112 (2017) 93–99, https://doi.org/
375–381, https://doi.org/10.1016/j.scitotenv.2018.06.327. 10.1016/j.watres.2017.01.042.
[48] P. Eisentraut, E. Dümichen, A.S. Ruhl, M. Jekel, M. Albrecht, M. Gehde, U. Braun, [70] S.A. Carr, J. Liu, A.G. Tesoro, Transport and fate of microplastic particles in
Two birds with one stone-Fast and simultaneous analysis of microplastics: wastewater treatment plants, Water Res. 91 (2016) 174–182, https://doi.org/
Microparticles derived from thermoplastics and tire wear, Environ. Sci. Technol. 10.1016/j.watres.2016.01.002.
Let. 5 (10) (2018) 608–613, https://doi.org/10.1021/acs. [71] S.M. Mintenig, I. Int-Veen, M.G.J. Löder, S. Primpke, G. Gerdts, Identification of
estlett.8b0044610.1021/acs.estlett.8b00446.s001. microplastic in effluents of waste water treatment plants using focal plane array-
[49] I.E. Napper, R.C. Thompson, Release of synthetic microplastic plastic fibres from based micro-Fourier-transform infrared imaging, Water Res. 108 (2017)
domestic washing machines: Effects of fabric type and washing conditions, Mar. 365–372, https://doi.org/10.1016/j.watres.2016.11.015.
Pollut. Bull. 112 (1-2) (2016) 39–45, https://doi.org/10.1016/j. [72] X. Xu, Y. Jian, Y. Xue, Q. Hou, L. Wang, Microplastics in the wastewater treatment
marpolbul.2016.09.025. plants (WWTPs): Occurrence and removal, Chemosphere 235 (2019) 1089–1096,
[50] S. Wagner, T. Huffer, P. Klockner, M. Wehrhahn, T. Hofmann, T. Reemtsma, Tire https://doi.org/10.1016/j.chemosphere.2019.06.197.
wear particles in the aquatic environment - A review on generation, analysis, [73] S.A. Mason, D. Garneau, R. Sutton, Y. Chu, K. Ehmann, J. Barnes, P. Fink,
occurrence, fate and effects, Water Res. 139 (2018) 83–100, https://doi.org/ D. Papazissimos, D.L. Rogers, Microplastic pollution is widely detected in US
10.1016/j.watres.2018.03.051. municipal wastewater treatment plant effluent, Environ. Pollut. 218 (2016)
[51] J. Bayo, S. Olmos, J. López-Castellanos, Microplastics in an urban wastewater 1045–1054, https://doi.org/10.1016/j.envpol.2016.08.056.
treatment plant: The influence of physicochemical parameters and environmental [74] H. Hidayaturrahman, T.-G. Lee, A study on characteristics of microplastic in
factors, Chemosphere 238 (2020) 124593, https://doi.org/10.1016/j. wastewater of South Korea: Identification, quantification, and fate of
chemosphere.2019.124593. microplastics during treatment process, Mar. Pollut. Bull. 146 (2019) 696–702,
[52] R. Dris, J. Gasperi, V. Rocher, M. Saad, N. Renault, B. Tassin, Microplastic https://doi.org/10.1016/j.marpolbul.2019.06.071.
contamination in an urban area: A case study in Greater Paris, Environ. Chem. 12 [75] M. Fischer, B.M. Scholz-Böttcher, Simultaneous trace identification and
(2015) 592–599, https://doi.org/10.1071/EN14167. quantification of common types of microplastics in environmental samples by
[53] E.A. Gies, J.L. LeNoble, M. Noël, A. Etemadifar, F. Bishay, E.R. Hall, P.S. Ross, pyrolysis-gas chromatography–mass spectrometry, Environ. Sci. Technol. 51 (9)
Retention of microplastics in a major secondary wastewater treatment plant in (2017) 5052–5060, https://doi.org/10.1021/acs.est.6b0636210.1021/acs.
Vancouver, Canada, Mar. Pollut. Bull. 133 (2018) 553–561, https://doi.org/ est.6b06362.s001.
10.1016/j.marpolbul.2018.06.006. [76] F. Corradini, P. Meza, R. Eguiluz, F. Casado, E. Huerta-Lwanga, V. Geissen,
[54] M. Lares, M.C. Ncibi, M. Sillanpää, M. Sillanpää, Occurrence, identification and Evidence of microplastic accumulation in agricultural soils from sewage sludge
removal of microplastic particles and fibers in conventional activated sludge disposal, Sci. Total Environ. 671 (2019) 411–420, https://doi.org/10.1016/j.
process and advanced MBR technology, Water Res. 133 (2018) 236–246, https:// scitotenv.2019.03.368.
doi.org/10.1016/j.watres.2018.01.049. [77] X. Li, L. Chen, Q. Mei, B. Dong, X. Dai, G. Ding, E.Y. Zeng, Microplastics in sewage
[55] X. Liu, W. Yuan, M. Di, Z. Li, J. Wang, Transfer and fate of microplastics during sludge from the wastewater treatment plants in China, Water Res. 142 (2018)
the conventional activated sludge process in one wastewater treatment plant of 75–85, https://doi.org/10.1016/j.watres.2018.05.034.
China, 365, Chem. Eng. J. 362 (2019) 176–182, https://doi.org/10.1016/j. [78] A.M. Mahon, B. O’Connell, M.G. Healy, I. O’Connor, R. Officer, R. Nash,
cej.2019.01.033. L. Morrison, Microplastics in sewage sludge: Effects of treatment, Environ. Sci.
[56] X. Lv, Q. Dong, Z. Zuo, Y. Liu, X. Huang, W.-M. Wu, Microplastics in a municipal Technol. 51 (2) (2017) 810–818, https://doi.org/10.1021/acs.
wastewater treatment plant: Fate, dynamic distribution, removal efficiencies, and est.6b0404810.1021/acs.est.6b04048.s001.
control strategies, J. Cleaner Prod. 225 (2019) 579–586, https://doi.org/ [79] B. Ma, W. Xue, Y. Ding, C. Hu, H. Liu, J. Qu, Removal characteristics of
10.1016/j.jclepro.2019.03.321. microplastics by Fe-based coagulants during drinking water treatment,
[57] K. Magnusson, F. Norén, Screening of microplastic particles in and down-stream a J. Environ. Sci. 78 (2019) 267–275, https://doi.org/10.1016/j.jes.2018.10.006.
wastewater treatment plant, IVL Swedish Environ. Res. Inst. C 55 (2014). https: [80] Z. Wang, T. Lin, W. Chen, Occurrence and removal of microplastics in an
//www.diva-portal.org/smash/get/diva2:773505/FULLTEXT01.pdf. advanced drinking water treatment plant (ADWTP), Sci. Total Environ. 700
[58] F. Murphy, C. Ewins, F. Carbonnier, B. Quinn, Wastewater treatment works (2020) 134520, https://doi.org/10.1016/j.scitotenv.2019.134520.
(WwTW) as a source of microplastics in the aquatic environment, Environ. Sci. [81] B. Ma, W. Xue, C. Hu, H. Liu, J. Qu, L. Li, Characteristics of microplastic removal
Technol. 50 (11) (2016) 5800–5808, https://doi.org/10.1021/acs. via coagulation and ultrafiltration during drinking water treatment, Chem. Eng. J.
est.5b0541610.1021/acs.est.5b05416.s001. 359 (2019) 159–167, https://doi.org/10.1016/j.cej.2018.11.155.
[59] K. Conley, A. Clum, J. Deepe, H. Lane, B. Beckingham, Wastewater treatment [82] M. Pivokonsky, L. Cermakova, K. Novotna, P. Peer, T. Cajthaml, V. Janda,
plants as a source of microplastics to an urban estuary: Removal efficiencies and Occurrence of microplastics in raw and treated drinking water, Sci. Total Environ.
loading per capita over one year, Water Res. X 3 (2019) 100030, https://doi.org/ 643 (2018) 1644–1651, https://doi.org/10.1016/j.scitotenv.2018.08.102.
10.1016/j.wroa.2019.100030. [83] S.M. Mintenig, M.G.J. Löder, S. Primpke, G. Gerdts, Low numbers of microplastics
[60] H. Lee, Y. Kim, Treatment characteristics of microplastics at biological sewage detected in drinking water from ground water sources, Sci. Total Environ. 648
treatment facilities in Korea, Mar. Pollut. Bull. 137 (2018) 1–8, https://doi.org/ (2019) 631–635, https://doi.org/10.1016/j.scitotenv.2018.08.178.
10.1016/j.marpolbul.2018.09.050. [84] A.A. Horton, A. Walton, D.J. Spurgeon, E. Lahive, C. Svendsen, Microplastics in
[61] H.A. Leslie, S.H. Brandsma, M.J.M. Van Velzen, A.D. Vethaak, Microplastics en freshwater and terrestrial environments: Evaluating the current understanding to
route: Field measurements in the Dutch river delta and Amsterdam canals, identify the knowledge gaps and future research priorities, Sci. Total Environ. 586
wastewater treatment plants, North Sea sediments and biota, Environ. Int. 101 (2017) 127–141, https://doi.org/10.1016/j.scitotenv.2017.01.190.
(2017) 133–142, https://doi.org/10.1016/j.envint.2017.01.018. [85] M. Filella, Questions of size and numbers in environmental research on
[62] R.M. Blair, S. Waldron, C. Gauchotte-Lindsay, Average daily flow of microplastics microplastics: Methodological and conceptual aspects, Environ. Chem. 12 (2015)
through a tertiary wastewater treatment plant over a ten-month period, Water 527–538, https://doi.org/10.1071/EN15012.
Res. 163 (2019) 114909, https://doi.org/10.1016/j.watres.2019.114909. [86] Z. Pan, X.W. Sun, H.G. Guo, S.Z. Cai, H.Z. Chen, S.M. Wang, Y.B. Zhang, H. Lin,
[63] S. Magni, A. Binelli, L. Pittura, C.G. Avio, C.T. Torre, C.P. Parenti, S. Gorbi, J. Huang, Prevalence of microplastic pollution in the Northwestern Pacific Ocean,
F. Regoli, The fate of microplastics in an Italian wastewater treatment plant, Sci. Chemosphere 225 (2019) 735–744, https://doi.org/10.1016/j.
Total Environ. 652 (2019) 602–610, https://doi.org/10.1016/j. chemosphere.2019.03.076.
scitotenv.2018.10.269. [87] C.G. Liu, J. Li, Y.L. Zhang, L. Wang, J. Deng, Y. Gao, L. Yu, J.J. Zhang, H.W. Sun,
[64] J. Talvitie, M. Heinonen, J.-P. Pääkkönen, E. Vahtera, A. Mikola, O. Setälä, Widespread distribution of PET and PC microplastics in dust in urban China and
R. Vahala, Do wastewater treatment plants act as a potential point source of their estimated human exposure, Environ. Int. 128 (2019) 116–124, https://doi.
microplastics? Preliminary study in the coastal Gulf of Finland, Baltic Sea, Water org/10.1016/j.envint.2019.04.024.
Sci. Technol. 72 (2015) 1495–1504, https://doi.org/10.2166/wst.2015.360. [88] M.X. Di, J. Wang, Microplastics in surface waters and sediments of the Three
[65] J. Talvitie, A. Mikola, O. Setälä, M. Heinonen, A. Koistinen, How well is Gorges Reservoir, China, Sci. Total Environ. 616 (2018) 1620–1627, https://doi.
microlitter purified from wastewater? – A detailed study on the stepwise removal org/10.1016/j.scitotenv.2017.10.150.
of microlitter in a tertiary level wastewater treatment plant, Water Res. 109 [89] P.A. Helm, Improving microplastics source apportionment: A role for microplastic
(2017) 164–172, https://doi.org/10.1016/j.watres.2016.11.046. morphology and taxonomy? Anal. Methods 9 (9) (2017) 1328–1331, https://doi.
[66] J. Talvitie, A. Mikola, A. Koistinen, O. Setälä, Solutions to microplastic pollution - org/10.1039/C7AY90016C.
Removal of microplastics from wastewater effluent with advanced wastewater

17
Y.L. Cheng et al. Chemical Engineering Journal 410 (2021) 128381

[90] J. Li, H. Liu, J.P. Chen, Microplastics in freshwater systems: A review on [100] R.A. Naik, L.S. Rowles, A.I. Hossain, M. Yen, R.M. Aldossary, O.G. Apul, J. Conkle,
occurrence, environmental effects, and methods for microplastics detection, N.B. Saleh, Microplastic particle versus fiber generation during photo-
Water Res. 137 (2018) 362–374, https://doi.org/10.1016/j.watres.2017.12.056. transformation in simulated seawater, Sci. Total Environ. 736 (2020) 139690,
[91] J. Huang, Y.-C. Jin, Numerical modeling of Type I circular sedimentation tank, https://doi.org/10.1016/j.scitotenv.2020.139690.
J. Environ. Eng. 137 (3) (2011) 196–204, https://doi.org/10.1061/(ASCE) [101] R. Mao, M. Lang, X. Yu, R. Wu, X. Yang, X. Guo, Aging mechanism of
EE.1943-7870.0000310. microplastics with UV irradiation and its effects on the adsorption of heavy
[92] G.R. Pophali, S.N. Kaul, T. Nandy, S. Devotta, Development of a novel circular metals, J. Hazard. Mater. 393 (2020) 122515, https://doi.org/10.1016/j.
secondary clarifier for improving solids liquid separation in wastewater jhazmat.2020.122515.
treatment, Water Environ. Res. 81 (2) (2009) 140–149, https://doi.org/10.2175/ [102] W. Wei, Q.-S. Huang, J. Sun, J.-Y. Wang, S.-L. Wu, B.-J. Ni, Polyvinyl chloride
106143008X390816. microplastics affect methane production from the anaerobic digestion of waste
[93] M. Lapointe, J. Farner, L.M. Hernandez, N. Tufenkji, Understanding and activated sludge through leaching toxic bisphenol-A, Environ. Sci. Technol. 53 (5)
improving microplastics removal during water treatment: Impact of coagulation (2019) 2509–2517, https://doi.org/10.1021/acs.est.8b0706910.1021/acs.
and flocculation, Environ. Sci. Technol. 54 (2020) 8719–8727, https://doi.org/ est.8b07069.s001.
10.1021/acs.est.0c00712. [103] W. Wei, Y.-T. Zhang, Q.-S. Huang, B.-J. Ni, Polyethylene terephthalate
[94] K. Rajala, O. Grönfors, M. Hesampour, A. Mikola, Removal of microplastics from microplastics affect hydrogen production from alkaline anaerobic fermentation of
secondary wastewater treatment plant effluent by coagulation/flocculation with waste activated sludge through altering viability and activity of anaerobic
iron, aluminum and polyamine-based chemicals, Water Res. 183 (2020) 116045, microorganisms, Water Res. 163 (2019) 114881, https://doi.org/10.1016/j.
https://doi.org/10.1016/j.watres.2020.116045. watres.2019.114881.
[95] L. Lu, T. Luo, Y. Zhao, C. Cai, Z. Fu, Y. Jin, Interaction between microplastics and [104] W. Wei, Q.-S. Huang, J. Sun, X. Dai, B.-J. Ni, Revealing the mechanisms of
microorganism as well as gut microbiota: A consideration on environmental polyethylene microplastics affecting anaerobic digestion of waste activated
animal and human health, Sci. Total Environ. 667 (2019) 94–100, https://doi. sludge, Environ. Sci. Technol. 53 (16) (2019) 9604–9613, https://doi.org/
org/10.1016/j.scitotenv.2019.02.380. 10.1021/acs.est.9b0297110.1021/acs.est.9b02971.s001.
[96] D. Nasuhoglu, S. Isazadeh, P. Westlund, S. Neamatallah, V. Yargeau, Chemical, [105] W.J. Shim, S.H. Hong, S.E. Eo, Identification methods in microplastic analysis: A
microbial and toxicological assessment of wastewater treatment plant effluents review, Anal. Methods 9 (9) (2017) 1384–1391, https://doi.org/10.1039/
during disinfection by ozonation, Chem. Eng. J. 346 (2018) 466–476, https://doi. C6AY02558G.
org/10.1016/j.cej.2018.04.037. [106] E. Dümichen, A.-K. Barthel, U. Braun, C.G. Bannick, K. Brand, M. Jekel, R. Senz,
[97] Z. Zhang, Y. Chen, Effects of microplastics on wastewater and sewage sludge Analysis of polyethylene microplastics in environmental samples, using a thermal
treatment and their removal: A review, Chem. Eng. J. 382 (2020) 122955, decomposition method, Water Res. 85 (2015) 451–457, https://doi.org/10.1016/
https://doi.org/10.1016/j.cej.2019.122955. j.watres.2015.09.002.
[98] X. Zhang, J. Chen, J.i. Li, The removal of microplastics in the wastewater [107] O. Mallow, S. Spacek, T. Schwarzböck, J. Fellner, H. Rechberger, A new
treatment process and their potential impact on anaerobic digestion due to thermoanalytical method for the quantification of microplastics in industrial
pollutants association, Chemosphere 251 (2020) 126360, https://doi.org/ wastewater, Environ. Pollut. 259 (2020) 113862, https://doi.org/10.1016/j.
10.1016/j.chemosphere.2020.126360. envpol.2019.113862.
[99] V.P. Kelkar, C.B. Rolsky, A. Pant, M.D. Green, S. Tongay, R.U. Halden, Chemical
and physical changes of microplastics during sterilization by chlorination, Water
Res. 163 (2019) 114871, https://doi.org/10.1016/j.watres.2019.114871.

18

You might also like