Next Article in Journal
Does Digital Inclusive Finance Help County Level Governance in the Five Provinces of Northwest China, from the Perspective of Economic Resilience?
Previous Article in Journal
Integrated Adaptation Strategies for Human–Leopard Cat Coexistence Management in Taiwan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microplastics in Water: A Review of Characterization and Removal Methods

1
College of Chemistry and Chemical Engineering, Central South University, Changsha 410083, China
2
College of Environmental Science and Engineering, Hunan University and Key Laboratory of Environmental Biology and Pollution Control (Hunan University), Ministry of Education, Changsha 410082, China
3
Hunan Haili Chemical Industry Co., Ltd., Changsha 410007, China
4
Hunan Provincial Key Laboratory of the Research and Development of Novel Pharmaceutical Preparations, the “Double-First Class” Application Characteristic Discipline of Hunan Province (Pharmaceutical Science), Changsha Medical University, Changsha 410219, China
*
Authors to whom correspondence should be addressed.
Sustainability 2024, 16(10), 4033; https://doi.org/10.3390/su16104033
Submission received: 29 March 2024 / Revised: 27 April 2024 / Accepted: 1 May 2024 / Published: 11 May 2024

Abstract

:
Microplastics (MPs), as an emerging persistent pollutant, exist and accumulate in the environment, which has garnered them considerable global attention. While the origin, dispersion, distribution, and impact of MPs have been extensively documented, the characterization and removal strategies for MPs present ongoing challenges. In this literature review, we introduce in detail the advantages and disadvantages of seven characterization methods, from macroscopic to microscopic, from visual observation to microscopic characterization, and discuss their scope of application. In addition, 12 treatment schemes were summarized from the three treatment directions of physics, chemistry, and biology, including filtration, adsorption, extraction, magnetic separation, oil film separation, Fenton oxidation, electrochemical oxidation, persulfate advanced oxidation, photocatalytic oxidation, coagulation, electrocoagulation, foam flotation, anaerobic–anoxic–aerobic activated sludge, enzymatic degradation, bacterial degradation, and fungal degradation. Additionally, we present a critical assessment of the advantages and drawbacks associated with these removal strategies. Building upon the findings of our research team, we propose a novel approach to degrade MPs, which combines three-dimensional electrocatalytic oxidation technology with persulfate advanced oxidation technology. This advanced oxidation technology achieves 100% degradation of antibiotics in water, can degrade large molecules into environmentally harmless small molecules, and should also be a very good strategy for the degradation of MPs. Compared with two-dimensional electrocatalytic technology, the degradation efficiency is higher and the degradation cost is lower This review intends to propel further advancements for addressing the issue of MP pollution.

1. Introduction

1.1. Definition

In the past 70 years, the global production of plastic has exceeded 8 billion tons, and this number is still increasing rapidly. According to statistics, by 2050, the global annual production of primary plastic may exceed 34 billion tons. Plastics undergo a gradual decomposition process, resulting in the formation of small plastic fragments, through physical, chemical, and biological mechanisms. These fragments can be categorized into various sizes, such as megaplastics (more than 50 cm), macroplastics (5–50 cm), mesoplastics (0.5–5 cm), microplastics (MP, 0.1 μm–5 cm), and nanoplastics (NP, less than 1 μm) [1]. As far back as 2004, Thompson et al. from Plymouth University made a groundbreaking discovery of plastic debris in marine water and sediment, coining the term “microplastics” (MPs) [2]. MPs are defined as plastic fragments and particles with a diameter less than 5 mm [3,4,5]. In 2013, the Marine Waste Technical Group of the European Marine Waste Management Agency proposed a classification of microplastics into small microplastics (SMPs, 1 μm–1 mm) and large microplastics (LMPs, 1–5 mm) [6,7]. The particle size of MPs can vary from a few microns to a few millimeters, forming a heterogeneous mixture of plastic particles with different shapes, making them nearly imperceptible to the naked eye [8,9]. They are often described as the “PM2.5 of the sea”, drawing a vivid analogy to airborne fine particulate matter.

1.2. Sources

Gaining insight into the origins of MPs was a crucial aspect of comprehending their formation. Plastic waste, encompassing packaging materials (PE, PP, PS, PVC, PET, PC, PEP, PMMA), medical devices (PVC, PE, PP, PS, PTFE), and agricultural films (PE, PVC, EVA), emerged as a significant contributor to the presence of MPs [10]. The degradation of plastic waste, whether through natural processes or human intervention, yielded fragments teeming with MPs, resembling detritus or scum [11,12,13]. Cleaning and washing products (EVOH, PU, PS, PP, PC) in daily life were also one of the sources of MPs [14,15,16]. Notably, Dris et al. discovered that clothing released a substantial amount of fibers (PTT, PET, PBT, PA, PAN, PVA, PU, PP, PVC, etc.) during the drying process, with natural drying methods exacerbating fiber shedding compared to machine drying [17]. In addition, the wear and erosion of textiles and microfibers from tire usage constituted another significant wellspring of MPs [18,19].

1.3. Discharge, Distribution, and Transportation

The migration of MPs occurs through surface runoff and atmospheric deposition, representing significant pathways for their dispersal. All kinds of plastics existing in soil and water will degrade into microplastics of different sizes through photolysis, weathering, hydrolysis, and microbial degradation. Microplastics circulate and transfer in water, soil, and the atmosphere through surface runoff, underground infiltration, atmospheric deposition, and atmospheric flow (Figure 1). Regrettably, sewage treatment plants within urban centers cannot handle these pollutants, leading to their ultimate accumulation within the vast expanse of the ocean [20,21]. Due to the characteristics of microplastics, such as being thin, small, and easily adsorbed, the introduction, migration, and transportation of microplastics in water are the fastest. Microplastics in land waters or oceans are suspended in the atmosphere by air currents and then moved around. Urban activities and the utilization of everyday household items play a pivotal role in the widespread distribution and transportation of MPs on land. From the wear and tear of tires to the flaking of paint, from plastic manufacturing to the textile industry, and from cosmetics to exfoliating agents and washing products, these seemingly innocuous elements all contribute to the proliferation of MPs [22,23].
Of particular significance are the superfine fibers emanating from our everyday apparel and bedding, which are released into the air and subsequently emerge as the primary source of atmospheric MPs. Under the influence of rainwater, these minute microfibers settle upon the land, infiltrating rivers and ultimately finding their way into the depths of the ocean [24]. The Philippines, India, Malaysia, China, and Indonesia are the top five countries responsible for annual marine plastic emissions. Rainfall, acting as a potent driving force, facilitates the transportation of MPs, while also serving as a conduit for their migration from the surface to the waterways. Astonishingly, the research team led by Leslie et al. made a groundbreaking discovery, detecting microplastics within the blood of human volunteers for the first time, further underscoring the omnipresence of these pollutants in our daily lives [25].
Plants and animals, through a variety of pathways, absorb MPs from the atmosphere, soil, and water, perpetuating the continuous transfer and enrichment of these contaminants throughout the intricate web of the food chain. A momentous revelation emerged from the research conducted by Aves et al.’s team, as they detected microplastics within newly fallen snow in Antarctica, shedding light on the far-reaching extent of MPs’ influence [26]. Thus, it becomes evident that MPs have permeated every facet of our existence.

1.4. Negative Effects

MPs in the environment affect biome composition and nitrogen cycling. Numerous studies conducted both domestically and internationally have revealed the propensity of MPs to permeate the very fabric of our ecosystem, infiltrating crops, fish, earthworms, chickens, bees, marine, and terrestrial animals, as well as humans, thereby impeding their growth, development, and reproductive capabilities [27,28]. MPs can adsorb other pollutants, and these vagabond MPs themselves and their adsorbed pollutions are easily consumed by mussels, zooplankton, and other organisms residing at the lower echelons of the food chain, and possess an inherent resistance to digestion within the confines of the stomach. Consequently, they accumulate and settle within the delicate tissues and bodily structures of humans, unleashing an onslaught of maladies and even death in their wake [29,30,31].

2. Characterization of MPs

The classification of MPs encompasses two distinct categories: primary MPs and secondary MPs [32,33], as illustrated in Figure 2. Primary MPs pertain to plastic granules derived from industrial products that find their way into the aquatic environment via rivers and sewage treatment plants. Examples include the minute particles encapsulated within cosmetic formulations, as well as plastic and resin granules utilized as raw materials in industrial processes. On the other hand, secondary microplastics arise from the physical, chemical, and biological transformations undergone by larger plastic waste, resulting in fragmentation and volume reduction [34,35,36].
Microplastics (MPs) have gained recognition as a burgeoning category of pollutants worldwide, warranting diligent attention from researchers in terms of detection and characterization [37,38]. Characterizing MPs poses several challenges, primarily due to the complex composition of environmental samples, which often comprise diverse substances that are difficult to differentiate from MPs. Consequently, elucidating the morphological attributes of MPs stands as a significant obstacle within the realm of MP research [39,40]. Presently, prevailing methods of characterization encompass visual discrimination, microscopic discrimination, scanning electron microscopy (SEM), atomic force microscopy (AFM), Fourier transform infrared spectroscopy (FT-IR), Raman spectroscopy, and pyrolysis analysis [41,42,43].

2.1. Visual Discrimination

The method of visual identification primarily targets larger microplastics, ranging from 1 to 5 mm in size, which are predominantly found in coastal areas. This approach typically involves the use of tweezers and trays to directly separate and identify the particles [44]. However, the presence of numerous organic and inorganic substances in the general sample, which closely resemble the size and appearance of plastics, makes it challenging to differentiate them accurately. Occasionally, smaller but colorful plastics can be identified visually [45]. It is worth noting that naturally occurring fibers tend to produce white and transparent microplastics, while synthetic fibers often exhibit vibrant colors [46]. Although the visual method is straightforward and convenient, it can lead to significant errors due to the similarity between plastics and other substances in the sample. De Witte et al. devised a technique that involved touching the sample with a heated needle tip, allowing them to determine if the material was plastic based on whether it melted or curled. However, this method has its limitations, as the properties of certain plastics may not change unless the temperature of the needle tip is sufficiently high. Additionally, this approach is most effective when specific characteristics of the plastic are known beforehand [47]. While the visual method is easy to implement, it is prone to substantial errors.

2.2. Microscopic Discrimination

Traditional optical microscopy has been widely employed for the identification of microplastics measuring several hundred microns in size [48]. By enlarging the image, this method offers valuable insights into the surface texture and structural characteristics of the particles, enabling the differentiation of plastics from other materials with similar appearances. While this technique is capable of detecting smaller microplastics, accurate discrimination becomes challenging when dealing with colorless and shapeless particles measuring less than 100 μm [49]. It is worth noting that previous studies have uncovered a significant discrepancy of up to 20% in the classification of plastics using microscopy, with transparent particles accounting for 70% of the misidentified samples, a finding later verified through spectral analysis [50].

2.3. Scanning Electron Microscopy

Scanning electron microscopy (SEM) employs a powerful electron beam to illuminate the sample, initiating interactions that generate secondary electrons. These electrons serve as valuable signals, revealing the intricate morphology of the specimen [51]. SEM yields high-resolution, magnified images of plastic particles, enabling the discrimination of minute microplastics and organic particles [52]. In a comprehensive study, Cooper et al. utilized SEM to meticulously examine the morphological characteristics of various plastic fragments collected from beach environments. Their meticulous observations confirmed that both mechanical and chemical weathering processes occurring on the shoreline induce the development of cracks, grooves, and notches on the surfaces of plastic fragments. Over time, these alterations contribute to the fragmentation of plastics into smaller particles [53]. Nevertheless, this technique is not without its challenges. For instance, samples must be examined in a vacuum environment, restricting the range of applicable specimens. Moreover, SEM provides only two-dimensional plane images, lacking height and directional information. Consequently, liquid samples cannot be observed using this method [54].

2.4. Atomic Force Microscope

Currently, the examination of microplastics (MPs) through the use of atomic force microscopy (AFM) allows for the detection of minute particles, reaching a minimum size of several microns, while providing a genuine three-dimensional representation of the surface topography [55]. Moreover, this method circumvents the need for special sample preparation and operates effectively under normal pressure, even in liquid environments. Demir-Yilmaz et al. employed atomic force microscopy to investigate the biophysical attributes of MPs, revealing their nanostructure characterized by rough, irregular, and hydrophobic surfaces [56]. By integrating AFM with microfluidics [57], it becomes possible to precisely assess the interaction between microalgae and MPs, while accurately determining their hydrophobic properties. Nevertheless, AFM presents limitations, including a restricted imaging range, slow imaging speed, and susceptibility to probe interference [58].

2.5. Fourier Transform Infrared Spectroscopy

Fourier transform infrared spectroscopy (FT-IR) boasts numerous advantages, including its non-invasive nature, uncomplicated sample preparation, and qualitative precision, rendering it a favored technique for the structural analysis of materials [59,60]. Employing graphical analysis, FT-IR can circumvent false-positive outcomes in the absence of microplastics (MPs) and minimize misidentification of MPs lacking distinct coloration or material characteristics [61]. By utilizing infrared radiation to detect molecular vibration frequencies and distinct functional groups, FT-IR provides insight into the weathering degree of MPs through oxygen-demanding bonds. Nonetheless, this method is susceptible to interference from water and organic pollutants, posing challenges in detecting oxidation functional groups of MPs and MPs with diameters below 20 microns. These factors, whether directly or indirectly, impact the qualitative efficacy of FT-IR in identifying MPs [62]. To enhance the recognition accuracy of MPs, Wander et al. ingeniously combined Principal Component Analysis (PCA), a statistical feature extraction method based on minimizing the mean square error, with FT-IR, effectively reducing the data dimensions of MP samples and visually depicting particle familiarity [63,64]. In addition to identifying sample composition, FT-IR also permits quantitative analysis of MP quantities [65].

2.6. Raman Spectrometer

Raman spectrometry is an exquisite technique of vibrational spectroscopy, which relies on the captivating phenomenon of inelastic light scattering to bestow upon us mesmerizing vibrational spectra. This method has found wide application in the meticulous analysis of minute particles [66,67]. Not only does Raman analysis have the capability to unravel the identity of these particles, but it also holds the power to divulge precious insights into the composition of the samples under scrutiny. Astonishingly, the Raman spectrometer boasts remarkable sensitivity, capable of detecting particles as minuscule as 1 μm in size [68]. What further elevates the allure of this technique is its non-contact nature, which ensures the preservation of the pristine structure of the samples, and, in turn, facilitates subsequent analyses [69]. In contrast to its counterpart, FT-IR, Raman spectroscopy triumphs in the identification of both organic and inorganic additives and coatings, in addition to the matrix polymers. However, it is worth noting that the Raman signal from the matrix can easily be obfuscated by the more pronounced scattering from additives and coatings, rendering their identification a formidable task. Meanwhile, the presence of fluorescence in the samples poses the most formidable challenge to Raman detection. Infrared spectroscopy, on the other hand, proves to be a more suitable tool for the identification of additives and coatings, particularly when confronted with samples that exhibit fluorescence [70]. Nevertheless, it is essential to acknowledge that the efficacy of this characterization method is contingent upon the stringent demands imposed upon the samples.

2.7. Thermal Cleavage

Thermal cleavage analysis, an emerging spectroscopic technique that exploits the thermal stability of samples to discern changes in their physicochemical properties, has emerged as a promising tool for the identification of microplastics (MPs) [71]. The combination of thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) has proven effective in discerning polyethylene (PE) and polypropylene (PP) [72]. By integrating TGA with solid-phase extraction (SPE) and coupling it with thermal and inspiratory chromatography–mass spectrometry (TDS-GC-MS), one can harness the full detection potential of multiple approaches, revealing intricate details with exceptional resolution [73]. This methodology offers the advantage of direct identification of samples and mixed polymers, representing a relatively straightforward and expeditious means of analysis. However, it is crucial to note that this approach is inherently destructive, limiting its utility in chemical characterization alone, thereby precluding the acquisition of crucial information regarding the morphology, size, and quantification of microplastics [74]. Thus, ongoing efforts to refine and optimize thermal analysis are imperative to establish it as an efficient and widely adopted technology for microplastic detection.

3. MP Treatment Technologies

Currently, a plethora of approaches exist for addressing microplastics (MPs), categorically classified as physical, chemical, and biological treatment techniques (Figure 3). The physical methods encompass various strategies such as filtration [75], adsorption [76], extraction [77], magnetic separation [78], and oil film separation [79]. On the other hand, the chemical methodologies encompass Fenton oxidation, electrocoagulation, advanced oxidation technologies, coagulation, and foam flotation. Lastly, the biological approaches encompass the anaerobic–anoxic–aerobic activated sludge method, enzymatic degradation, bacterial degradation, and fungal degradation [80,81,82,83,84,85].

3.1. Physical Methods

3.1.1. Filtration Method

The process of filtration involves the interception of solid particles and other substances within a suspension, thereby separating them from the liquid medium. In the early stages of microplastic (MP) separation, the filtration method emerged as a widely employed technique due to its simplicity and rapidity of operation [86,87]. For instance, Tadsuwan et al. utilized a series of filters ranging from 5 mm to 0.05 mm in size to eliminate MPs from wastewater obtained from Thai municipal treatment plants, achieving a removal rate of 33.33% [88]. Wang et al. employed a biochar filter to effectively eliminate MPs with a diameter of 10 μm, resulting in a remarkably efficient removal exceeding 95% [89]. Similarly, Ziajahromi et al. employed gravel as a filtering medium to extract MPs from sludge and biosolids obtained from Australian wastewater treatment plants, resulting in a removal rate of 69–79% [90]. While the filtration method remains a common strategy for managing MPs, it is worth noting that it can generate smaller MPs, thereby augmenting the challenges associated with post-processing. This method only collects microplastics and does not achieve the purpose of eradication. Moreover, this method has higher requirements for the filter, and too high a cost will be invested if the processing volume is too large. In future studies, researchers should design more advanced filters to increase throughput and reduce treatment costs.

3.1.2. Adsorption Method

The adsorption technique entails utilizing a porous solid adsorbent to capture one or more types of adsorbates from the water sample’s surface. Subsequently, appropriate solvents, or heating or blowing methods, are employed to release the adsorbate, thereby achieving the objective of separation and enrichment. Adsorption is frequently employed for the elimination of water pollutants [91,92,93]. Currently, adsorption is commonly employed for the removal of microplastics (MPs) from water [94,95]. For example, Wang et al. fabricated a natural biodegradable sponge material with exceptional mechanical properties by utilizing plant protein as a chemical crosslinking agent. A polystyrene MP waste solution was prepared using deionized water. The adsorbent’s removal efficiency reached 38% with an adsorption time of 10 s. In simulated wastewater, the material exhibited an adsorption efficiency of 81.2% for MPs. Even after 20 cycles, the adsorbent maintained its rapid adsorption capability. The primary mechanisms of MP removal by the adsorption material are hydrophobicity and particle diffusion [96]. Sun et al. employed chitin and graphene oxide (ChGO) as raw materials to fabricate a solid and compressible sponge adsorption material. Deionized water was used to prepare the MP waste solution. The adsorption rates of ChGO for pure polystyrene, carboxylic-acid-modified polystyrene, and amine-modified polystyrene were 89.8%, 72.4%, and 88.9%, respectively. This adsorption material primarily captures MPs through electrostatic interactions, hydrogen bond interactions, and π-π interactions. Additionally, the material exhibits excellent activation and regeneration capabilities [97]. Yuan et al. employed three-dimensional reduced graphene oxide as an adsorbent to remove polystyrene MPs. It was discovered that the main adsorption mechanism was the strong π-π interaction between graphene oxide and polystyrene MPs. Deionized water was utilized to prepare the MPs waste solution. The maximum adsorption capacity of three-dimensional reduced graphene oxide for polystyrene MPs was 617.28 mg/g. Furthermore, the material displayed remarkable regeneration capabilities [98]. The adsorption method offers the advantages of simple operation, minimal equipment requirements, and high efficiency. However, the cost, structural stability, and adsorption selectivity of the adsorbent restrict its widespread application. Moreover, the adsorption method solely separates MPs from water, necessitating the implementation of additional methods for the treatment of MPs in subsequent stages. In future studies, researchers need to design an adsorbent that is efficient in handling microplastics, has good recycling performance, and is easy to recycle.

3.1.3. Extraction Method

The utilization of extraction techniques has gained considerable attention for the treatment of industrial wastewater containing high concentrations of phenols, nitrogen heterocycles, dyes, heavy metals, and other pollutants [99,100]. In recent years, numerous scholars worldwide have devoted their efforts to the development and application of extraction methods for the treatment of microplastics (MPs) [101,102]. For instance, Li et al. employed a custom-designed separation and extraction apparatus to effectively isolate and extract MP particles, presenting a readily available device for MP extraction. This self-fashioned equipment successfully extracted three types of biodegradable MPs (polybutylene succinate, poly(adipic acid) butylene terephthalate, and polylactic acid) as well as four types of non-degradable MPs (low-density polyethylene, polystyrene, polypropylene, and polyvinyl chloride). The recovery rates for these MPs ranged between 92% and 99.6%, thus highlighting the accuracy and precision of their separation and extraction device [103]. In a separate study, Nuelle et al. implemented a two-step approach to extract MPs from sediment samples. The extraction recoveries for polyethylene, polypropylene, polyvinyl chloride, poly(ethylene terephthalate), polystyrene, and polyurethane (1 mm) were found to be as high as 91% to 99% [104]. Similarly, Han et al. employed an extraction method to isolate MP particles from soil and sediment samples. By refining the flotation process and flotation solution, they successfully extracted and separated six commonly found MP compounds, namely polyethylene, polyethylene terephthalate, polypropylene, polyvinyl chloride, polystyrene, and expanded polystyrene. The extraction recoveries achieved for MPs were remarkably high, ranging from 80% to 100% [105]. Wang et al. applied an extraction technique to separate styrene MP spheres of various sizes (0.05, 1.0, 2.6, 4.8, and 100 μm) from biosolid and soil samples. While styrene nanoparticles with a diameter of 100 μm could be extracted completely from biosolids and soil, the extraction efficiency for smaller particles ranged from 5% to 80% [106]. This method boasts the advantages of simple instrumentation, automated control, and high operational safety. However, it is worth noting that the cost associated with this method remains high, and the separation of the dissolved solute in the extraction solvent poses a challenge. In future studies, researchers should seek a green and economical extractant to treat MPs.

3.1.4. Magnetic Separation

Magnetic separation technology encompasses the application of magnetic fields to manipulate substances, often harnessed in the realm of water treatment [107]. Presently, magnetic separation finds frequent employment in the isolation of microplastics (MPs). As an illustration, Tang et al. devised hydrophobic iron nanoparticles for the magnetic separation and extraction of MPs. Remarkably, their findings reveal that this material effectively eliminates over 90% of MPs ranging from 10 to 20 μm and exceeding 1 mm in size from seawater. Furthermore, this material boasts an ability to remove 84% and 78% of MPs measuring 200 μm to 1 mm from freshwater and sediment, respectively [108]. In a separate study, Tang et al. synthesized magnetic carbon nanotubes, leveraging their magnetic properties to isolate MPs from aqueous solutions. When employing a dose of 5 g/L, the efficient removal for MPs reached a remarkable 100% within a span of 300 min. Even after undergoing four cycles of use, the efficiency remained at 80% for MPs at a concentration of 5 g/L. Scanning electron microscopy (SEM) images exhibited the adsorption of MPs onto the surface of magnetic carbon nanotubes [109]. The magnetic separation approach boasts notable advantages, such as its capacity for high-volume treatment, minimal generation of waste sludge, and the potential for long-range magnetic enhancement in separation. However, certain drawbacks persist, including the tendency for magnetic seeds, MPs, and other lipophilic/oleophobic substances to aggregate on surfaces. In future studies, researchers need to develop a suitable magnetic separation method for different MPs to improve the universality of this method as much as possible.

3.1.5. Oil Film Separation

Oil film separation represents a hydrophobic and density-agnostic technique, frequently employed in the realm of microplastic (MP) separation [110]. For instance, Crichton et al. introduced an innovative and economical oil film method for MP removal. Impressively, the efficient removal of total MPs, fibers, and particles reached 96.1% ± 7.4, 92.7% ± 4.3, and 99% ± 1.4, respectively, underscoring the immense potential of this approach in MP elimination [111]. In a similar vein, Mani et al. harnessed castor oil membranes to effectuate the separation of MPs from aqueous matrices, yielding an impressive average MP removal rate of up to 99%. Notably, this method achieved an MP removal efficiency of 74 ± 13% in the Rhine River, further solidifying its environmental friendliness, harmlessness, and efficacy in MP separation [112]. The method’s virtues include its density-agnostic nature, cost-effectiveness, and low risk. However, it is worth mentioning that the separation funnel was prone to blockage during the separation process [113]. In future studies, researchers should work to solve the problem of separation equipment blockage and establish a set of treatment methods for different sizes of microplastics.

3.2. Chemical Methods

3.2.1. Fenton Oxidation and Advanced Oxidation Technology

Fenton oxidation technology finds extensive application in the realm of water treatment [114,115]. Presently, researchers have reported the utilization of chemical oxidation methods to address wastewater containing microplastics (MPs) [116]. For instance, Liu et al. employed the heat-activated K2S2O8 oxidation process and the Fenton oxidation process in the treatment of MPs. K2S2O8 generates a considerable quantity of sulfate radicals and hydroxyl radicals under Fenton-like and thermal activation, which facilitate the oxidation and degradation of MPs. Consequently, the scanning electron microscopy (SEM) diagram in Figure 4 demonstrates the deformation of the surface of polystyrene (PS) and polyethylene (PE), signifying a certain extent of degradation of MPs composed of these materials [117]. Prominent advanced oxidation technologies presently encompass electrochemical oxidation, persulfate advanced oxidation, and photocatalytic oxidation. Photocatalytic degradation of MPs also represents a prevalent approach [118,119]. For instance, Venkataramana et al. employed a 350 W metal halide lamp to irradiate polyethylene MPs, resulting in a weight reduction rate of 12.5% after 5 h, indicating the partial degradation of MPs through the photocatalytic method [120]. Uheida et al. proposed a sustainable and environmentally friendly photocatalytic technique for eliminating polypropylene MPs from water activated by visible light. Following two weeks of visible-light irradiation, the average particle volume decreased by 65% due to reduction. Gas chromatography–mass spectroscopy (GC/MS) analysis revealed that the predominant degradation by-products were ethynyloxy/acetyl radicals, hydroxypropyl, butyraldehyde, acetone, acrolein (propenal), and the pentyl group [121]. Additionally, electrochemical oxidation also represents a common method for wastewater treatment [122,123]. Kiendrebeogo et al. employed an electrochemical oxidation process to address the issue of synthetic polystyrene microplastics (MPs) in wastewater. Through the application of an electric field, a significant number of hydroxyl radicals and sulfate radicals were generated, showcasing their potent oxidizing capabilities against polystyrene MPs. Ultimately, the mineralization of polystyrene MPs into CO2 was achieved. The removal efficiency reached an impressive 89 ± 8% within a span of 6 h, utilizing a Na2SO4 dosage of 0.06 M. SEM characterization instruments substantiated that the degradation of polystyrene MPs did not result in their fragmentation into smaller particles, but rather their direct conversion into gaseous products (Figure 5) [124].
Miao et al. adopted a two-dimensional electrocatalytic oxidation method to degrade polyvinyl chloride (PVC) microplastics. Following 6 h of electrocatalytic oxidation, remarkable dechlorination efficiency (75%) and weight loss (56%) were achieved for PVC. Mechanistic insights, as depicted in Figure 6, illustrated that polyvinyl chloride obtains electrons directly from the TiO2/C cathode plate and undergoes dechlorination under elevated temperature conditions. Concurrently, hydroxyl radicals oxidize polyvinyl chloride microplastics, resulting in the formation of oxygen-containing groups such as C=O and O-H. Eventually, these substances are partially mineralized into CO2 and H2O. Gas chromatography–mass spectroscopy (GC/MS) and high-performance liquid chromatography (HPLC) analyses revealed that the predominant degradation by-products were alkenes, alcohols, monocarboxylic acids, dicarboxylic acids, and esters [125].
Fenton oxidation and advanced oxidation technology offer a multitude of advantages, including high removal efficiency, dependable and consistent outcomes, simple equipment, convenient operation and maintenance, and cost-effectiveness. However, it is important to acknowledge that these technologies also present a range of challenges, such as suboptimal treatment efficacy, elevated costs, the potential for secondary pollution, and stringent process requirements.
Based on our team’s comprehensive investigation into electrochemical oxidation and persulfate advanced oxidation, we have devised a sophisticated water treatment system that combines three-dimensional electrocatalytic oxidation with persulfate advanced oxidation. Our findings indicate that the combination of electrocatalytic oxidation, PMS advanced oxidation, and waste coal cinder yielded the highest efficient removal (99.95%) and mineralization efficiency (90.16%) for sulfadiazine over a period of 90 min [126]. The three-dimensional electrocatalytic coupled PMS advanced oxidation system achieved an efficient removal and efficient mineralization of 99.56% and 88.63%, respectively, for sulfamethazine within the same time frame [127]. Furthermore, we employed a self-assembled three-dimensional electrocatalytic oxidation degradation reactor to degrade sulfonamide, resulting in an efficient removal and efficient mineralization of 99.845% and 88.958%, respectively [128]. Similarly, sulfonamide exhibited an efficient removal and efficient mineralization of 99.867% and 89.675%, respectively [129]. Additionally, a meticulously prepared spherical bimetallic clay catalyst was utilized to degrade bromobenzonitrile, achieving a complete efficient removal of 100% [130]. These degradation systems generate a substantial amount of highly reactive oxygen species, which possess potent oxidation capabilities. Moreover, these degradation systems demonstrate remarkable recyclability. Furthermore, the presence of chloride ions enhances the degradation efficiency of these systems.
Based on the aforementioned reports, it was postulated that the integration of three-dimensional electrocatalytic oxidation with persulfate advanced oxidation would impart a fresh perspective on the degradation of microplastics (MPs). This treatment method is relatively new and its mechanism may not be clear. In future studies, researchers should broaden the study of the treatment mechanisms, so as to avoid the environmental impact of some degradation intermediates.

3.2.2. Coagulation Method

The coagulation technique involves the introduction of coagulants into wastewater, thereby destabilizing and aggregating the organic pollutants present into larger clumps of alum, measuring hundreds of microns or even millimeters. Subsequently, the pollutants in the wastewater could be eliminated through gravity sedimentation or other methods of solid–liquid separation [131]. Currently, coagulation is employed in the treatment of pollutants containing microplastics (MPs) [132,133]. For instance, Zhou et al. utilized poly-aluminum chloride (PAC) and FeCl3 as coagulants to eliminate polystyrene (PS) and polyethylene (PE) MPs. As depicted in Figure 7, charge neutralization transpires between the flocculant and MPs. Scanning electron microscope (SEM) images reveal the occurrence of aggregation and adsorption between the MPs and the coagulant, while Fourier transform infrared (FTIR) spectra demonstrate the formation of novel chemical bonds during the interaction between the MPs and the coagulant. Furthermore, there is a discernible alteration in the zeta potential before and after adsorption, indicating the successful removal of PS and PE MPs [134]. Ma et al. employed aluminum-based (AlCl3·6H2O) and iron-based salt (FeCl3·6H2O) coagulants to eliminate polyethylene MPs. The findings indicate that aluminum salt outperforms iron salt in the removal of polyethylene. Notably, the efficiency of removal increases as the particle size of polyethylene decreases. Nevertheless, even when the dosage of aluminum-based salt was as high as 15 Mm, the maximum average removal rate reached a mere 36.89% [135]. Shahi et al. employed alum coagulant and alum composite cationic polyamine-coated sand coagulant to eliminate MPs from wastewater in drinking water treatment plants. The results demonstrate that the removal rate of the alum composite cationic polyamine-coated sand coagulant surpassed that of alum alone by 26.8%. The data underscore the significance of the particle size, morphology, and surface characteristics of MPs in the removal process within drinking water treatment plants [136].
Moreover, electrocoagulation serves as a technique employed for the eradication of pollutants through the application of a pulsed high voltage, thereby facilitating electrochemical reactions. Presently, electrocoagulation has found utility in the elimination of MP pollutants. For instance, Perren et al. employed electrocoagulation in the purification of synthetic wastewater containing varying concentrations of polyethylene MP spheres. Their findings corroborate that the efficacy of electrocoagulation in removing pollutants can surpass 90% when the pH level ranges from 3 to 10. Astonishingly, the removal rate achieved a staggering 99.24% at a pH value of 7.5 [137].
In summary, the coagulation process exhibits numerous advantages, including its simplicity in operation, minimal equipment requirements, and swift treatment duration. Nevertheless, it is important to note that pH exerts a profound influence on this method, and many coagulants possess reducibility and coloration. Should the dosage be excessive, adverse consequences such as heightened chromaticity and diminished removal rates may ensue. In future studies, researchers should increase their exploration of the factors that affect this method and work out the most appropriate treatment scheme.

3.2.3. Foam Flotation Method

Furthermore, foam flotation represents a technique utilized for the separation of minerals from impurities. Raw ore powder is agitated with water and reagents, which selectively interact with the desired minerals to modify their surface properties. Subsequently, air is introduced into the mixture, causing the targeted minerals to rise to the top and form a froth. Currently, numerous researchers have employed this method for the removal of MPs [138,139]. For instance, Imhof et al. devised a foam-flotation-based approach for the separation of MPs, achieving a removal efficiency of 55% [140]. Additionally, Nguyen et al. have highlighted the presence of unpredictable factors that may impede the separation of MPs [141]. Talvitie et al. developed an air flotation technique to eliminate MPs from water, achieving an impressive removal efficiency of up to 95%, thereby reducing the MP concentration from 2 MP/L to 0.1 MP/L in an aqueous solution [142]. Enfrin et al. and Sun et al. advocate the use of foam flotation as a means to treat MPs, given its simple operation, low cost, and potential to mitigate MP discharge into sewage [143,144]. Additionally, Jiang et al. applied froth flotation for the removal of MPs from beach and lake sediments (Figure 8). In their study, sodium oleate was employed to restore the hydrophobicity of MPs, facilitating their effective removal from sediments [145]. Consequently, the foam flotation method offers advantages in its uncomplicated equipment and affordability. However, the experiment’s reproducibility remains a significant challenge, with temperature exerting a substantial influence. Regrettably, few scientists have expounded upon the unstable performance of foam flotation, leaving it as a focal point for future research endeavors.

3.3. Biological Methods

3.3.1. Anaerobic–Anoxic–Aerobic Activated Sludge Method

The anaerobic–anoxic–aerobic activated sludge (AAO) process was employed to eliminate organic pollutants in water through a combination of anaerobic, anoxic, and aerobic zones, along with different sludge return strategies. This technique primarily targeted the removal of BOD [146]. Presently, numerous researchers have applied this approach to address MPs [147,148]. For instance, Yang et al. introduced a technology based on the anaerobic–anoxic–aerobic activated sludge process to treat authentic MPs obtained from a sewage treatment facility in Beijing, resulting in a 54.47% removal rate [149]. Jia et al. proposed a similar approach for the treatment of real MPs in wastewater obtained from a wastewater treatment plant in Shanghai, achieving a removal rate of 26.01% [150]. Jiang et al. implemented an anaerobic–anoxic–aerobic activated sludge process to treat real MPs in wastewater derived from wastewater treatment plants in northern China, resulting in a removal rate of 16.9% [151]. Similarly, Liu et al. applied an anaerobic–anoxic–aerobic activated sludge process to treat authentic MPs in wastewater from a sewage treatment plant in a specific region of China, yielding a removal rate of 16.6% [152]. The AAO process offers advantages such as cost-effectiveness, a straightforward process flow, and a short hydraulic retention time. However, this method is time-consuming, exhibits low removal efficiency, is susceptible to bacterial demise, and generates a substantial amount of sludge. In future studies, researchers need to screen and domesticate a high-quality bacterial community that can achieve good results in the degradation of microplastics in different environments.

3.3.2. Enzymatic Degradation

Currently, the field of in situ degradation of microplastics (MPs) by enzymes under gentle conditions is a highly active area of research [153,154]. In this approach, either indigenous or introduced microorganisms are employed to degrade or metabolize MPs, transforming them into harmless end products. Biocatalysis itself embodies the principles of environmental friendliness, and an ideal enzyme possesses the ability to accomplish MP degradation with exceptional efficiency [155]. In 2016, Yoshida et al. pioneered the development and utilization of an enzyme capable of effectively breaking down polyethylene terephthalate plastics. However, the inherent instability of this enzyme hindered its practical application in the field of biodegradation [156]. Subsequently, Son et al. utilized a thermally stable variant of PETase to degrade polyethylene terephthalate MPs. Nevertheless, this enzyme exhibited limited durability, with a significant loss of activity within 24 h at 37 °C [157]. Presently, numerous research groups have made substantial contributions to the study of PETase [158,159,160]. Recently, Cui et al. introduced a novel computational design strategy for enhancing protein stability, known as the Greedy Accumulated Strategy for Protein Engineering (GRAPE). As a result of this innovative approach, a catalytic enzyme called DuraPETase was engineered and its ability to degrade poly(ethylene terephthalate) MPs was successfully demonstrated. Figure 9A illustrates the degradation of poly(ethylene terephthalate) into smaller molecules or non-toxic substances. The scanning electron microscopy (SEM) image in Figure 9B clearly demonstrates noticeable alterations on the surface of poly(ethylene terephthalate) MPs. Furthermore, the high-performance liquid chromatography analysis in Figure 9C confirms the effective degradation of poly(ethylene terephthalate) MPs by DuraPETase enzymes [161]. The enhancement of ispetase stability was achieved through the utilization of state-of-the-art computational protein design techniques. This groundbreaking approach yielded a redesigned enzyme with remarkable resilience, effectively addressing the long-standing issues of enzyme instability and fragility. Importantly, this breakthrough opens up new possibilities for the utilization of biodegradable plastics. However, it is worth noting that the widespread implementation of this method is hampered by its prohibitive costs and the intricate process required for enzyme preparation [162]. In future studies, researchers should control treatment costs while maintaining the efficient treatment of microplastics.

3.3.3. Bacterial Degradation Method

Bacteria, a prominent group of microorganisms, reign supreme as the most abundant lifeforms across all organisms. The morphological diversity of bacteria is striking, ranging from spherical to rod-shaped and even spiral forms [163,164]. Presently, bacteria find wide applications in the production of cheese, yogurt, wine, and antibiotics. Furthermore, their remarkable potential in wastewater pollutant degradation cannot be overlooked [165,166]. Researchers have harnessed the power of bacteria in the degradation of microplastics (MPs). These bacteria primarily originate from sediments, sludge, and MP-laden water bodies [167,168,169]. For instance, Auta et al. isolated two pure bacteria from mangrove sediments and employed them in the degradation of polypropylene MPs. After a span of 40 days, the degradation rates for polypropylene MPs by Rhodococcus 36 and Bacillus 27 were recorded at 6.4% and 4.0%, respectively. The degradation process induced the formation of various porous structures and irregularities on the surface of the MPs, providing compelling evidence of the bacteria’s efficacy in polypropylene MP degradation. Auta et al. employed Bacillus cereus and Bacillus gotthiilii bacteria to facilitate the degradation of various types of microplastics (MPs). Their findings revealed that Bacillus cereus achieved weight reductions of 1.6%, 6.6%, and 7.4% for polyphylene, polyphylene terephthalate, and polystyrene, respectively. On the other hand, Bacillus gotthielii displayed weight reduction rates of 6.2%, 3.0%, 3.6%, and 5.8% for polyphylene, polyphylene terephthalate, polypropelene, and polystyrene, respectively [170]. Yang et al. successfully isolated Enterobacter asburiae YT1 from plastic-eating waxworms and employed it in the degradation of polyethylene MPs. After a duration of 28 days, the weight reduction rate of polyethylene was observed to be 6.1% ± 0.3 [171]. Similarly, Shah et al. isolated a strain of Bacillus subtilis mza-75 from soil samples and utilized it to degrade polyurethane MPs. Following a 28-day period, the scanning electron microscope (SEM) image revealed the emergence of extensive crack formations on the polyurethane surface, while Fourier-transform infrared (FTIR) spectroscopy indicated a decline in the functional groups of polyurethanes. These phenomena collectively indicate the efficacy of Bacillus subtilis mza-75 in the degradation of polyurethane MPs [172]. Furthermore, Yoshida et al. isolated a strain of ideanella sakaiensis 201-f6 from contaminated samples and employed it in the degradation of polyethylene and polyethylene terephthalate MPs. Remarkably, after a duration of 60 days, the SEM image confirmed the capability of ideonella sakaiensis 201-f6 to degrade polyethylene and polyethylene terephthalate MPs [173]. In conclusion, the bacteria-mediated degradation of MPs represents a widely utilized treatment technology in the management of MPs. However, this approach is not without its drawbacks, including the complexity of bacterial culture, the stringent environmental conditions required for bacterial degradation, and the substantial time investment involved. In future studies, researchers should simplify the steps of bacterial culture and train more adaptable bacteria. On the one hand, the processing cost is reduced, and on the other hand, the anti-risk capability of this method is enhanced. At the same time, it is also necessary to pay attention to other hazards to the environment caused by bacteria that may leak during the treatment process.

3.3.4. Fungal Degradation Method

Fungi have been frequently employed for the degradation of hazardous pollutants [174,175]. Presently, numerous researchers have embarked on employing fungi for the degradation of microplastics (MPs) [176,177,178]. For instance, Yamada-Onodera et al. employed Penicillium simplicissimum YK fungi to degrade polyethylene MPs, leading to the subsequent observation of polyethylene MPs with reduced molecular weight after a 3-month period of liquid culture. These findings unequivocally demonstrate the remarkable proficiency of Penicillium simplicissimum YK fungi in degrading polyethylene MPs [179]. Volke-Sepúlveda et al. isolated and cultivated Aspergillus niger and Penicillium pinophilum fungi, which were then employed for the degradation of low-density polyethylene (TO-LDPE) MPs. After 31 months, Aspergillus niger and Penicillium pinophilum achieved weight reduction rates of 0.57% and 0.37% for TO-LDPE, respectively [180]. Devi et al. isolated and cultured Aspergillus tubingensis VRKPT1 and Aspergillus flavus VRKPT2 from discarded polyethylene waste, subsequently utilizing them for the degradation of high-density polyethylene (HDPE) MPs. After 30 days, Aspergillus tubingensis VRKPT1 and Aspergillus flavus VRKPT2 demonstrated weight reduction rates of 6.88 ± 0.1% and 9.34 ± 0.2%, respectively, for HDPE MPs [181]. El-Shafei et al. isolated and cultivated VRKPT1 and VRKPT2 fungi from the Nile Delta, employing them for the degradation of high-density polyethylene (HDPE) MPs. After 30 days, VRKPT1 and VRKPT2 achieved weight reduction rates of 6.02 ± 0.2% and 8.51 ± 0.1%, respectively [182]. In conclusion, these fungi possess remarkable capabilities for the in vitro degradation of MPs. Nonetheless, this approach has certain limitations, such as the need for intricate bacterial culture, demanding environmental prerequisites for fungal degradation, and time-intensive processes. Consequently, further refinement is imperative for the fungal-mediated degradation of MPs. In future studies, researchers should optimize the conditions for fungi to process microplastics, so that fungi can treat microplastics in complex environments with the shortest time, the lowest cost, and the highest efficiency.
As elucidated earlier, Table 1 succinctly delineates the merits and demerits of various methodologies employed for the treatment of microplastics (MPs). Consequently, it is evident that substantial progress is yet to be made in the realm of eradicating MPs.

4. Challenges, Future Research, and Research Limitations

In this paper, a large number of existing studies on the characterization and treatment techniques of microplastics are reviewed. However, these characterization and treatment techniques have some shortcomings: their universality is poor, and it is difficult to use a single method to properly characterize and treat microplastics. In view of the current waste production, the pollution of microplastics in the environment may become more and more serious, and the difficulty of detecting and disposing of microplastics will become more and more difficult. Therefore, we need to pay more attention to the source of microplastics, the transfer route, the toxicity of degradation intermediates, and other aspects in future research processes.

4.1. Challenges and Suggestions

(1)
MPs come from a wide range of sources, and the challenge remains of whether researchers can take certain measures from the time of plastic production to avoid the subsequent generation of MPs.
(2)
There are many ways in which MPs are transferred, which forms a cycle in the environment; it is important that we dispose of MPs at the stage where we are most likely to dispose of them.
(3)
Current research is mainly concerned with the removal rate of MPs, but it is ignored that MPs may produce toxic substances during the treatment process and cause other impacts on the environment.
(4)
The removal rate of MPs mentioned in the literature is very high, but whether the treated MPs are transformed into other substances that are still harmful to the environment is unknown. And there is no standardized way to judge the extent to which a treatment is environmentally friendly.
(5)
For biodegradable MPs, whether the degradation process has an impact on microorganisms, thereby causing harm to biological groups, and whether it will indirectly affect the environment are unknown.

4.2. Future Recommendations

(1)
It is recommended to link a variety of characterization and treatment technologies to find a more universal characterization and treatment technology to solve the complex MP pollution situation.
(2)
Toxicity studies are recommended to assess the toxicity of MPs’ degradation intermediates and degradation products.
(3)
At present, it seems that it may be easier to treat MPs in water, but they are more difficult to treat soil or the atmosphere, and it is recommended to develop a method that can deal with MP pollutants in soil and the atmosphere.
(4)
It is recommended to establish a set of evaluation methods to systematically evaluate the degradation efficiency and toxicity of the degradation products of MPs.
(5)
It is recommended to improve the classification of waste at the source and plastic recycling.

4.3 Research Limitations

The number of research papers on MPs has soared in recent years, but overall MPs are still understudied. Although many methods for characterizing and treating MPs have been summarized in this paper, there are still many methods that we have not summarized. Some of the methods described in this paper only describe a single characterization and treatment technology, and do not combine multiple methods to study MPs. As the study of MPs deepens, more and more technologies can be used to characterize and remove MPs, and more and more advanced methods will be derived.

5. Conclusions

The development of effective, sustainable, and uncomplicated methodologies for the eradication of microplastics (MPs) holds paramount significance in the amelioration of plastic pollution. In this study, we conducted a systematic appraisal of the techniques employed for characterizing and removing MPs. At the same time, according to the technical gaps not involved in the literature, some suggestions for future research are put forward. The characterization of MPs proved invaluable in discerning their diverse typologies. Moreover, we conducted a comprehensive evaluation of the merits and demerits associated with various removal technologies, encompassing filtration, adsorption, extraction, magnetic separation, oil film separation, Fenton oxidation, electrochemical oxidation, persulfate advanced oxidation, photocatalytic oxidation, coagulation, electrocoagulation, foam flotation, anaerobic–anoxic–aerobic activated sludge, enzymatic degradation, bacterial degradation, and fungal degradation. Drawing upon the findings of our research team, we proposed a viable and efficacious degradation system for the elimination of MPs. Each approach has its own advantages, but there are also major limitations. These physical methods have the advantages of simple operation and low cost, but they have certain limitations on the treatment effect of microplastics. Chemical methods have a good effect on the treatment of microplastics, but can easily to produce other substances that harm the environment. Biological methods are a new approach; they have the advantages of high efficiency and environmental protection, but their mechanism is not sufficient, and they need further research and development. However, it is imperative to note that numerous experiments have thus far been confined to laboratory settings, with limited exploration of removal techniques for minute MPs. Hence, expanding the scale of experimentation and devising strategies for the elimination of diminutive MPs emerge as the focal points for future researchers.

Author Contributions

Y.L.: Data Curation, Writing—Original Draft, Conceptualization, Methodology, Investigation, Formal Analysis. P.C.: Investigation. Y.T.: Investigation. Y.Y.: Investigation, Supervision. C.Z.: Investigation, Supervision. J.B.: Writing—Review and Editing, Supervision, Validation. S.Z.: Writing—Review and Editing, Visualization, Resources. All authors have read and agreed to the published version of the manuscript.

Funding

The National Natural Science Foundation of China (Grant No.21576295), the Hunan Provincial Natural Science Foundation of China (2019JJ50759), and the Fundamental Research Funds for the Central Universities of Central South University (2017zzts175 and 2018zzts371).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Data are available on request from the authors.

Acknowledgments

This study thanks for the funding of National Natural Science Foundation of China (Grant No.21576295), the Hunan Provincial Natural Science Foundation of China (2019JJ50759), and the Fundamental Research Funds for the Central Universities of Central South University (2017zzts175 and 2018zzts371).

Conflicts of Interest

Author Jiaqi Bu was employed by the company Hunan Haili Chemical Industry Company. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conmpanylict of interest.

References

  1. Van Cauwenberghe, L.; Devriese, L.; Galgani, F.; Robbens, J.; Janssen, C.R. Microplastics in sediments: A review of techniques, occurrence and effects. Mar. Environ. Res. 2015, 111, 5–17. [Google Scholar] [CrossRef] [PubMed]
  2. Thompson, R.C.; Olsen, Y.; Mitchell, R.P.; Davis, A.; Rowland, S.J.; John, A.W.G.; McGonigle, D.; Russell, A.E. Lost at sea: Where is all the plastic? Science 2004, 304, 838. [Google Scholar] [CrossRef] [PubMed]
  3. Andrady, A.L. Microplastics in the marine environment. Mar. Pollut. Bull. 2011, 62, 1596–1605. [Google Scholar] [CrossRef] [PubMed]
  4. Lu, X.; Hu, H.W.; Li, J.W.; Li, J.P.; Wang, L.J.; Liu, L.L.; Tang, Y.Y. Microplastics existence affected heavy metal affinity to ferrihydrite as a representative sediment mineral. Sci. Total Environ. 2023, 859, 160227. [Google Scholar] [CrossRef] [PubMed]
  5. Di, M.X.; Wang, J. Microplastics in surface waters and sediments of the Three Gorges Reservoir, China. Sci. Total Environ. 2018, 616, 1620–1627. [Google Scholar] [CrossRef] [PubMed]
  6. Ma, M.L.; Xu, D.Y.; Zhao, J.; Gao, B. Disposable face masks release micro particles to the aqueous environment after simulating sunlight aging: Microplastics or non-microplastics? J. Hazard. Mater. 2022, 443, 130146. [Google Scholar] [CrossRef] [PubMed]
  7. Schymanski, D.; Goldbeck, C.; Humpf, H.U.; Fuerst, P. Analysis of microplastics in water by micro-Raman spectroscopy: Release of plastic particles from different packaging into mineral water. Water Res. 2018, 129, 154–162. [Google Scholar] [CrossRef] [PubMed]
  8. Shi, Y.Z.; Yi, L.; Du, G.R.; Hu, X.; Huang, Y. Visual characterization of microplastics in corn flour by near field molecular spectral imaging and data mining. Sci. Total Environ. 2023, 862, 160714. [Google Scholar] [CrossRef] [PubMed]
  9. Oßmann, B.E.; Sarau, G.; Holtmannspötter, H.; Pischetsrieder, M.; Christiansen, S.H.; Dicke, W. Small-sized microplastics and pigmented particles in bottled mineral water. Water Res. 2018, 141, 307–316. [Google Scholar] [CrossRef]
  10. Liu, M.T.; Lu, S.B.; Song, Y.; Lei, L.L.; Hu, J.N.; Lv, W.W.; Zhou, W.Z.; Cao, C.J.; Shi, H.H.; Yang, X.F.; et al. Microplastic and mesoplastic pollution in farmland soils in suburbs of shanghai, china. Environ. Pollut. 2018, 242, 855–862. [Google Scholar] [CrossRef]
  11. Martín, C.; Fajardo, C.; Costa, G.; Sánchez-Fortún, S.; San Andres, M.D.; Gonzalez, F.; Nande, M.; Mengs, G.; Martín, M. Bioassays to assess the ecotoxicological impact of polyethylene microplastics and two organic pollutants, simazine and ibuprofen. Chemosphere 2021, 274, 129704. [Google Scholar] [CrossRef] [PubMed]
  12. Choi, D.; Hwang, J.; Bang, J.; Han, S.; Kim, T.; Oh, Y.; Hwang, Y.; Choi, J.; Hong, J. In vitro toxicity from a physical perspective of polyethylene microplastics based on statistical curvature change analysis. Sci. Total Environ. 2021, 752, 142242. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, L.X.; Wang, Y.X.; Xu, M.; Ma, J.; Zhang, S.P.; Liu, S.J.; Wang, K.; Tian, H.F.; Cui, J.S. Enhanced hepatic cytotoxicity of chemically transformed polystyrene microplastics by simulated gastric fluid. J. Hazard. Mater. 2021, 410, 124536. [Google Scholar] [CrossRef] [PubMed]
  14. Praveena, S.M.; Shaifuddin, S.N.M.; Akizuki, S. Exploration of microplastics from personal care and cosmetic products and its estimated emissions to marine environment: An evidence from Malaysia. Mar. Pollut. Bull. 2018, 136, 135–140. [Google Scholar] [CrossRef] [PubMed]
  15. Guerranti, C.; Martellini, T.; Perra, G.; Scopetani, C.; Cincinelli, A. Microplastics in cosmetics: Environmental issues and needs for global bans. Environ. Toxicol. Pharmacol. 2019, 68, 75–79. [Google Scholar] [CrossRef] [PubMed]
  16. Sun, Q.; Ren, S.Y.; Ni, H.G. Incidence of microplastics in personal care products: An appreciable part of plastic pollution. Sci. Total Environ. 2020, 742, 140218. [Google Scholar] [CrossRef] [PubMed]
  17. Stanton, T.; Johnson, M.; Nathanail, P.; MacNaughtan, W.; Gomes, R.L. Freshwater and airborne textile fibre populations are dominated by ‘natural’, not microplastic, fibres. Sci. Total Environ. 2019, 666, 377–389. [Google Scholar] [CrossRef]
  18. He, X.Q.; Li, H.B.; Zhu, J. A value-added insight of reusing microplastic waste: Carrier particle in fluidized bed bioreactor for simultaneous carbon and nitrogen removal from septic wastewater. Biochem. Eng. J. 2019, 151, 107300. [Google Scholar] [CrossRef]
  19. Hamidian, A.H.; Ozumchelouei, E.J.; Feizi, F.; Wu, C.X.; Zhang, Y.; Yang, M. A review on the characteristics of microplastics in wastewater treatment plants: A source for toxic chemicals. J. Clean. Prod. 2021, 295, 126480. [Google Scholar] [CrossRef]
  20. Silvestrova, K.; Stepanova, N. The distribution of microplastics in the surface layer of the Atlantic Ocean from the subtropics to the equator according to visual analysis. Mar. Pollut. Bull. 2021, 162, 111836. [Google Scholar] [CrossRef]
  21. Li, C.J.; Wang, X.H.; Liu, K.; Zhu, L.X.; Wei, N.; Zong, C.X.; Li, D.J. Pelagic microplastics in surface water of the Eastern Indian Ocean during monsoon transition period: Abundance, distribution, and characteristics. Sci. Total Environ. 2021, 755, 142629. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, K.; Wang, X.H.; Fang, T.; Xu, P.; Zhu, L.X.; Li, D.J. Source and potential risk assessment of suspended atmospheric microplastics in Shanghai. Sci. Total Environ. 2019, 675, 462–471. [Google Scholar] [CrossRef] [PubMed]
  23. De Falco, F.; Cocca, M.; Avella, M.; Thompson, R.C. Microfiber release to water, via laundering, and to air, via everyday use: A comparison between polyester clothing with differing textile parameters. Environ. Sci. Technol. 2020, 54, 3288–3296. [Google Scholar] [CrossRef]
  24. Xiao, S.L.; Cui, Y.F.; Brahney, J.; Mahowald, N.M.; Li, Q. Long-distance atmospheric transport of microplastic fibres influenced by their shapes. Nat. Geosci. 2023, 16, 863–870. [Google Scholar] [CrossRef]
  25. Leslie, H.A.; van Velzen, M.J.M.; Brandsma, S.H.; Vethaak, A.D.; Garcia-Vallejo, J.J.; Lamoree, M.H. Discovery and quantification of plastic particle pollution in human blood. Environ. Int. 2022, 163, 107199. [Google Scholar] [CrossRef] [PubMed]
  26. Aves, A.R.; Revell, L.E.; Gaw, S.; Ruffell, H.; Schuddeboom, A.; Wotherspoon, N.E.; LaRue, M.; McDonald, A.J. First evidence of microplastics in Antarctic snow. Cryosphere 2022, 16, 2127–2145. [Google Scholar] [CrossRef]
  27. Yin, K.; Wang, Y.; Zhao, H.J.; Wang, D.X.; Guo, M.H.; Mu, M.Y.; Liu, Y.C.; Nie, X.P.; Li, B.Y.; Li, J.Y.; et al. A comparative review of microplastics and nanoplastics: Toxicity hazards on digestive, reproductive and nervous system. Sci. Total Environ. 2021, 774, 145758. [Google Scholar] [CrossRef]
  28. Klingelhofer, D.; Braun, M.; Quarcoo, D.; Bruggmann, D.; Groneberg, D.A. Research landscape of a global environmental challenge: Microplastics. Water Res. 2020, 170, 115358. [Google Scholar] [CrossRef] [PubMed]
  29. Wright, S.L.; Kelly, F.J. Plastic and human health: A micro lssue? Environ. Sci. Technol. 2017, 51, 6634–6647. [Google Scholar] [CrossRef]
  30. Mak, C.W.; Yeung, K.C.F.; Chan, K.M. Acute toxic effects of polyethylene microplastic on adult zebrafish. Ecotoxicol. Environ. Saf. 2019, 182, 109442. [Google Scholar] [CrossRef]
  31. Yang, H.; Xiong, H.R.; Mi, K.H.; Xue, W.; Wei, W.Z.; Zhang, Y.Y. Toxicity comparison of nano-sized and micron-sized microplastics to goldfish Carassius auratus larvae. J. Hazard. Mater. 2020, 388, 122058. [Google Scholar] [CrossRef] [PubMed]
  32. Xie, Y.C.; Li, Y.; Feng, Y.; Cheng, W.; Wang, Y. Inhalable microplastics prevails in air: Exploring the size detection limit. Environ. Int. 2022, 162, 107151. [Google Scholar] [CrossRef] [PubMed]
  33. Jiang, X.F.; Yang, Y.; Wang, Q.; Liu, N.; Li, M. Seasonal variations and feedback from microplastics and cadmium on soil organisms in agricultural fields. Environ. Int. 2022, 161, 107096. [Google Scholar] [CrossRef] [PubMed]
  34. Fan, P.; Yu, H.; Xi, B.D.; Tan, W.B. A review on the occurrence and influence of biodegradable microplastics in soil ecosystems: Are biodegradable plastics substitute or threat? Environ. Int. 2022, 163, 107244. [Google Scholar] [CrossRef] [PubMed]
  35. Dai, Y.J.; Shi, J.J.; Zhang, N.X.; Pan, Z.L.; Xing, C.M.; Chen, X. Current research trends on microplastics pollution and impacts on agro-ecosystems: A short review. Sep. Sci. Technol. 2022, 57, 656–669. [Google Scholar] [CrossRef]
  36. Rezaei, M.; Abbasi, S.; Pourmahmood, H.; Oleszczuk, P.; Ritsema, C.; Turner, A. Microplastics in agricultural soils from a semi-arid region and their transport by wind erosion. Environ. Res. 2022, 212, 113213. [Google Scholar] [CrossRef] [PubMed]
  37. Perez, C.N.; Carré, F.; Hoarau-Belkhiri, A.; Joris, A.; Leonards, P.E.; Lamoree, M.H. Innovations in analytical methods to assess the occurrence of microplastics in soil. J. Environ. Chem. Eng. 2022, 10, 107421. [Google Scholar] [CrossRef]
  38. Grause, G.; Kuniyasu, Y.; Chien, M.F.; Inoue, C. Separation of microplastic from soil by centrifugation and its application to agricultural soil. Chemosphere 2022, 288, 132654. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, S.; Yang, X.; Gertsen, H.; Peters, P.; Salánki, T.; Geissen, V. A simple method for the extraction and identification of light density microplastics from soil. Sci. Total Environ. 2018, 616, 1056–1065. [Google Scholar] [CrossRef]
  40. Li, W.; Wufuer, R.; Duo, J.; Wang, S.; Luo, Y.; Zhang, D.; Pan, X. Microplastics in agricultural soils: Extraction and characterization after different periods of polythene film mulching in an arid region. Sci. Total Environ. 2020, 749, 141420. [Google Scholar] [CrossRef]
  41. Wander, L.; Lommel, L.; Meyer, K.; Braun, U.; Paul, A. Development of a low-cost method for quantifying microplastics in soils and compost using near-infrared spectroscopy. Meas. Sci. Technol. 2022, 33, 075801. [Google Scholar] [CrossRef]
  42. Way, C.; Hudson, M.D.; Williams, I.D.; Langley, G.J. Evidence of underestimation in microplastic research: A meta-analysis of recovery rate studies. Sci. Total Environ. 2022, 805, 150227. [Google Scholar] [CrossRef] [PubMed]
  43. Cowger, W.; Gray, A.; Christiansen, S.H.; DeFrond, H.; Deshpande, A.D.; Hemabessiere, L.; Lee, E.; Mill, L.; Munno, K.; Ossmann, B.E.; et al. Critical review of processing and classification techniques for images and spectra in microplastic research. Appl. Spectrosc. 2020, 74, 989–1010. [Google Scholar] [CrossRef] [PubMed]
  44. McDermid, K.J.; McMullen, T.L. Quantitative Analysis of Small-Plastic Debris on Beaches in the Hawaiian Archipelago. Mar. Pollut. Bull. 2004, 48, 790–794. [Google Scholar] [CrossRef] [PubMed]
  45. Li, W.C.; Tse, H.F.; Fok, L. Plastic waste in the marine environment: A review of sources, occurrence and effects. Sci. Total Environ. 2016, 566, 333–349. [Google Scholar] [CrossRef] [PubMed]
  46. González-Pleiter, M.; Edo, C.; Aguilera, Á.; Viúdez-Moreiras, D.; Pulido-Reyes, G.; González-Toril, E.; Osuna, S.; De Diego-Castilla, G.; Leganés, F.; Fernández-Piñas, F.; et al. Occurrence and transport of microplastics asmpled within and above the planetary boundary layer. Sci. Total Environ. 2021, 761, 143213. [Google Scholar] [CrossRef] [PubMed]
  47. De Witte, B.; Devriese, L.; Bekaert, K.; Hoffman, S.; Vandermeersch, G.; Cooreman, K.; Robbens, J. Quality assessment of the blue mussel (Mytilus edulis): Comparison between commercial and wild types. Mar. Pollut. Bull. 2014, 85, 146–155. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, Z.M.; Wagner, J.; Ghosal, S.; Bedi, G.; Wall, S. SEM/EDS and opticalmicroscopy analyses of microplastics in ocean trawl and fish guts. Sci. Total Environ. 2017, 603, 616–626. [Google Scholar] [CrossRef] [PubMed]
  49. Song, Y.K.; Hong, S.H.; Jang, M.; Han, G.M.; Rani, M.; Lee, J.; Shim, W.J. A comparison of microscopic and spectroscopic identification methods for analysis of microplastics in environmental samples. Mar. Pollut. Bull. 2015, 93, 202–209. [Google Scholar] [CrossRef]
  50. Eriksen, M.; Mason, S.; Wilson, S.; Box, C.; Zellers, A.; Edwards, W.; Farley, H.; Amato, S. Microplastic pollution in the surface waters of the Laurentian Great Lakes. Mar. Pollut. Bull. 2013, 77, 177–182. [Google Scholar] [CrossRef]
  51. Tirkey, A.; Upadhyay, L.S.B. Micoplastics: An overview on separation, identification and characterization of microplastics. Mar. Pollut. Bull. 2021, 170, 112604. [Google Scholar] [CrossRef] [PubMed]
  52. Fries, E.; Dekiff, J.H.; Willmeyer, J.; Nuelle, M.T.; Ebert, M.; Remy, D. Identification of polymer types and additives in marine microplastic particles using pyrolysis-GC/MS and scanning electron microscopy. Environ. Sci. Process. Impacts 2013, 15, 1949–1956. [Google Scholar] [CrossRef] [PubMed]
  53. Cooper, D.A.; Corcoran, P.L. Effects of mechanical and chemical processes on the dgradation of plastic beach debris on the island of Kauai, Hawaii. Mar. Pollut. Bull. 2010, 60, 650–654. [Google Scholar] [CrossRef] [PubMed]
  54. Qiu, Q.X.; Tan, Z.; Wang, J.D.; Peng, J.P.; Li, M.M.; Zhan, Z.W. Extraction, enumeration and identification methods for monitoring microplastics in the environment. Estuar. Coast. Shelf. Sci. 2016, 176, 102–109. [Google Scholar] [CrossRef]
  55. Nolte, T.M.; Hartmann, N.B.; Kleijn, J.M.; Garnæs, J.; van de Meent, D.; Jan Hendriks, A.; Baun, A. The toxicity of plastic nanoparticles to green algae as influenced by surface modification, medium hardness and cellular adsorption. Aquat. Toxicol. 2017, 183, 11–20. [Google Scholar] [CrossRef] [PubMed]
  56. Demir-Yilmaz, I.; Yakovenko, N.; Roux, C.; Guiraud, P.; Collin, F.; Coudret, C.; Halle, A.T.; Formosa-Dague, C. The role of microplastics in microalgae cells aggregation: A study at the molecular scale using atomic force microscopy. Sci. Total Environ. 2022, 832, 155036. [Google Scholar] [CrossRef] [PubMed]
  57. Meister, A.; Gabi, M.; Behr, P.; Studer, P.; Vörös, J.; Niedermann, P.; Bitterli, J.; Polesel-Maris, J.; Liley, M.; Heinzelmann, H.; et al. FluidFM: Combining atomic force microscopy and nanofluidics in a universal liquid delivery system for single cell applications and beyond. Nano Lett. 2009, 9, 2501–2507. [Google Scholar] [CrossRef] [PubMed]
  58. Kumari, A.; Chaudhary, D.R.; Jha, B. Destabilization of polyethylene and polyvinylchloride structure by marine bacterial strain. Environ. Sci. Pollut. Res. 2019, 26, 1507–1516. [Google Scholar] [CrossRef]
  59. He, D.F.; Luo, Y.M.; Lu, S.B.; Liu, M.T.; Song, Y.; Lei, L.L. Microplastics in soils: Analytical methods, pollution characteristics and ecological risks. TrAC Trends Anal. Chem. 2018, 109, 163–172. [Google Scholar] [CrossRef]
  60. Fan, C.H.; Huang, Y.Z.; Lin, J.N.; Li, J.W. Microplastic constituent identification from admixtures by Fourier-transform infrared (FTIR) spectroscopy: The use of polyethylene terephthalate (PET), polyethylene (PE), polypropylene (PP), polyvinyl chloride (PVC) and nylon (NY) as the model constituents. Environ. Technol. Innov. 2021, 23, 101798. [Google Scholar] [CrossRef]
  61. Caldwell, J.; Petri-Fink, A.; Rothen-Rutishauser, B.; Lehner, R. Assessing meso- and microplastic pollution in the ligurian and tyrrhenian seas. Mar. Pollut. Bull. 2019, 149, 110572. [Google Scholar] [CrossRef]
  62. Lin, J.Y.; Liu, H.T.; Zhang, J. Recent advances in the application of machine learning methods to improve identification of the microplastics in environment. Chemosphere 2022, 307, 136092. [Google Scholar] [CrossRef]
  63. Wander, L.; Vianello, A.; Vollertsen, J.; Westad, F.; Braun, U.; Paul, A. Exploratory analysis of hyperspectral FTIR data obtained from environmental microplastics samples. Anal. Methods 2020, 12, 781–791. [Google Scholar] [CrossRef]
  64. Chen, Y.L.; Li, T.C.; Hu, H.J.; Ao, H.Y.; Xiong, X.; Shi, H.H.; Wu, C.X. Transport and fate of microplastics in constructed wetlands: A microcosm study. J. Hazard. Mater. 2021, 415, 125615. [Google Scholar] [CrossRef]
  65. Renner, G.; Schmidt, T.C.; Schram, J. A New Chemometric Approach for Automatic Identification of Microplastics from Environmental Compartments Based on FT-IR Spectroscopy. Anal. Chem. 2017, 89, 12045–12053. [Google Scholar] [CrossRef] [PubMed]
  66. Araujo, C.F.; Nolasco, M.M.; Ribeiro, A.M.P.; Ribeiro-Claro, P.J.A. Identification of microplastics using raman spectroscopy: Latest developments and future prospects. Water Res. 2018, 142, 426–440. [Google Scholar] [CrossRef]
  67. Sobhani, Z.; Al Amin, M.; Naidu, R.; Megharaj, M.; Fang, C. Identification and visualisation of microplastics by Raman mapping. Anal. Chim. Acta 2019, 1077, 191–199. [Google Scholar] [CrossRef] [PubMed]
  68. Becucci, M.; Mancini, M.; Campo, R.; Paris, E. Microplastics in the Florence wastewater treatment plant studied by a continuous sampling method and Raman spectroscopy: A preliminary investigation. Sci. Total Environ. 2021, 808, 152025. [Google Scholar] [CrossRef] [PubMed]
  69. Cole, M.; Lindeque, P.; Fileman, E.; Halsband, C.; Goodhead, R.; Moger, J.; Galloway, T.S. Microplastic Ingestion by Zooplankton. Environ. Sci. Technol. 2013, 47, 6646–6655. [Google Scholar] [CrossRef]
  70. Kappler, A.; Fischer, D.; Oberbeckmann, S.; Schernewski, G.; Labrenz, M.; Eichhorn, K.J.; Voit, B. Analysis of environmental microplastics by vibrational microspectroscopy: FTIR, Raman or both? Anal. Bioanal. Chem. 2016, 408, 8377–8391. [Google Scholar] [CrossRef]
  71. Vilakati, B.; Sivasankar, V.; Nyoni, H.; Mamba, B.B.; Omine, K.; Msagati, T.A.M. The Py-GC-TOF-MS analysis and characterization of microplastics (MPs) in a wastewater treatment plant in Gauteng Province, South Africa. Ecotoxicol. Environ. Saf. 2021, 222, 112478. [Google Scholar] [CrossRef] [PubMed]
  72. Majewsky, M.; Bitter, H.; Eiche, E.; Horn, H. Determination of microplastic polyethylene (PE) and polypropylene (PP) in environmental samples using thermal analysis (TGA-DSC). Sci. Total Environ. 2016, 568, 507–511. [Google Scholar] [CrossRef] [PubMed]
  73. Dumichen, E.; Barthel, A.K.; Braun, U.; Bannick, C.G.; Brand, K.; Jekel, M.; Senz, R. Analysis of polyethylene microplastics in environmental samples, using a thermal decomposition method. Water Res. 2015, 85, 451–457. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, X.Q.; Tian, K.; Chen, Z.J.; Wei, W.; Xu, B.T.; Ni, B.J. Online TG-FTIR-MS analysis of the catalytic pyrolysis of polyethylene and polyvinyl chloride microplastics. J. Hazard. Mater. 2023, 441, 129881. [Google Scholar] [CrossRef] [PubMed]
  75. Sarkar, D.J.; Sarkar, S.D.; Das, B.K.; Praharaj, J.K.; Mahajan, D.K.; Purokait, B.; Mohanty, T.R.; Mohanty, D.; Gogoi, P.; Kumar, V.; et al. Microplastics removal efficiency of drinking water treatment plant with pulse clarifier. J. Hazard. Mater. 2021, 413, 125347. [Google Scholar] [CrossRef] [PubMed]
  76. Sundbaek, K.B.; Koch, I.D.W.; Villaro, C.G.; Rasmussen, N.S.; Holdt, S.L.; Hartmann, N.B. Sorption of fluorescent polystyrene microplastic particles to edible seaweed Fucus vesiculosus. J. Appl. Phycol. 2018, 30, 2923–2927. [Google Scholar] [CrossRef]
  77. Catarino, A.I.; Thompson, R.; Sanderson, W.; Henry, T.B. Development and optimization of a standard method for extraction of microplastics in mussels by enzyme digestion of soft tissues. Environ. Toxicol. Chem. 2017, 36, 947–951. [Google Scholar] [CrossRef]
  78. Shi, X.H.; Zhang, X.T.; Gao, W.; Zhang, Y.L.; He, D.F. Removal of microplastics from water by magnetic nano-Fe3O4. Sci. Total Environ. 2022, 802, 149838. [Google Scholar] [CrossRef] [PubMed]
  79. Lares, M.; Ncibi, M.C.; Sillanpaa, M.; Sillanpaa, M. Intercomparison study on commonly used methods to determine microplastics in wastewater and sludge samples. Environ. Sci. Pollut. Res. 2019, 26, 12109–12122. [Google Scholar] [CrossRef]
  80. Zou, W.; Xia, M.L.; Jiang, K.; Cao, Z.G.; Zhang, X.L.; Hu, X.G. Photo-Oxidative Degradation Mitigated the Developmental Toxicity of Polyamide Microplastics to Zebrafish Larvae by Modulating Macrophage-Triggered Proinflammatory Responses and Apoptosis. Environ. Sci. Technol. 2020, 54, 13888–13898. [Google Scholar] [CrossRef]
  81. Rajala, K.; Grönfors, O.; Hesampour, M.; Mikola, A. Removal of microplastics from secondary wastewater treatment plant effluent by coagulation/flocculation with iron, aluminum and polyamine-based chemicals. Water Res. 2020, 183, 116045. [Google Scholar] [CrossRef] [PubMed]
  82. Bolto, B.; Xie, Z.L. The Use of Polymers in the Flotation Treatment of Wastewater. Processes 2019, 7, 374. [Google Scholar] [CrossRef]
  83. Lee, H.; Kim, Y. Treatment characteristics of microplastics at biological sewage treatment facilities in Korea. Mar. Pollut. Bull. 2018, 137, 1–8. [Google Scholar] [CrossRef] [PubMed]
  84. Tournier, V.; Topham, C.M.; Gilles, A.; David, B.; Folgoas, C.; Moya-Leclair, E.; Kamionka, E.; Desrousseaux, M.L.; Texier, H.; Gavalda, S.; et al. An engineered PET depolymerase to break down and recycle plastic bottles. Nature 2020, 580, 216–219. [Google Scholar] [CrossRef] [PubMed]
  85. Park, S.Y.; Kim, C.G. Biodegradation of micro-polyethylene particles by bacterial colonization of a mixed microbial consortium isolated from a landfill site. Chemosphere 2019, 222, 527–533. [Google Scholar] [CrossRef]
  86. Liu, W.Y.; Zhang, J.L.; Liu, H.; Guo, X.N.; Zhang, X.Y.; Yao, X.L.; Cao, Z.G.; Zhang, T.T. A review of the removal of microplastics in global wastewater treatment plants: Characteristics and mechanisms. Environ. Int. 2021, 146, 106277. [Google Scholar] [CrossRef]
  87. Sol, D.; Laca, A.; Laca, A.; Díaz, M. Approaching the environmental problem of microplastics: Importance of WWTP treatments. Sci. Total Environ. 2020, 740, 140016. [Google Scholar] [CrossRef] [PubMed]
  88. Tadsuwan, K.; Babel, S. Microplastic contamination in a conventional wastewater treatment plant in Thailand. Waste Manag. Res. 2021, 39, 754–761. [Google Scholar] [CrossRef]
  89. Wang, Z.H.; Sedighi, M.; Lea-Langton, A. Filtration of microplastic spheres by biochar: Removal efficiency and immobilisation mechanisms. Water Res. 2020, 184, 116165. [Google Scholar] [CrossRef]
  90. Ziajahromi, S.; Neale, P.A.; Silveira, I.T.; Chua, A.; Leusch, F.D.L. An audit of microplastic abundance throughout three Australian wastewater treatment plants. Chemosphere 2021, 263, 128294. [Google Scholar] [CrossRef]
  91. Bu, J.Q.; Yuan, L.; Ren, Y.L.; Lv, Y.X.; Meng, Y.; Peng, X. Enhanced removal of Eriochrome Black T in wastewater by zirconium-based MOF/graphene oxide. Can. J. Chem. 2020, 98, 90–97. [Google Scholar] [CrossRef]
  92. Bu, J.Q.; Yuan, L.; Zhang, N.; Liu, D.; Meng, Y.; Peng, X. High-efficiency adsorption of methylene blue dye from wastewater by a thiosemicarbazide functionalized graphene oxide composite. Diam. Relat. Mater. 2020, 101, 107604. [Google Scholar] [CrossRef]
  93. Bu, J.Q.; Yuan, L.; Zhang, N.; Meng, Y.; Peng, X. Novel Adsorbent of N-Phenylthiourea-Functionalized Graphene Oxide and Its Removal of Methyl Orange in Aqueous Solutions. J. Chem. Eng. Date 2021, 66, 199–209. [Google Scholar] [CrossRef]
  94. Wang, J.C.; Yue, D.B.; Wang, H. In situ Fe3O4 nanoparticles coating of polymers for separating hazardous PVC from microplastic mixtures. Chem. Eng. J. 2021, 407, 127170. [Google Scholar] [CrossRef]
  95. Sajid, M.; Ihsanullah, I.; Tariq Khan, M.; Baig, N. Nanomaterials-based adsorbents for remediation of microplastics and nanoplastics in aqueous media: A review. Sep. Purif. Technol. 2023, 305, 122453. [Google Scholar] [CrossRef]
  96. Wang, Z.G.; Sun, C.Z.; Li, F.M.; Chen, L.Y. Fatigue resistance, re-usable and biodegradable sponge materials from plant protein with rapid water adsorption capacity for microplastics removal. Chem. Eng. J. 2021, 415, 129006. [Google Scholar] [CrossRef]
  97. Sun, C.Z.; Wang, Z.G.; Chen, L.Y.; Li, F.M. Fabrication of robust and compressive chitin and graphene oxide sponges for removal of microplastics with different functional groups. Chem. Eng. J. 2020, 393, 124796. [Google Scholar] [CrossRef]
  98. Yuan, F.; Yue, L.Z.; Zhao, H.; Wu, H.F. Study on the adsorption of polystyrene microplastics by three-dimensional reduced graphene oxide. Water Sci. Technol. 2020, 81, 2163–2175. [Google Scholar] [CrossRef]
  99. Cao, M.W.; Shen, Y.; Yan, Z.S.; Wei, Q.; Jiao, T.F.; Shen, Y.T.; Han, Y.C.; Wang, Y.L.; Wang, S.J.; Xia, Y.Q.; et al. Extraction-like removal of organic dyes from polluted water by the graphene oxide/PNIPAM composite system. Chem. Eng. J. 2021, 405, 126647. [Google Scholar] [CrossRef]
  100. Warrag, S.E.E.; Darwish, A.S.; Adeyemi, I.A.; Hadj-Kali, M.K.; Kroon, M.C.; AlNashef, I.M. Extraction of pyridine from n-alkane mixtures using methyltriphenylphosphonium bromide-based deep eutectic solvents as extractive denitrogenation agents. Fluid Phase Equilibr. 2020, 517, 112622. [Google Scholar] [CrossRef]
  101. Wagner, J.; Wang, Z.M.; Ghosal, S.; Rochman, C.; Gassel, M.; Wall, S. Novel method for the extraction and identification of microplastics in ocean trawl and fish gut matrices. Anal. Methods 2017, 9, 1479–1490. [Google Scholar] [CrossRef]
  102. Hurley, R.R.; Lusher, A.L.; Olsen, M.; Nizzetto, L. Validation of a Method for Extracting Microplastics from Complex, Organic-Rich, Environmental Matrices. Environ. Sci. Technol. 2018, 52, 7409–7417. [Google Scholar] [CrossRef]
  103. Li, C.T.; Cui, Q.; Zhang, M.; Vogt, R.D.; Lu, X.Q. A commonly available and easily assembled device for extraction of bio/non-degradable microplastics from soil by flotation in NaBr solution. Sci. Total Environ. 2021, 759, 143482. [Google Scholar] [CrossRef]
  104. Nuelle, M.T.; Dekiff, J.H.; Remy, D.; Fries, E. A new analytical approach for monitoring microplastics in marine sediments. Environ. Pollut. 2014, 184, 161–169. [Google Scholar] [CrossRef]
  105. Han, X.X.; Lu, X.Q.; Vogt, R.D. An optimized density-based approach for extracting microplastics from soil and sediment samples. Environ. Pollut. 2019, 254, 113009. [Google Scholar] [CrossRef]
  106. Wang, Z.; Taylor, S.E.; Sharma, P.; Flury, M. Poor extraction efficiencies of polystyrene nano- and microplastics from biosolids and soil. PLoS ONE 2018, 13, e0208009. [Google Scholar] [CrossRef]
  107. Li, X.D.; Wang, Y.H.; Lu, D.F.; Zheng, X.Y. Influence of Separation Angle on the Dry Pneumatic Magnetic Separation. Minerals 2022, 12, 1192. [Google Scholar] [CrossRef]
  108. Grbic, J.; Nguyen, B.; Guo, E.; You, J.B.; Sinton, D.; Rochman, C.M. Magnetic extraction of microplastics from environmental samples. Environ. Sci. Technol. Lett. 2019, 6, 68–72. [Google Scholar] [CrossRef]
  109. Tang, Y.; Zhang, S.H.; Su, Y.L.; Wu, D.; Zhao, Y.P.; Xie, B. Removal of microplastics from aqueous solutions by magnetic carbon nanotubes. Chem. Eng. J. 2021, 406, 126804. [Google Scholar] [CrossRef]
  110. Scopetani, C.; Chelazzi, D.; Mikola, J.; Leiniö, V.; Heikkinen, R.; Cincinelli, A.; Pellinen, J. Olive oil-based method for the extraction, quantification and identification of microplastics in soil and compost samples. Sci. Total Environ. 2020, 733, 139338. [Google Scholar] [CrossRef]
  111. Crichton, E.M.; Noel, M.; Gies, E.A.; Ross, P.S. A novel, density-independent and FTIR-compatible approach for the rapid extraction of microplastics from aquatic sediments. Anal. Methods 2017, 9, 1419–1428. [Google Scholar] [CrossRef]
  112. Mani, T.; Frehland, S.; Kalberer, A.; Burkhardt-Holm, P. Using castor oil to separate microplastics from four different environmental matrices. Anal. Methods 2019, 11, 1788–1794. [Google Scholar] [CrossRef]
  113. Thanh Truc, N.T.; Lee, B. Sustainable hydrophilization to separate hazardous chlorine PVC from plastic wastes using H2O2/ultrasonic irrigation. Waste Manag. 2019, 88, 28–38. [Google Scholar] [CrossRef]
  114. Yang, Y.T.; Chen, J.; Chen, Z.; Yu, Z.; Xue, J.C.; Luan, T.G.; Chen, S.S.; Zhou, S.G. Mechanisms of polystyrene microplastic degradation by the microbially driven Fenton reaction. Water Res. 2022, 223, 118979. [Google Scholar] [CrossRef]
  115. Xu, X.; Zong, S.; Chen, W.; Liu, D. Comparative study of bisphenol A degradation via heterogeneously catalyzed H2O2 and persulfate: Reactivity, products, stability and mechanism. Chem. Eng. J. 2019, 369, 470–479. [Google Scholar] [CrossRef]
  116. Liu, Y.; Zhang, J.D.; Cai, C.Y.; He, Y.; Chen, L.Y.; Xiong, X.; Huang, H.J.; Tao, S.; Liu, W.X. Occurrence and characteristics of microplastics in the Haihe River: An investigation of a seagoing river flowing through a megacity in northern China. Environ. Pollut. 2020, 262, 114261. [Google Scholar] [CrossRef]
  117. Liu, Y.L.; Lou, Z.M.; Yang, K.L.; Wang, Z.N.; Zhou, C.C.; Li, Y.Z.; Cao, Z.; Xu, X.H. Coagulation removal of Sb(V) from textile wastewater matrix with enhanced strategy: Comparison study and mechanism analysis. Chemosphere 2019, 237, 124494. [Google Scholar] [CrossRef] [PubMed]
  118. Lin, J.L.; Yan, D.Y.; Fu, J.W.; Chen, Y.H.; Ou, H.S. Ultraviolet-C and vacuum ultraviolet inducing surface degradation of microplastics. Water Res. 2020, 186, 116360. [Google Scholar] [CrossRef]
  119. Wang, L.L.; Kaeppler, A.; Fischer, D.; Simmchen, J. Photocatalytic TiO2 Micromotors for Removal of Microplastics and Suspended Matter. ACS Appl. Mater. Interfaces 2019, 11, 32937–32944. [Google Scholar] [CrossRef]
  120. Venkataramana, C.; Botsa, S.M.; Shyamala, P.; Muralikrishna, R. Photocatalytic degradation of polyethylene plastics by NiAl2O4 spinels-synthesis and characterization. Chemosphere 2021, 265, 129021. [Google Scholar] [CrossRef]
  121. Uheida, A.; Mejía, H.G.; Abdel-Rehim, M.; Hamd, W.; Dutta, J. Visible light photocatalytic degradation of polypropylene microplastics in a continuous water flow system. J. Hazard. Mater. 2021, 406, 124299. [Google Scholar] [CrossRef]
  122. Bensalah, N.; Midassi, S.; Ahmad, M.I.; Bedoui, A. Degradation of hydroxychloroquine by electrochemical advanced oxidation processes. Chem. Eng. J. 2020, 402, 126279. [Google Scholar] [CrossRef] [PubMed]
  123. De Vidales, M.J.M.; Rua, J.; De Juan, J.L.M.; Fernández-Martínez, F.; Dos Santos-García, A.J. Degradation of Contaminants of Emerging Concern by Electrochemical Oxidation: Coupling of Ultraviolet and Ultrasound Radiations. Materials 2020, 13, 5551. [Google Scholar] [CrossRef] [PubMed]
  124. Kiendrebeogo, M.; Estahbanati, M.R.K.; Mostafazadeh, A.K.; Drogui, P.; Tyagi, R.D. Treatment of microplastics in water by anodic oxidation: A case study for polystyrene. Environ. Pollut. 2021, 269, 116168. [Google Scholar] [CrossRef]
  125. Miao, F.; Liu, Y.F.; Gao, M.M.; Yu, X.; Xiao, P.W.; Wang, M.; Wang, S.G.; Wang, X.H. Degradation of polyvinyl chloride microplastics via an electro-Fenton-like system with a TiO2/graphite cathode. J. Hazard. Mater. 2020, 399, 123023. [Google Scholar] [CrossRef]
  126. Bu, J.Q.; Wan, Q.Q.; Deng, Z.W.; Liu, H.; Li, T.H.; Zhou, C.Y.; Zhong, S.A. Waste coal cinder catalyst enhanced electrocatalytic oxidation and persulfate advanced oxidation for the degradation of sulfadiazine. Chemosphere 2022, 303, 134880. [Google Scholar] [CrossRef]
  127. Bu, J.Q.; Wan, Q.Q.; Deng, Z.W.; Liu, H.; Li, T.H.; Zhou, C.Y.; Zhong, S.A. High-efficient degradation of sulfamethazine by electro-enhanced peroxymonosulfate activation with bimetallic modified Mud sphere catalyst. Sep. Purif. Technol. 2022, 292, 120977. [Google Scholar] [CrossRef]
  128. Bu, J.Q.; Deng, Z.W.; Liu, H.; Li, T.H.; Yang, Y.J.; Zhong, S.A. Bimetallic modified halloysite particle electrode enhanced electrocatalytic oxidation for the degradation of sulfanilamide. J. Environ. Manag. 2022, 312, 114975. [Google Scholar] [CrossRef]
  129. Bu, J.Q.; Deng, Z.W.; Liu, H.; Li, T.H.; Yang, Y.J.; Zhong, S.A. The degradation of sulfamilamide wastewater by three-dimensional electrocatalytic oxidation system composed of activated carbon bimetallic particle electrode. J. Clean. Prod. 2021, 324, 129256. [Google Scholar] [CrossRef]
  130. Wan, Q.Q.; Bu, J.Q.; Deng, Z.W.; Liu, H.; Li, T.H.; Luo, T.; Zhou, C.Y.; Zhong, S.A. Peroxymonosulfate activation by bimetallic modified syderolite pellets catalyst for degradation of brominobenzonitrile. Process. Saf. Environ. 2022, 165, 505–513. [Google Scholar] [CrossRef]
  131. Liu, P.; Qian, L.; Wang, H.Y.; Zhan, X.; Lu, K.; Gu, C.; Gao, S.X. New Insights into the Aging Behavior of Microplastics Accelerated by Advanced Oxidation Processes. Environ. Sci. Technol. 2019, 53, 3579–3588. [Google Scholar] [CrossRef] [PubMed]
  132. Xu, Q.X.; Huang, Q.S.; Luo, T.Y.; Wu, R.L.; Wei, W.; Ni, B.J. Coagulation Removal and Photocatalytic Degradation of Microplastics in Urban Waters. Chem. Eng. J. 2021, 416, 129123. [Google Scholar] [CrossRef]
  133. Lapointe, M.; Farner, J.M.; Hernandez, L.M.; Tufenkji, N. Understanding and Improving Microplastic Removal during Water Treatment: Impact of Coagulation and Flocculation. Environ. Sci. Technol. 2020, 54, 8719–8727. [Google Scholar] [CrossRef] [PubMed]
  134. Zhou, G.Y.; Wang, Q.G.; Li, J.; Li, Q.S.; Xu, H.; Ye, Q.; Wang, Y.Q.; Shu, S.H.; Zhang, J. Removal of polystyrene and polyethylene microplastics using PAC and FeCl3 coagulation: Performance and mechanism. Sci. Total Environ. 2021, 752, 141837. [Google Scholar] [CrossRef] [PubMed]
  135. Ma, B.W.; Xue, W.J.; Hu, C.Z.; Liu, H.J.; Qu, J.H.; Li, L.L. Characteristics of microplastic removal via coagulation and ultrafiltration during drinking water treatment. Chem. Eng. J. 2019, 359, 159–167. [Google Scholar] [CrossRef]
  136. Shahi, N.K.; Maeng, M.; Kim, D.; Dockko, S. Removal behavior of microplastics using alum coagulant and its enhancement using polyamine-coated sand. Process. Saf. Environ. 2020, 141, 9–17. [Google Scholar] [CrossRef]
  137. Perren, W.; Wojtasik, A.; Cai, Q. Removal of Microbeads from Wastewater Using Electrocoagulation. ACS Omega 2018, 3, 3357–3364. [Google Scholar] [CrossRef]
  138. Bayo, J.; Olmos, S.; López-Castellanos, J. Microplastics in an urban wastewater treatment plant: The influence of physicochemical parameters and environmental factors. Chemosphere 2020, 238, 124593. [Google Scholar] [CrossRef]
  139. Zhang, Y.S.; Jiang, H.R.; Wan, H.; Wang, C.Q. Separation of hazardous polyvinyl chloride from waste plastics by flotation assisted with surface modification of ammonium persulfate: Process and mechanism. J. Hazard. Mater. 2020, 389, 121918. [Google Scholar] [CrossRef]
  140. Imhof, H.K.; Schmid, J.; Niessner, R.; Ivleva, N.P.; Laforsch, C. A novel, highly efficient method for the separation and quantification of plastic particles in sediments of aquatic environments. Limnol. Oceanogr. Methods 2012, 10, 524–537. [Google Scholar] [CrossRef]
  141. Nguyen, B.; Claveau-Mallet, D.; Hernandez, L.M.; Xu, E.G.; Farner, J.M.; Tufenkji, N. Separation and analysis of microplastics and nanoplastics in complex environmental samples. Acc. Chem. Res. 2019, 52, 858–866. [Google Scholar] [CrossRef]
  142. Talvitie, J.; Mikola, A.; Koistinen, A.; Setala, O. Solutions to microplastic pollution-removal of microplastics from wastewater effluent with advanced wastewater treatment technologies. Water Res. 2017, 123, 401–407. [Google Scholar] [CrossRef]
  143. Enfrin, M.; Dumee, L.F.; Lee, J. Nano/microplastics in water and wastewater treatment processes—Origin, impact and potential solutions. Water Res. 2019, 161, 621–638. [Google Scholar] [CrossRef]
  144. Sun, J.; Dai, X.H.; Wang, Q.L.; van Loosdrecht, M.C.M.; Ni, B.J. Microplastics in wastewater treatment plants: Detection, occurrence and removal. Water Res. 2019, 152, 21–37. [Google Scholar] [CrossRef] [PubMed]
  145. Jiang, H.R.; Zhang, Y.S.; Bian, K.; Wang, C.Q.; Xie, X.; Wang, H.; Zhao, H.L. Is it possible to efficiently and sustainably remove microplastics from sediments using froth flotation? Chem. Eng. J. 2022, 448, 137692. [Google Scholar] [CrossRef]
  146. Mirghorayshi, M.; Zinatizadeh, A.A.; van Loosdrecht, M. Simultaneous biodegradability enhancement and high-efficient nitrogen removal in an innovative single stage anaerobic/anoxic/aerobic hybrid airlift bioreactor (HALBR) for composting leachate treatment: Process modeling and optimization. Chem. Eng. J. 2021, 407, 127019. [Google Scholar] [CrossRef]
  147. Carr, S.A.; Liu, J.; Tesoro, A.G. Transport and fate of microplastic particles in wastewater treatment plants. Water Res. 2016, 91, 174–182. [Google Scholar] [CrossRef]
  148. Hidayaturrahman, H.; Lee, T.G. A study on characteristics of microplastic in wastewater of South Korea: Identification, quantification, and fate of microplastics during treatment process. Mar. Pollut. Bull. 2019, 146, 696–702. [Google Scholar] [CrossRef] [PubMed]
  149. Yang, L.B.; Li, K.X.; Cui, S.; Kang, Y.; An, L.H.; Lei, K. Removal of microplastics in municipal sewage from China’s largest water reclamation plant. Water Res. 2019, 155, 175–181. [Google Scholar] [CrossRef] [PubMed]
  150. Jia, Q.L.; Chen, H.; Zhao, X.; Li, L.; Nie, Y.H.; Ye, J.F. Removal of microplastics by different treatment processes in Shanghai large municipal wastewater treatment plants. Environ. Sci. 2019, 40, 4105–4112. [Google Scholar] [CrossRef]
  151. Jiang, J.H.; Wang, X.W.; Ren, H.Y.; Cao, G.L.; Xie, G.J.; Xing, D.F.; Liu, B.F. Investigation and fate of microplastics in wastewater and sludge filter cake from a wastewater treatment plant in China. Sci. Total Environ. 2020, 746, 141378. [Google Scholar] [CrossRef] [PubMed]
  152. Liu, X.N.; Yuan, W.K.; Di, M.X.; Li, Z.; Wang, J. Transfer and fate of microplastics during the conventional activated sludge process in one wastewater treatment plant of China. Chem. Eng. J. 2019, 362, 176–182. [Google Scholar] [CrossRef]
  153. Fecker, T.; Galaz-Davison, P.; Engelberger, F.; Narui, Y.; Sotomayor, M.; Parra, L.P.; Ramírez-Sarmiento, C.A. Active Site Flexibility as a Hallmark for Efficient PET Degradation by I. sakaiensis PETase. Biophys. J. 2018, 114, 1302–1312. [Google Scholar] [CrossRef] [PubMed]
  154. Joo, S.; Cho, I.J.; Seo, H.; Son, H.F.; Sagong, H.Y.; Shin, T.J.; Choi, S.Y.; Lee, S.Y.; Kim, K.J. Structural Insight into Molecular Mechanism of Poly(ethylene terephthalate) Degradation. Nat. Commun. 2018, 9, 382. [Google Scholar] [CrossRef] [PubMed]
  155. Han, X.; Liu, W.D.; Huang, J.W.; Ma, J.T.; Zheng, Y.Y.; Ko, T.P.; Xu, L.M.; Cheng, Y.S.; Chen, C.C.; Guo, R.T. Structural insight into catalytic mechanism of PET hydrolase. Nat. Commun. 2017, 8, 2106. [Google Scholar] [CrossRef] [PubMed]
  156. Yoshida, S.; Hiraga, K.; Takehana, T.; Taniguchi, I.; Yamaji, H.; Maeda, Y.; Toyohara, K.; Miyamoto, K.; Kimura, Y.; Oda, K. Response to Comment on “A bacterium that degrades and assimilates poly(ethylene terephthalate)”. Science 2016, 353, 6301. [Google Scholar] [CrossRef]
  157. Son, H.F.; Cho, I.J.; Joo, S.; Seo, H.; Sagong, H.Y.; Choi, S.Y.; Lee, S.Y.; Kim, K.J. Rational Protein Engineering of Thermo-Stable PETase from Ideonella sakaiensis for Highly Efficient PET Degradation. ACS Catal. 2019, 9, 3519–3526. [Google Scholar] [CrossRef]
  158. Palm, G.J.; Reisky, L.; Bottcher, D.; Muller, H.; Michels, E.A.P.; Walczak, M.C.; Berndt, L.; Weiss, M.S.; Bornscheuer, U.T.; Weber, G. Structure of the Plastic-degrading Ideonella Sakaiensis MHETase Bound to a Substrate. Nat. Commun. 2019, 10, 1717. [Google Scholar] [CrossRef] [PubMed]
  159. Austin, H.P.; Allen, M.D.; Donohoe, B.S.; Rorrer, N.A.; Kearns, F.L.; Silveira, R.L.; Pollard, B.C.; Dominick, G.; Duman, R.; El Omari, K.; et al. Characterization and engineering of a plastic-degrading aromatic polyesterase. Proc. Natl. Acad. Sci. USA 2018, 115, E4350–E4357. [Google Scholar] [CrossRef]
  160. Liu, B.; He, L.H.; Wang, L.P.; Li, T.; Li, C.C.; Liu, H.Y.; Luo, Y.Z.; Bao, R. Protein crystallography and site-direct mutagenesis analysis of the poly(ethylene terephthalate) hydrolase PETase from ideonella sakaiensis. ChemBioChem 2018, 19, 1471–1475. [Google Scholar] [CrossRef]
  161. Cui, Y.L.; Chen, Y.C.; Liu, X.Y.; Dong, S.J.; Tian, Y.E.; Qiao, Y.X.; Mitra, R.; Han, J.; Li, C.L.; Han, X.; et al. Computational Redesign of a PETase for Plastic Biodegradation under Ambient Condition by the GRAPE Strategy. ACS Catal. 2021, 11, 1340–1350. [Google Scholar] [CrossRef]
  162. Liu, C.C.; Shi, C.; Zhu, S.J.; Wei, R.S.; Yin, C.C. Structural and functional characterization of polyethylene terephthalate hydrolase from ideonella sakaiensis. Biochem. Biophys. Res. Commun. 2019, 508, 289–294. [Google Scholar] [CrossRef] [PubMed]
  163. Li, J.L.; Wang, P.D.; Salam, N.; Li, X.; Ahmad, M.; Tian, Y.; Duan, L.; Huang, L.N.; Xiao, M.; Mou, X.Z.; et al. Unraveling bacteria-mediated degradation of lignin-derived aromatic compounds in a freshwater environment. Sci. Total Environ. 2020, 749, 141236. [Google Scholar] [CrossRef] [PubMed]
  164. Mei, J.F.; Shen, X.B.; Gang, L.P.; Xu, H.J.; Wu, F.F.; Sheng, L.Q. A novel lignin degradation bacteria-Bacillus amyloliquefaciens SL-7 used to degrade straw lignin efficiently. Bioresour. Technol. 2020, 310, 123445. [Google Scholar] [CrossRef] [PubMed]
  165. Sun, C.; Yu, Q.L.; Zhao, Z.Q.; Zhang, Y.B. Syntrophic metabolism of phenol in the anodic degradation within a Phenol-Cr(VI) coupled microbial electrolysis cell. Sci. Total Environ. 2020, 723, 137990. [Google Scholar] [CrossRef] [PubMed]
  166. Zhang, Y.S.; Jiang, H.R.; Bian, K.; Wang, H.; Wang, C.Q. A critical review of control and removal strategies for microplastics from aquatic environments. J. Environ. Chem. Eng. 2021, 9, 105463. [Google Scholar] [CrossRef]
  167. Denaro, R.; Aulenta, F.; Crisafi, F.; Di Pippo, F.; Viggi, C.C.; Matturro, B.; Tomei, P.; Smedile, F.; Martinelli, A.; Di Lisio, V.; et al. Marine hydrocarbon-degrading bacteria breakdown poly(ethylene terephthalate) (PET). Sci. Total Environ. 2020, 749, 141608. [Google Scholar] [CrossRef]
  168. Janssen, P.H.; Yates, P.S.; Grinton, B.E.; Taylor, P.M.; Sait, M. Improved culturability of soil bacteria and isolation in pure culture of novel members of the divisions Acidobacteria, Actinobacteria, Proteobacteria, and Verrucomicrobia. Appl. Environ. Microbiol. 2002, 68, 2391–2396. [Google Scholar] [CrossRef]
  169. Teeraphatpornchai, T.; Nakajima-Kambe, T.; Shigeno-Akutsu, Y.; Nakayama, M.; Nomura, N.; Nakahara, T.; Uchiyama, H. Isolation and characterization of a bacterium that degrades various polyester-based biodegradable plastics. Biotechnol. Lett. 2003, 25, 23–28. [Google Scholar] [CrossRef]
  170. Auta, H.S.; Emenike, C.U.; Fauziah, S.H. Screening of Bacillus strains isolated from mangrove ecosystems in Peninsular Malaysia for microplastic degradation. Environ. Pollut. 2017, 231, 1552–1559. [Google Scholar] [CrossRef]
  171. Yang, J.; Yang, Y.; Wu, W.M.; Zhao, J.; Jiang, L. Evidence of polyethylene biodegradation by bacterial strains from the guts of plastic-eating waxworms. Environ. Sci. Technol. 2014, 48, 13776–13784. [Google Scholar] [CrossRef] [PubMed]
  172. Shah, Z.; Krumholz, L.; Aktas, D.F.; Hasan, F.; Khattak, M.; Shah, A.A. Degradation of polyester polyurethane by a newly isolated soil bacterium, Bacillus subtilis strain MZA-75. Biodegradation 2013, 24, 865–877. [Google Scholar] [CrossRef] [PubMed]
  173. Yoshida, S.; Hiraga, K.; Takehana, T.; Taniguchi, I.; Yamaji, H.; Maeda, Y.; Toyohara, K.; Miyamoto, K.; Kimura, Y.; Oda, K. A bacterium that degrades and assimilates poly(ethylene terephthalate). Science 2016, 351, 1196–1199. [Google Scholar] [CrossRef] [PubMed]
  174. Germain, J.; Raveton, M.; Binet, M.N.; Mouhamadou, B. Screening and metabolic potential of fungal strains isolated from contaminated soil and sediment in the polychlorinated biphenyl degradation. Ecotoxicol. Environ. Saf. 2021, 208, 111703. [Google Scholar] [CrossRef] [PubMed]
  175. Pandey, R.; Choudhury, P.P. Aspergillus niger-mediated degradation of orthosulfamuron in rice soil. Environ. Monit. Assess. 2020, 192, 813. [Google Scholar] [CrossRef] [PubMed]
  176. Sánchez, C. Fungal potential for the degradation of petroleum-based polymers: An overview of macro- and microplastics biodegradation. Biotechnol. Adv. 2020, 40, 107501. [Google Scholar] [CrossRef] [PubMed]
  177. Yamamoto-Tamura, K.; Hoshino, Y.T.; Tsuboi, S.; Huang, C.; Kishimoto-Mo, A.W.; Sameshima-Yamashita, Y.; Kitamoto, H. Fungal community dynamics during degradation of poly(butylene succinate -co- adipate) film in two cultivated soils in Japan. Biosci. Biotechnol. Biochem. 2020, 84, 1077–1087. [Google Scholar] [CrossRef] [PubMed]
  178. Paco, A.; Duarte, K.; da Costa, J.P.; Santos, P.S.M.; Pereira, R.; Pereira, M.E.; Freitas, A.C.; Duarte, A.C.; Rocha-Santos, T.A.P. Biodegradation of polyethylene microplastics by the marine fungus Zalerion maritimum. Sci. Total Environ. 2017, 586, 10–15. [Google Scholar] [CrossRef] [PubMed]
  179. Yamada-Onodera, K.; Mukumoto, H.; Katsuyaya, Y.; Saiganji, A.; Tani, Y. Degradation of polyethylene by a fungus, Penicillium simplicissimum YK. Polym. Degrad. Stabil. 2001, 72, 323–327. [Google Scholar] [CrossRef]
  180. Volke-Sepúlveda, T.; Saucedo-Castañeda, G.; Gutiérrez-Rojas, M.; Manzur, A.; Favela-Torres, E. Thermally treated low density polyethylene biodegradation by Penicillium pinophilum and Aspergillus niger. J. Appl. Polym. Sci. 2002, 83, 305–314. [Google Scholar] [CrossRef]
  181. Devi, R.S.; Kannan, V.R.; Nivas, D.; Kannan, K.; Chandru, S.; Antony, A.R. Biodegradation of HDPE by Aspergillus spp. from marine ecosystem of Gulf of Mannar, India. Mar. Pollut. Bull. 2015, 96, 32–40. [Google Scholar] [CrossRef]
  182. El-Shafei, H.A.; El-Nasser, N.H.A.; Kansoh, A.L.; Ali, A.M. Biodegradation of disposable polyethylene by fungi and Streptomyces species. Polym. Degrad. Stabil. 1998, 62, 361–365. [Google Scholar] [CrossRef]
  183. Dris, R.; Gasperi, J.; Rocher, V.; Saad, M.; Renault, N.; Tassin, B. Microplastic contamination in an urban area: A case study in Greater Paris. Environ. Chem. 2015, 12, 592–599. [Google Scholar] [CrossRef]
  184. Bhattacharya, P.; Lin, S.J.; Turner, J.P.; Ke, P.C. Physical Adsorption of Charged Plastic Nanoparticles Affects Algal Photosynthesis. J. Phys. Chem. C 2010, 114, 16556–16561. [Google Scholar] [CrossRef]
  185. Sanchez-Nieva, J.; Perales, J.A.; Gonzalez-Leal, J.M.; Rojo-Nieto, E. A new analytical technique for the extraction and quantification of microplastics in marine sediments focused on easy implementation and repeatability. Anal. Methods 2017, 9, 6371–6378. [Google Scholar] [CrossRef]
  186. Baruah, S.; Khan, M.; Dutta, J. Perspectives and applications of nanotechnology in water treatment. Environ. Chem. Lett. 2016, 14, 1–14. [Google Scholar] [CrossRef]
  187. Li, M.; Zhao, F.P.; Mika, S.; Meng, Y.; Yin, D.L. Electrochemical degradation of 2-diethylamino-6-methyl-4-hydroxypyrimidine using three-dimensional electrodes reactor with ceramic particle electrodes. Sep. Purif. Technol. 2015, 156, 588–595. [Google Scholar] [CrossRef]
  188. Ariza-Tarazona, M.C.; Villarreal-Chiu, J.F.; Barbieri, V.; Siligardi, C.; Cedillo-González, E.I. New strategy for microplastic degradation: Green photocatalysis using a protein-based porous N-TiO2 semiconductor. Ceram. Int. 2019, 45, 9618–9624. [Google Scholar] [CrossRef]
  189. Xian, G.; Zhang, G.M.; Chang, H.Z.; Zhang, Y.; Zou, Z.G.; Li, X.Y. Heterogeneous activation of persulfate by Co3O4-CeO2 catalyst for diclofenac removal. J. Environ. Manag. 2019, 234, 265–272. [Google Scholar] [CrossRef]
  190. Wei, N.; Zhang, Z.G.; Liu, D.; Wu, Y.; Wang, J.; Wang, Q.H. Coagulation behavior of polyaluminum chloride: Effects of pH and coagulant dosage. Chin. J. Chem. Eng. 2015, 23, 1041–1046. [Google Scholar] [CrossRef]
  191. Wang, K.; Zhang, Y.; Zhong, Y.; Luo, M.; Du, Y.; Wang, L.; Wang, H. Flotation separation of polyethylene terephthalate from waste packaging plastics through ethylene glycol pretreatment assisted by sonication. Waste Manag. 2020, 105, 309–316. [Google Scholar] [CrossRef] [PubMed]
  192. Inoue, Y.; Fukunaga, Y.; Katsumata, H.; Ohji, S.; Hosoyama, A.; Mori, K.; Ando, K. Aerobic degradation of cis-dichloroethene by the marine bacterium Marinobacter salsuginis strain 5N-3. J. Gen. Appl. Microbiol. 2020, 66, 215–219. [Google Scholar] [CrossRef] [PubMed]
  193. Li, Q.Q.; Li, J.B.; Jiang, L.F.; Sun, Y.T.; Luo, C.L.; Zhang, G. Diversity and structure of phenanthrene degrading bacterial communities associated with fungal bioremediation in petroleum contaminated soil. J. Hazard. Mater. 2021, 403, 123895. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Cyclic process of emission, transfer, and decomposition of MPs.
Figure 1. Cyclic process of emission, transfer, and decomposition of MPs.
Sustainability 16 04033 g001
Figure 2. Sources of primary and secondary MPs.
Figure 2. Sources of primary and secondary MPs.
Sustainability 16 04033 g002
Figure 3. MP treatment technologies.
Figure 3. MP treatment technologies.
Sustainability 16 04033 g003
Figure 4. The FESEM images of PS and PE MPs with different alteration degrees. PS-K-5, PSF-5, PE-K-5, and PE-F-5 denote that PS or PE microplastics were altered in heat-activated K2S2O8 and Fenton processes for 5 days. Reprinted with permission from ref. [117]. Copyright 2019, Copyright Xu X.H.
Figure 4. The FESEM images of PS and PE MPs with different alteration degrees. PS-K-5, PSF-5, PE-K-5, and PE-F-5 denote that PS or PE microplastics were altered in heat-activated K2S2O8 and Fenton processes for 5 days. Reprinted with permission from ref. [117]. Copyright 2019, Copyright Xu X.H.
Sustainability 16 04033 g004
Figure 5. Degradation equipment, mechanism, and efficiency diagram of polystyrene MPs. Reprinted with permission from ref. [124]. Copyright 2021, Copyright Tyagi R.D.
Figure 5. Degradation equipment, mechanism, and efficiency diagram of polystyrene MPs. Reprinted with permission from ref. [124]. Copyright 2021, Copyright Tyagi R.D.
Sustainability 16 04033 g005
Figure 6. The proposed degradation process of polyvinyl chloride plastics. Reprinted with permission from ref. [125]. Copyright 2020, Copyright Wang X.H.
Figure 6. The proposed degradation process of polyvinyl chloride plastics. Reprinted with permission from ref. [125]. Copyright 2020, Copyright Wang X.H.
Sustainability 16 04033 g006
Figure 7. Diagram of the removal mechanism of polystyrene (PS) and polyethylene (PE) microplastics by PAC and FeCl3 coagulation. Reprinted with permission from ref. [134]. Copyright 2021, Copyright Zhang J.
Figure 7. Diagram of the removal mechanism of polystyrene (PS) and polyethylene (PE) microplastics by PAC and FeCl3 coagulation. Reprinted with permission from ref. [134]. Copyright 2021, Copyright Zhang J.
Sustainability 16 04033 g007
Figure 8. MP removal from sediments using froth flotation. Reprinted with permission from ref. [145]. Copyright 2022, Copyright Wang H.
Figure 8. MP removal from sediments using froth flotation. Reprinted with permission from ref. [145]. Copyright 2022, Copyright Wang H.
Sustainability 16 04033 g008
Figure 9. The degradation mechanism of PEN and PBT (A); the SEM images of PEN and PBT before and after degradation (B); HPLC chromatogram of the products released from the PEN and PBT films (C). Reprinted with permission from ref. [161]. Copyright 2021, Copyright Han X.
Figure 9. The degradation mechanism of PEN and PBT (A); the SEM images of PEN and PBT before and after degradation (B); HPLC chromatogram of the products released from the PEN and PBT films (C). Reprinted with permission from ref. [161]. Copyright 2021, Copyright Han X.
Sustainability 16 04033 g009
Table 1. Advantages and disadvantages of MP treatment methods.
Table 1. Advantages and disadvantages of MP treatment methods.
MethodAdvantagesDisadvantagesReference
Physical
methods
FiltrationSimple operation, high efficiency, high volumePoor structural stability of adsorbent (membrane) and high cost[183]
Physical
methods
AdsorptionSimple operation, simple equipment, high efficiencyPoor structural stability of adsorbent and the risk of introducing secondary pollution[184]
Physical
methods
ExtractionSimple operation, simple equipment, high safety operationThe cost of this method was high and it was difficult to separate the solute after dissolving it in the extraction solvent[185]
Physical
methods
Magnetic separationLess waste sludge, high efficiency, high volumeSurface agglomeration of magnetic seeds, MPs, and other lipophilic/oleophobic substances[161]
Physical
methods
Oil film separationSimple operation, moderate cost, independent from densityA hydrophobic surface is required, organic contaminant entrainment[110]
Chemical
methods
Fenton oxidationSimple operation, high efficiency, high volumeThe equipment is easily blocked by a large amount of sludge, and the utilization rate of hydrogen peroxide is low[186]
Chemical
methods
Electrochemical oxidationMild reaction conditions, no secondary pollution, flexibility, simple operation, and controllabilityLow mass transfer efficiency and low current efficiency[187]
Chemical
methods
Photocatalytic oxidationSimple operation, simple equipment, high safety operationThe absorption range of photocatalyst is narrow, the utilization rate of light energy is low, some suspended solids and darker chroma will have a great impact, it is difficult to recover, and the electron–hole pair is easy to inactivate[188]
Chemical
methods
Persulfate advanced oxidationSimple operation, simple equipment, high safety operation, low costThe pH requirement is high, the catalyst recovery is difficult, and the generated redox potential of hydroxide is not as high as that of sulfate[189]
Chemical
methods
CoagulationSimple operation, simple equipment, short treatment timepH has a great influence on this method, and many coagulants have reducibility and color. If the dosage is large, it can easily cause high chroma and low removal rate[190]
Chemical
methods
ElectrocoagulationNo requirement for chemical coagulants, less sludgeHigh electricity and conductivity are required, and oxidation is caused by electrodes[132]
Chemical
methods
Foam flotationSimple operation, simple equipment, high safety operation, low costThe repeatability of the experiment was very poor, and the temperature has a great influence[191]
Biological
methods
Anaerobic–anoxic–aerobic activated sludgeLow cost, high volume, simple operationTime-consuming, low removal efficiency, easy death of bacteria, and a large amount of sludge[192]
Biological
methods
Enzymatic degradationHigh efficiency and specificityComplex operation process, high cost, and harsh reaction conditions[175]
Biological
methods
Bacterial degradationHigh safety, low cost, and simple operationComplex bacterial culture, high environmental requirements for bacterial degradation, high time consumption[193]
Biological
methods
Fungal degradationHigh safety, low cost, and simple operationComplex bacterial culture, high environmental requirements for fungi degradation, high time consumption[172]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Y.; Chen, P.; Tang, Y.; Yang, Y.; Zhou, C.; Bu, J.; Zhong, S. Microplastics in Water: A Review of Characterization and Removal Methods. Sustainability 2024, 16, 4033. https://doi.org/10.3390/su16104033

AMA Style

Li Y, Chen P, Tang Y, Yang Y, Zhou C, Bu J, Zhong S. Microplastics in Water: A Review of Characterization and Removal Methods. Sustainability. 2024; 16(10):4033. https://doi.org/10.3390/su16104033

Chicago/Turabian Style

Li, Yun, Ping Chen, Yalan Tang, Yanjing Yang, Chengyun Zhou, Jiaqi Bu, and Shian Zhong. 2024. "Microplastics in Water: A Review of Characterization and Removal Methods" Sustainability 16, no. 10: 4033. https://doi.org/10.3390/su16104033

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop