Next Article in Journal
Evolution and Future Prospects of Education Evaluation Research in China over the Last Decade
Previous Article in Journal
Young Adults’ Intentions toward the Prevention of Microplastic Pollution in Taiwan: Examining Personality and Information Processing in Fear-Appeal Communication
Previous Article in Special Issue
Comparative Assessment of Microplastics in Surface Water and Sediments of Meishe River, Haikou, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of Micro(Nano)Plastics in the Environment: Sampling, Identification, Risk Assessment and Control

1
Key Laboratory of Agro-Forestry Environmental Processes and Ecological Regulation of Hainan Province, Haikou 570228, China
2
College of Ecology and Environment, Hainan University, Haikou 570228, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Sustainability 2022, 14(21), 14338; https://doi.org/10.3390/su142114338
Submission received: 13 September 2022 / Revised: 25 October 2022 / Accepted: 26 October 2022 / Published: 2 November 2022

Abstract

:
Advances in urban infrastructure, a flourishing polymer sector, and more traffic have all contributed to a rise in micro(nano)plastics in the environment. Researchers are exploring the production, fate, toxicity threshold, and severity of micro(nano)plastic exposure. Albeit, understanding sampling protocols, preservation of samples, and characterization of micro(nano)plastics obtained from the different mediums (e.g., soil, water, air, and living bodies) is still challenging. Particularly identification of micro(nano)plastics, on the other hand, is restricted and limited to the typical generic definition of contaminating sources. In addition, before micro(nano)plastics degrade naturally, many challenges must be overcome, enhancing the need for research on assisted degradation. Thus, a systematic review is presented, which begins by discussing micro(nano)plastic identification, sampling, and handling; then showcases the environmental and health consequences and how to control them; finally, it discusses environmental micro(nano)plastics management options. According to studies, biological and chemical methods to break down micro(nano)plastics have risen in popularity. However, these methods often only cover one type of plastic. Furthermore, these solutions can transform polymers into micro(nano)plastics and may also produce byproducts, increasing environmental contamination risk. Therefore, control, prevention, and management strategies are all investigated to generate more realistic and long-term solutions. The literature suggests a combination of different microorganisms (e.g., different bacterial species) and different approaches (e.g., filtration with degradation) could be more effective in the treatment of micro(nano)plastics. Furthermore, according to the literature, relevant health risks associated with micro(nano)plastics to humans from various exposure routes are currently unclear. Likewise, standardization of methods supported with sophisticated state-of-the-art apparatus for detecting micro(nano)plastics is required. Overall, precision in micro(nano)plastic identification and treatment strategy selection is critical, and their usage should be regulated if their environmental behavior is not properly addressed.

1. Introduction

Plastics have been massively produced since the 1940s, when society started to consume large quantities of plastics in every segment of our lives, including industry, agriculture, healthcare, and others [1,2]. The production of plastics has been expanded in the last decades because of their favored characteristics, such as lightness, durability, flexibility, versatility, and cost-effectiveness [3]. However, the extensive consumption of plastics has resulted in severe environmental consequences. It is of particular relevance that once plastic litters enter the environment, they can persist for a long time and will be broken down into small debris (alias mesoplastics), microplastics (MPs, <5 mm), and even nanoplastics (NPs, <100 nm) via physical, chemical and biological processes before reaching complete decomposition [4]. The MPs and NPs have attracted public concern due to their small size, worldwide distribution, and potential ecotoxicological effects [5]. As a result, the studies have been increasing in the past few years (Figure 1, retrieval time from 01.2016 to 01.2022). The figure shows that the studies of MPs and NPs have expanded in the past few years.
Figure 1. Visual analysis diagram of research studies relating to MPs and NPs since 2016. To date, many previous studies have investigated the global distribution of MPs and recently focused on the NPs [6]. For example, as compared with the studies of MPs, the analysis methodologies of NPs are pretty limited, and only 405 publications (including 71 review papers and 334 research papers) can be retrieved according to the literature retrieval of ‘nanoplastic(s)’with other keywords such as ‘analysis’, ‘extraction’, ‘methodology’, ‘separation’, ‘characterization’, ‘identification’, or ‘quantification’ (Figure 2) (Date up to 1 May 2022). The figure shows that most NPs studies focused on the fate and toxicity in the environment, while studies regarding the measuring methodology are still scarce.
Figure 1. Visual analysis diagram of research studies relating to MPs and NPs since 2016. To date, many previous studies have investigated the global distribution of MPs and recently focused on the NPs [6]. For example, as compared with the studies of MPs, the analysis methodologies of NPs are pretty limited, and only 405 publications (including 71 review papers and 334 research papers) can be retrieved according to the literature retrieval of ‘nanoplastic(s)’with other keywords such as ‘analysis’, ‘extraction’, ‘methodology’, ‘separation’, ‘characterization’, ‘identification’, or ‘quantification’ (Figure 2) (Date up to 1 May 2022). The figure shows that most NPs studies focused on the fate and toxicity in the environment, while studies regarding the measuring methodology are still scarce.
Sustainability 14 14338 g001
Figure 2. Visual analysis diagram of MPs and NPs measurement since 2016. Note: The retrieval was accomplished with the keywords such as “nanoplastic”, “analysis”, “extraction”, “methodology”, “separation”, “characterization”, “identification”, and “quantification”, via database, including Web of Science, Science Direct.
Figure 2. Visual analysis diagram of MPs and NPs measurement since 2016. Note: The retrieval was accomplished with the keywords such as “nanoplastic”, “analysis”, “extraction”, “methodology”, “separation”, “characterization”, “identification”, and “quantification”, via database, including Web of Science, Science Direct.
Sustainability 14 14338 g002
Recently, Cai et al. [7] summarized the progress and challenges of environmental NPs analysis, in which 95 studies have been reviewed while only 12 of them focused on the field studies (Date up to 1 May 2022). For example, Fang et al. [8] introduced the identification and visualization of MPs and NPs by a newly developed Raman mapping image method; Peiponen et al. [9] also demonstrated the optical identification of MPs and NPs in aquatic environments. However, they all pointed out that improving analytical reliability and accuracy is still challenging, especially for NPs. For instance, how to enhance the weak Raman signal from NPs, and the identification requires a combination of optical detection, etc. In other words, it is a common challenge that the analytic techniques remain to be standardized and explored, especially for improving the detection efficiency and accuracy.
Various physical, biological/microbial, and chemical techniques have been implemented to combat the threat posed by micro(nano)plastic. Currently, some research is being carried out on (micro)plastic degradation. The degradation of (micro)plastics is split into biological and chemical degradation [10]. Sudhakar et al. [11] conducted in situ experiments using polyethylene (PE) and polypropylene (PP) sheets in marine water. They studied Pseudomonas spp., Clostridium spp., unidentified anaerobic, heterotrophic, and iron-reducing bacteria and fungi and found that degradation dependent on study site, plastic type, and season. Another study addressed an in vitro study of Nylon pellets degradation by Bacillus cereus, B. sphericus, Vibrio furnisii, and Brevundimonas vesicularis and found that degradation varied across microbial taxa (highest for B. cereus) [12]. On the other hand, in vitro degradation of PE sheets with B. cereus subgroup A and B. sphericus GC subgroup IV showed degradation rates varied with PE [13]. Bao and coworkers recently studied microplastics released from biodegradable blended plastic and demonstrated bacterial communities in contrasting environments such as water and air, showing variable trends [14]. For instance, plastics having hydrolyzable covalent links (such as polyethylene terephthalate (PET)) can be hydrophilic, which accelerates the hydrolytic degradation of the polymer [14]. Microplastics’ C–C and C–H bonds can be destroyed by ultraviolet (UV) light, causing them to deteriorate [15,16]. However, more research is needed on the chemical breakdown of microplastics, which is currently being carried out via advanced oxidation processes (AOPs) [17].
This work intends to provide detailed insight into recent advancements in micro and nano plastics characteristics, including sampling, identification, and handling. Furthermore, different environmental segments, such as seawater, soil, and sediments, were assessed for micro and nano plastics. Studies comprised plastics detection in birds, fish, and other organisms, and their characteristics inside living bodies were also discussed. The emphasis will be on risk assessment and management of micro and nano plastic pollution leading to their control and degradation strategies.

2. Overall Introduction on Techniques for MPs and NPs Measurement

The studies of MPs and NPs pollution are gaining more and more attention around the globe; however, it should be cautioned that their protocols remain to be standardized profoundly in terms of sampling, preservation, digestion, separation, and identification in particular to some comprehensive, fast and efficient approaches [18]. At present, most of the studies adopted the approaches recommended by NOAA. Figure 3 summarizes the most common methods for the determination of MPs in water [19,20], soil (or sediment, sand) [21,22], organism [23,24], and air [25] in the previous studies

2.1. Separation Methods for MPs and NPs in Different Matrices

The analytical steps for MPs and NPs in different environmental matrices are generally similar, including sample collection, processing (e.g., centrifugation, digestion, staining), flotation, filtration, and then a series of physical and chemical methods for characterization (Figure 3). In general, trawl or Neuston nets are usually used for MPs sampling from the aqueous environment such as freshwater and seawater. At the same time, stainless steel products (e.g., shovels and spoons) are commonly applied in sampling soil, sediment, and sludge. By comparison, pumps and atmospheric wet and dry deposition (e.g., active and passive sampling) are frequently applied to sampling MPs in the air; and planktonic nets are used to collect aquatic organisms [26,27,28]. After sampling, flotation is essential in separating MPs from the collected samples. At this stage, the large materials will be removed using saturated sodium chloride and sodium iodide [29,30]. Sodium chloride applies to the flotation of low-density plastics, while sodium iodide is for the flotation of high-density plastics. After flotation, the supernatant will be filtered using filter membranes made of cellulose nitrate, nylon, PTFE, sartorius, polycarbonate, and glass membranes [31,32,33]. However, digestion is essential to the flotation of MPs in complex samples (e.g., sludge, wastewater), and a 30% hydrogen peroxide solution is commonly used for digestion [34,35]. In addition, the digestion solution varies with the types of samples and the characteristics of MPs, aiming to achieve maximum recovery efficiency [36]. For example, potassium hydroxide solution is generally used for the digestion of biological samples [37], while proteinase K is for tissues [38] and Fenton’s reagent for sediments [39].

2.2. Identification of MPs and NPs

Various methods, including visual, spectroscopic, and scanning electron microscopy (SEM), are recommended to identify the MPs in environmental samples. Among them, the microscope’s visual method is aimed at observing MPs’ surface morphology (e.g., shape, length, color, etc.) [40], and high-resolution images of surface structure MPs can be achieved using SEM. SEM pictures can be used to compare MPs before and after aging or to see MP biofilm [41]. Further, the spectroscopic method, mainly through Micro Fourier Transform Interferometer (μ-FTIR) or μ-Raman can identify the composition of polymers by analyzing the functional groups on MPs [42], in which the μ-FTIR is applicable to the identification of MPs in size of ca. 20 μm, while μ-Raman to those in size of ca. 1 μm and even smaller [43]. However, since both of these procedures are costly and time-consuming, it is useful to quickly establish techniques for detecting MPs. Next, some new approaches were suggested, such as a support vector machine (SVM) approach is recommended to determine MPs based on hyperspectral imaging and achieving steady detection of various kinds of plastics [18]. Some studies investigated a portable optical sensor to detect transparent and translucent MPs in freshwater. For understanding a variety of MPs and plastic additives, gas chromatography-mass spectrometry (GC-MS) and liquid chromatography-mass spectrometry (LC-MS) are recommended in addition to visual and spectral verification. For instance, LC-MS/GC-MS can be used to examine the degradation products and additives in natural and manufactured fibers [44]. In addition, other integrated methods such as thermal desorption-proton transfer reaction-mass spectrometry (TS-PTR-MS), thermochemolysis coupled to GC-MS, pyrolysis-gas chromatography/mass spectrometry (py-GC-MS) and surface-enhanced Raman spectroscopy (SERS) have been shown to be efficient in the identification of NPs in real-world studies [45]. However, exploring novel methods for rapid quantitative and qualitative identification of MPs in various media, particularly for tiny plastic particles such as NPs, remains challenging [46].

3. Techniques for MPs and NPs Measurement in Various Matrices

3.1. Techniques for MPs Studies in Seawater

As shown in Figure 3 and Table 1, the main steps for MPs studies in seawater include sampling, preservation, digestion, filtration, and identification.

3.1.1. Sampling

Seawater samples can be collected by plankton net [61], manta trawl [47], catamaran, neuston net [48] (Figure 4), teflon pump [49] and intake system of the vessel [51], with a common range of mesh size from 120~335 μm [47,54]. Manta trawl and Neuston net with a mesh size of 330 μm or 333 μm [62,63] is the most frequently used in seawater sampling. In general, the vessel is dragged at a speed of 1~5 knots [50,54] for 5~30 min [61,64]. Next, a flow meter is installed at the entrance of the sampling device to measure the water flow [52]. Finally, transfer the sampled surface water to a glass bottle for further analysis.

3.1.2. Preservation and Digestion

In general, the aforementioned seawater samples are directly used for detecting MPs [19,33], except for some specific purpose of investigating MPs distribution in the environment [56]. Under such circumstances, it is suggested to preserve fresh seawater with reagents (e.g., 4% or 5% formalin solution) [52], 70 or 95% ethanol [53], or directly refrigerated at 4 °C [65], since that some MPs may transfer from seawaters to phytoplankton, zooplankton and other organisms in higher trophic levels. Subsequently, seawater is digested to degrade the organic matter inside by adding reagents such as 5 M NaOH [66], 65% HNO3 [56], 30–35% H2O2 [20], or wet peroxide oxidation (WPO, a combination of 30% H2O2 and 0.05 M Fe(II) solution (i.e., 7.5 g of FeSO4·7H2O to 500 mL of water and 3 mL of concentrated sulfuric acid)) [53] and others, as discussed later in Section 3.2. Next, the digested solution of the samples on the net will be transferred to a density separator for flotation, which separates the light component MPs (with a density of 0.8–1.4 g/mL) from heavy ones and collects the floating solids using 0.3-mm custom sieve [57]. In contrast, the samples in the volume-reduced water can be directly separated from the mixtures through sieving [53] and filtration with a set of vacuum suction filter devices with membrane pore sizes of 0.45~0.75 μm [54]. Next, the filter paper is subsequently transferred into a petri dish with a cover for drying at 25~60 °C [53].

3.1.3. Observation, Enumeration, and Identification

The items similar to MPs (based on morphology and color) will be selected from the above-dried samples for further analysis. In general, MPs can be analyzed by qualitative analysis and quantitative analysis, in which the qualitative analysis applies to the analysis of the physical morphologies and the identification of the chemical components of MPs [67], while the quantitative analysis is mainly to calculate the quantity and weight of MPs by inspection [68]. First, the MPs’ shape, color, and size were observed by a stereoscopic microscope [55] and characterized by SEM [69]. Next, the suspected MPs particles or fibers can be selected using tweezers. The quantity of MPs is counted by manual counting (prone to the miscarriage of justice) [68]. At the same time, the qualitative analysis is achieved by μ-FTIR [61], μ-Raman [60] are commonly used in previous studies, other approaches such as SEM-EDS [53], ESEM-EDS [67], Pyr-GG-MS [70], or other methods such as tagging method [71].

3.2. Techniques for MPs Studies in Sediment and Soil

Among the types of sediments, such as sea (core) sand, beach sand, and municipal soil [34,38], beach sand is the most studied type. Beach sands are collected by stainless steel spoon, shovel, or frame with a quadrat from the surface to a depth of 1~5 cm [59,72] (As shown in Table 2). Sands are usually preserved in aluminum foil or stainless containers and then dried through freeze-dried [50], air-dried [73], or oven at 40–60 °C [74,75] [76]. Once dried, the samples can be sieved with a mesh size of 1~5 mm [75,77].
Next, the dried and sieved samples will be separated by density flotation [79], which separates the light component MPs (with a density of 0.8–1.4 g/mL) from the reconstituted sediment (with a density of 2.65 g/mL) [68]. In detail, a certain volume of sediment samples can be mixed evenly by stirring and shaking, letting it settle down until reaching solid-liquid separation, and collecting the MPs suspended or floating on the supernatant using filter paper.
As summarized in Table 3, the common saturated salt solution includes 1.2 g/mL NaCl [75], 1.71 g/mL ZnBr2 [81], 1.6 g/mL ZnCl2 [82], and 1.57 g/mL NaI [77], and others. NaCl is the most widely used because of its low cost and no harm to the environment [79]. However, some high-density plastic particles cannot float in the solution. Flotation reagents such as NaI, KI [83], ZnCl2, NaBr [84], and ZnBr2 [81] are also used to float denser polymers in some studies, but they are costly and may cause chemical hazards to filter paper and therefore disturb the identification of samples, and be a hazard to the environment.
After floatation, the supernatant is filtered with a mesh size of 0.47~2 μm through vacuum suction filter equipment [72,86]. More MPs would be collected by the use of a 0.7 μm or smaller filter [50] as compared with larger ones (e.g., 2 μm or larger mesh size) [76]. At last, the suspected items similar to MPs can be identified according to the same procedures as the seawater sample [87].

3.3. Techniques for MPs Studies in Organisms

The existence of MPs in marine organisms have been studied in birds [88], fish [55], mussel [89], terrestrial invertebrate [90], earthworms [91], bivalves [92], and invertebrates [93], in terms of the whole animal [93], digestive tract [88], all soft tissue [92], and gastrointestinal tract [94]. Two sampling devices are commonly used to collect organism samples in water [95]. One is the direct use of plankton nets such as Bongo nets and shrimp nets, the residual organisms on the net can be collected after being washed with 10% formalin solution. The other is that collect a certain volume of the water column with a clean container and then extract the organism from the water column for further analysis.
Preservation methods used in biological organisms are stricter than seawater or sediment samples. Immediately, frozen the organism samples at −20 °C [96] or preserved them in formalin solution [97] or ethanol [98] (Table 4). The pretreatment of organism samples is a combination of digestion and filtration. According to the size of the biological sample, digestion can be divided into two types. One is the complete digestion (i.e., digestion of the whole biological sample [93], which applies to small organisms such as mussels [89] and oysters [99]. The other is partial digestion dissolves part of the soft tissue or digestive organs obtained by dissection and tissue section [100], which applies to larger biological samples such as fish [101], waterfowl [102], and others (e.g., turtles) [103].
Digestion is classified as acid, alkaline, and enzyme digestion [67] (Table 5). Based on the previous studies, acid digestion solutions mainly include H2O2 (30% or 35%) [47], HNO3 (65% or 69%) [56], NaClO [66] and other combined acids (i.e., 5% or 37% HCl, 65% or 69% HNO3, 40% HF) [66,104] at room temperature (usually 25 °C) [104], 65 °C [106], and 100 °C [76], in which 65 °C is the most frequently used. They show different advantages and disadvantages. For example, H2O2 has no effects on plastic particles, but it cannot remove all organic materials; NaClO can completely eliminate organisms, but it fails to reach an optimum digestion efficiency in a short time [66], and the concentrated acid, especially corrosive acid may adversely impair the structural and chemical integrity of MPs and may result in strong chemical hazards to the environment as well [56,66,76]. By comparison, enzymatic digestion has high efficiency, but it may destroy nylon fiber plastics and cannot completely digest biogenic materials [92,96]. Once the biological tissues are completely digested, a vacuum suction filter device is further used to filter the digested mixtures, and the identification of MPs follows the procedures of seawater samples.

3.4. Research Methods of MPs in Air

Air is a mixture of gases and suspended particles [107]. Recent literature suggests the presence of MPs in the air [28]. There are quite limited studies on the occurrence of MPs in the air [108,109,110]. In general, atmospheric fallout with dry and wet deposition can be collected through a stainless-steel funnel, which connects with a 20 L glass bottle at the bottom [111,112]. Once the fallout was collected, rinse the funnel with reverse osmosed water to recover all particles (including MPs) adhering to the funnel. After being rinsed, all samples in a 20 L glass bottle were immediately filtered on quartz fiber GF/A (with a size of 1.6 μm). At last, observe the MPs on the filter using a stereomicroscope and measure the length of the fibers during the enumeration (e.g., the software ‘Histolab (Microvision instruments—Evry—France)’ coupled with the stereomicroscope) [111,112]. The other method for indoor air sampling is as follows, a pump (Stand-alone sampling pump GH300, Deltanova, France) at 8 L/min allowed to sample indoor air on quartz fiber GF/A (1.6 μm, 47 mm), and the height and time of samplings depends on the location selected. Next, 50 mL zinc chloride solution (1.6 g/cm3) is used as a flotation solution to separate the 5.5 mg sample in a separation funnel. Next, a subsample of 1 mL was taken from the homogenized floating fraction and filtered on quartz fiber GF/A (1.6 μm, 47 mm). At last, all samples were observed by stereomicroscope and qualitatively analyzed by using μ-FTIR or μ-Raman [111].

3.5. Measurement of MPs in Other System

In addition to the natural environmental media, there are some techniques for detecting the MPs in other specific media, such as wastewater [65], drinking water or tap water [113,114], and sludge [115]. Among them, grabbing samples (i.e., single samples collected at one time) is the most typical method for collecting wastewater samples from wastewater treatment plants [43]. In detail, a telescopic sampling rod (Telescoop, Waterra Ltd., Solihull, UK) was used to collect wastewater samples, which can be further transferred into a 50 mL centrifuge tube for centrifugation at 2038× g for 2 min, and then the resulting pellet was treated with 30 mL of Fenton’s reagent [61]. By comparison, the sampling and pretreatment of tap water and drinking water is much easier because their components are simpler than wastewater. The samples are mainly taken from bottled water and water plants [116,117]. Next, the samples will be processed in terms of staining [118], filtration [119], or digestion [120] for subsequent observation. Since the sludge in wastewater treatment plants commonly contains complex components, they are usually to be air-dried, sieved, and kept as soil samples [34] and then analyzed using the same methods as that of MPs analysis in soil.

4. Accumulation and Translocation of MPs in the Human Body

MPs can transport and deposit in tissues and organs when breathed or inhaled [28,121]. The non-absorbed component of MPs is excreted out [122]. Few MPs have been known to infiltrate the respiratory tract. The different densities and configurations determine MPs’ aerodynamics, which is used to estimate MPs settling velocities [123]. Small aerodynamic equivalent diameter MPs are more ready to attain the deeper respiratory track. Fibers were found in the lungs compared to other shapes, suggesting their high risk [124]. Mehmood and Peng [28] proposed that the persistence, transportation, and health risks posed to MPs are directly related to their types, sizes, shapes, and environment, such as indoor and outdoor locations. After 180 days in synthetic lung fluid, PP and PE fibers showed no disintegration or alterations, indicating a high potential for MPs persistence in the respiratory system [125]. Some nanoscale particles have been demonstrated to pass through the various protecting barriers, cell membranes, blood, brain, and placenta barriers [126,127]. Nevertheless, there is no direct proof that MPs are distributed and accumulated in human organs. According to the sole mouse-model-based study, MPs can be collected in different organs, including the kidney, gut, and liver [128].
Little information is available on the possibility of MPs intake and the accompanying health effects in humans. However, the harmful effects of MPs have been observed in other living organisms, including animal studies, but their findings cannot be extrapolated to human toxicity. Ingestion of MPs showed inflammation in the digestion system in Mytilus [129]. Fish’s immune systems are similarly vulnerable to micro(nano)plastic attacks. For example, Fathead minnow showed a substantial increase in primary granule degranulation and innate immune system release neutrophil extracellular when exposed to nano-plastics [130]. Polystyrene microspheres were administered intramuscularly into rats to increase blood coagulability or to cause vascular occlusions, resulting in pulmonary hypertension, inflammation, and chemotactic activity [131]. Intracellular structure damage was observed in vivo polystyrene exposure in macrophages, erythrocytes, and rat alveolar epithelial cells [132]. Microbes, pathogens, metals, persistent organic contaminants, and chemicals are among the pollutants in soil, water, and air [133]. When MPs coexist with such contaminants, MPs can absorb these pollutants, which can then be released into the human body, exacerbating MPs toxicity [134,135]. The consequences of MPs on human health are still being debated [121]. Since plants can also uptake several types of soil pollutants, which ultimately enter the food chain [136], MPs has also potential to accumulate in plant biomass [137]. Some researchers underlined the concerns of food chain transfer, while others argued that MPs or MPs additions have no negative effects [134]. Furthermore, the deposition of suspended particles in the air on plants, whether dry or wet, can disrupt photosynthesis and be intake by humans and other animals through the consumption of such polluted plants. The majority of the disagreements arise from the ambiguity in the estimation intake of MPs, so further research and modeling are required for a more comprehensive demonstration of intake routes, fate, and persistence.
Despite the fact that MPs toxicity is still in its early stages, occupational illnesses have been related to inhaling MPs particles. Workers exposed to PP have respiratory symptoms that are 3.6 times more severe [124]. Long-term inhalation and excessive exposure to fine particles is more dangerous to health and also result in gene mutation [138]. Exposure to PP and other synthetic plastics over 10 to 20 years enhanced the number of cancer patients among textile workers. Workers who worked with polyvinyl chloride had a higher risk of lung cancer as they got older, worked more years, and spent more time in the factories [123]. More research is needed to estimate MPs concentrations in the atmosphere and tissues and understand their mechanical toxicity. In additions, the toxicity associated with micro(nano)plastics should further studies based on different age groups and genders. Previous studies suggest children are more susceptible to MPs exposure compared to adults due to their activities such as playing on artificial turf in grounds and parks [28]. Likewise, MPs may pose a threat to fetus development thus their toxicity to women during pregnancy should also explore in detail [139,140].

5. Management and Treatment of Plastic Pollution

5.1. Eco-Friendly Safety Procedures against the MPs

Due to higher sales and market share in the international market and a wide range of use in daily life, plastic appears as an unavoidable commodity [141,142,143]. A very small share of the total produced plastic is collected for recycling, so plastic waste is continuously gathered in the environment due to its high stability and persistence [142]. Although the global community has aimed directly at a range of regulatory initiatives in recent years to regulate plastic pollution and its environmental damage, there are presently restrictions on the manufacturing and using plastics.
Some nations worldwide have implemented regulations limiting the manufacturing of plastics (e.g., beverage bottles and carrier bags). For example, in the Canadian province of Ontario, a law prohibiting the manufacture of microbeads was passed [144]. As a result, legislation has been passed in the United States prohibiting the use of small plastics in order to reduce pollution [145]. These approaches, taken together, aid in lowering the micro(nano)plastic existence on the ground, in the steam, and the air.
The manufacturing industry also performs an integral role in decreasing micro(nano)plastic in the overall supply chain. IKEA (Ingvar Kamprad Elmtaryd and Agunnaryd), for instance, has included the EPR (Extended Manufacturer) strategy in its marketing strategy by encouraging the reusing of polymeric products across its production network [146]. The growth of a circular financial system, such as recycling, reuse, and waste management process, may also assist in controlling plastic utilization, with significant zigzag social and ecological consequences. Among the most effective and sustainable solutions to avoid pollution has been recommended: regulating micro(nano)plastic in a well. For instance, outlawing plastic bags can successfully prevent excess plastic use, and pursuing that goal can minimize plastic contamination with MP [147]. Other producers must also label the plastics commodities with proper guidance and information regarding corresponding recycling and environmental feasibility. Furthermore, campaigns raise awareness of MPs challenges among colleges, schools, organizations, and networks, as well as teach individual responsibility for decreasing MPs by deciding to cast-off, limit, recycle, and reuse. Children’s educational initiatives (e.g., sea waste education) will increase their understanding, visions of outcomes, and willingness to report them [148].

5.2. Governmental Policies and Nongovernmental Campaigns to Curb MPs Pollution

In 1970s, the distribution of plastic wastes in the Pacific Ocean firstly attracted the attention of researchers [149], and it was more than two decades later when Thompson coined the term “microplastic” in 2004 [4]. Micro(nano)plastic pollution has attracted public concerns in recent years. For example, MP was detected in some brands of table salts in China [150,151] and Turkey [152] and even in 83% of all tap water samples collected from cities and towns from five continents [153]. Under such circumstances, some policies and campaigns relating to plastic litter have been enacted or launched, aiming to control the quantities of plastic litter from the human side. Some selected policies or campaigns curbing MP pollution are shown with the time axis in Figure 5.
In the past few years, more and more countries made all kinds of efforts to slow down the release of plastic wastes (including MPs and NPs) into the natural environment, in terms of banning the supplement of pellets in daily skincare products, limiting the use of disposable plastic products, or strengthening the monitoring and remediation of plastics pollution. For example, some provinces in China banned the production, supplement, sale, and use of pellets, disposable non-degradable plastic products in the past years [27], and the European Commission held a side event as a response to Resolution UNEP/EA.4/R.9 on “Addressing Single-use Plastic Products Pollution” in 2021. It is expected that the distribution and transfer of (micro-/nano-)plastics pollution can be controlled with the efforts of all parties.

5.3. Biological Control

Researchers have found biological management viruses positively impact micro(nano)plastic degradation. Biodegradation is one of the most important strategies to control micro(nano)plastic. Plastics are decomposed by microbes, fungi, and meal worms. It has already been discovered as a dependable and environmentally beneficial solution for dealing with plastic contamination by decomposing it without causing any negative consequences. Bombelli et al. [154] found that ethylene glycol produced by wax moth larvae Galleria mellonella (at a rate of 0.23 mg cm−2 per hour) can rapidly bio-degrade PE.
Similarly, microorganisms of the earth’s crust (Lumbricus Terrestris) degrade low-density PE (LDPE) MP particles [155]. MPs’ hazardous has also been found to be declined by the Bacillus gottheilii bacteria. Bacillus gottheilii and Bacillus cereus were utilized by Auta et al. [156] to eliminate MPs from PE, PP, PET, and polystyrene. For morphological and structural alterations and weight loss in MPs, the findings were examined using (FTIR).
B. cereus and B. gottheilii presented a higher reduction of polystyrene and polyethylene and showed removal rates constant of 0.0019 and 0.0016 per day, respectively, while the corroding half-life was 363.16 and 431.25 days, indicating that B. gottheilii has the potential to degrade micro(nano)plastic. These plastic predators have demonstrated a novel approach to reducing plastic trash and MPs [156]. In other studies, Bacillus siamensis demonstrated 8.46% degradation of LDPE in 90 days [157], 5.5% of HDPE in100 days by Aspergillus flavus [158], 14.7% of PE in 60 days by Bacillus sp. In addition, Paenibacillus sp. [159]. Similarly, Aspergillus nomius RH06 degraded 6.63, while Trichoderma viride RH03 showed 5.13% of LDPE in 45 days [160]. Comparatively, Bacillus cereus strain A5 and Aspergillus oryzae strain A5 showed 35.72 and 36.4% degradation of LDPE in 112 days [161]. Previously, Thomas et al. [162] reported 18.0% degradation of PE by Pseudomonas fluorescens in 270 days and Klebsiella pneumoniae CH001 showed 18.4% degradation of HDPE in 60 days [143]. At the same time, the highest plastic degradation efficiency was found by Bacillus vallismortis bt-dsce01, which degraded 75% of LDPE in 120 days [163].
As a result, governments must invest more in research to uncover species capable of adequately breaking down plastic. Destruction of MPs and their associated species is a promising way to taint micro(nano)plastic’s reputation as a green method.

6. Methods and Mechanisms of Degradation of (Micro)Plastics

The following section discusses the most popular micro(nano)plastic degradation approaches, i.e., (1) biodegradation; (2) chemical degradation. Biodegradation includes bacterial-assisted biodegradation, fungal-assisted biodegradation, algal-assisted biodegradation, Integrated enzymatic biodegradation, and combined biodegradation [133].

6.1. Biodegradation

Biodegradation is the disintegration of pollutants (e.g., plastics) by microbes such as fungi, bacteria, and viruses accompanied by microbial catabolism [133]. The development of biofilms on micro(nano)plastic and plastic surfaces leads to the breakdown of their fundamental structure frame and cross-links by particular enzymes, resulting in monomers, oligomers, dimers, and other depolymerization byproducts (e.g., CO2 and H2O) [164]. Inorganic and organic metabolites or conversion constituents result from incomplete mineralization [165]. In an alkaline environment, biodegraded polymers produce H2O and CO2 [166], while in anaerobic conditions, they are CH4 and CO2 [167]. Plastic biodegradation has gained widespread acceptability due to its environmental neutrality and convenience. Aside from ecological considerations, the variety of microbes, the nature of plastics, the shape of polymers, and the enzyme selection are all major candidates influencing the biodegradation of plastic. Biodegrading by bacteria, enzymes, fungi, and collective biodegrading is being researched. Biodegradation occurs in various ways, each with its own mode of action.

6.1.1. Bacterial Assisted Degradation

Many scientists are studying the bacterial degradation of plastics and the formation of micro(nano)plastics. Based on Web of Science and Scopus online sources, bacteria that decompose plastics originated from landfills belong to the phyla Proteobacteria, Actinobacteria, and Firmicutes in the majority of biodegradable plastic-related studies [168]. The gut of a super worm contained Pseudomonas aeruginosa DSM 50071 (bacteria), which exhibited polystyrene degradation in lab experiment [169]. Likewise, Bacillus strains isolated from Peninsular Malaysian mangrove environments showed successful PP degradation in 40 days [170].
In association with bacteria in the body, certain bacteria performed biodegradation indirectly. In yellow mealworm larvae, polyvinylchloride can disintegrate and be converted into monomers, but gentamicin slows down this process by inhibiting intestinal microorganisms [171]. Gut microorganisms are claimed to be involved in polystyrene breakdown. For instance, polystyrene consumption substantially increased the gut microbial population in Achatina fulica (snails) and enhanced flora growth [172]. Likewise, Tribolium castaneum larvae have Acinetobacter bacterium, and Endosymbiotic bacteria in citrus mealybugs demonstrated successful polystyrene and polyethylene degradation, respectively [173,174]. Noticeably, a mixture of different microorganisms showed more efficient micro(nano)plastic degradation than a single bacterium whose degradation performance in a mixture of varying micro(nano)plastic is poor. Micro(nano)plastic is more complex and degraded by the environment, making its makeup more challenging to interpret. As a result, bacterial degradation is limited.

6.1.2. Fungal-Assisted Degradation

A laboratory study demonstrated that Zalerion maritimum (marine fungus) utilized polyethylene as a growth substrate and significantly reduced the proportion of polyethylene particles in 28 days [175]. Polyester yarn mineral media showed filamentous fungal tip proliferation in the presence of F. oxysporum and F. solani [166]. Fungi can degrade micro(nano)plastic using polymers for carbon and energy use [176]. Fungi can use enzymes that are not substrate-specific to detoxify contaminants and invade substrates. Mycelium may bind to hydrophobic substances and even infiltrate three-dimensional surfaces thanks to fungi’s hydrophobic proteins. For this purpose, the wax borer’s intestinal contents were used to isolate the Aspergillus flavus PEDX3 polyethylene-degrading fungus strain and successfully employed in HDPE conversion into lower molecular weight MPs during a four-week experiment [37]. In contrast with bacteria, fungi can degrade a wide range of plastics, which can only degrade specific plastics based on their enzymes. Thus, fungi can function on non-specific substrates, transcending the restrictions of bacterial degradation alone. Fungi have a lot of potential for breaking down plastics and micro(nano)plastic.

6.1.3. Algae-Assisted Biodegradation

Algae have also been found to produce chemicals to break down polymers utilizing these plastics as a carbon source. Future study into algal degradation holds a promising prospect [177]. Conventional bio-based biodegradable plastics have low water tolerance and mechanical qualities, whereas algal synthesized MPs are light, hydrophobic, and strong [178]. Furthermore, algal is simple to produce and harvest, making them time and space-efficient [179]. Although algal-based polymer synthesis is still limited to libraries scale, these polymers have enormous scope for application. Only a few algae species have been able to degrade plastic objects. The surface of plastics provides feasible sites for algal growth [177], where the colonies of algae degrade plastics by lignin and extracellular polysaccharides [179,180]. Plastic degradation has also been witnessed on plastic surfaces with algae (blue-green algae, green algae, diatom, Navicula pupula, and Scenedesmus dimorphus [181]. Furthermore, Khoironi et al. [181] discovered PET plastics degrade faster than PE plastics, implying plastic-type as a determinant factor in algal-based degradation. Furthermore, several bacteria and fungi which has the potential to degrade plastic are not well-tailored to the marine ecology, wherein a majority of plastic garbage build-up [182]. Still, such limitation endures by some harmless and non-toxic algae by establishing algal biofilms in varied polluted watercourses successfully impair plastic-oriented waste [182]. Algae can be utilized as microbial processing plants to create tailored PETase and directly degrade the waste of plastics [183]. Synthetic biology is a viable method for developing a green solution for lessening biological PET utilizing microalgae [177]. Establishing of Diatom, Navicula pupula, blue-green alga, Dimorphus, and Anabaena spiroides on the plastic surface showed breakdown [181]. Furthermore, Khoironi et al. [181,184] reported PET plastics break down faster than PE plastics, demonstrating that algae-based MPs biodegradation depends on the types of MPs. However, further research and analysis into the process and algae efficacy as micro(nano)plastic degraders is required.

6.1.4. Integrated Enzymatic Biodegradation (IEB)

The integrated enzymatic breakdown is among the main characteristics of biodegradable polymers. About all-polymer breakdown in the environment requires IEB [185]. Plastic-degrading microbes have been discovered in the digestive tracts of insects, on land, and in water, but they are quite particular and have very variable disintegrated [186,187]. IEB is a convenient tool to degrade plastics and micro(nano)plastic, despite relying on the enzyme of the precise organism, a distinct catalytic process, and a long incubation period. This is especially true nowadays, as a study on engineered enzymes raises the dominance of this biological breakdown process [185]. Only naturally occurring enzymes are insufficient to reply to breakdown MPs because their potency is low, and their range is not as diverse as we would prefer. As a result, research is being conducted to manufacture enzymes to maximize the effectiveness of enzyme breakdown.

6.1.5. Combined Biodegradation

Many studies have shown that one bacterium’s polymer biodegradation is poor, resulting in reduced microbial growth due to hazardous end products produced by polymer breakdown [188]. Some other investigated using many bacteria to build resilient microbial colonies to biodegrade MPs to overcome this limitation [186]: (1) products of microbial degradation by one bacterial strain often work as a substrate for the proliferation of a second bacterial strain [165]; (2) an ample and durable microbial group can confront sole microorganism precision and disintegrate in various MPs; (3) A healthy microbial community can efficiently degrade micro(nano)plastic in a synergistic symbiotic way. Many scientists have succeeded in developing a biodegradation method based on a range of microorganisms [14]. Taniguchi and colleagues [166] successfully demonstrated PET degradation by Three routes that can degrade PET: Ideonella sakaiensis 201-F6; Microbial Consortium No. 46; and MHETase and PETase. Park and Kim [159] speed up the decomposition of polyethylene MPs using hybrid bacterial culture primarily comprised of Paenibacillus sp. and Bacillus sp. obtained from landfills. To a significant extent, cumulative biological degradation surpasses the limits of sole bacterial degradation [159]. Nevertheless, due to the direct contact of numerous microorganisms and diverse enzymes, the breakdown and consumption of plastics and micro(nano)plastic by bacterial communities is a rather complex procedure, and the degradation mechanism and kinetics are still not precise. Moreover, the co-existence of other pollutants may alter the microbial community structure and their response to target pollutants [189,190]. Resultantly interaction of microorganisms with MPs and the efficiency of MPs degradation may change. Therefore, exploration of the main determinant elements and mechanisms is required.

6.2. Chemical Degradation

Miao and colleagues proposed a TiO2/graphite (TiO2/C) cathode, and six hours of electrolyzed Ag/AgCl at 0.7 V showed 75% dichlorination, oxidized the hydroxyl radicals, and degraded PVC simultaneously [191]. In another study, a nanocomposite membrane made of PLA/PE and TiO2 nanoparticles was subjected to simulated sunshine, which accelerates PLA and PE breakdown and modifies polymer structure [192]. By combining LDPE with a photothermosensitizer containing cobalt stearate or nanoparticles of silver, eco-friendly materials can be produced with promising PE degradation potential [193].
The process of hydrolysis chemically degrades plastics. The potential of plastic to disintegrate in water is determined by the presence of hydrolyzable covalent bonds such as ester, ether, anhydride, amide, carbamide, or ester amide groups, etc. The effectiveness of hydrolysis is influenced by water properties, time, pH, and temperature [194]. Plastics having hydrolyzable covalent linkages (for example, PET) can be hydrophilic, accelerating the polymer’s hydrolytic breakdown [14]. Polyester can cause hydrolysis when hydrogen ions target the ester bonds under alkaline or acidic conditions. Hydrolytic breakdown in alkaline conditions causes surface corrosion of polyesters in parallel to chain breaking [195]. Likewise, Lin et al. discovered that elevated UV radiation (3600 mJ cm2) could cause changes (such as cracks, wrinkles, and protrusions) in PET, PVC, and PS when compared to a regular dose of UV [196]. This phenomenon was likened to intense UV, causing the chemical bond to crack microplastics. UV radiation degrades micro(nano)plastics by cracking their C–H and C–C linkages [15].
Another efficient form of chemical degradation for MPs degradation is oxidative degradation. When oxygen is introduced to polymers, functional groups such as hydroxyl and carbon monoxide are generated, which assist in biological degradation [10]. Light or heat can stimulate the degradation activity, and nonhydrolyzable materials are especially susceptible to light and heath initiative OD [18]. Oxygen is found to be the most potent variable which causes material breakdown; ambient oxygen can strike covalent bonds in polymers, exacerbating the formation of free radicals; it has an oxidative degradation significantly affected by the structure of polymers, and OD produces peroxy radicals which break bonds and initiate side reactions [10]. Furthermore, the aging of plastics is substantially enhanced in the presence of ozone, causing them to degrade faster [194]. Due to its low economic efficiency, complete sulfate breakdown with a greater level of mineralization, and capacity to break down randomly tough plastics by bypassing the selectivity of microbial enzymatic biodegradation, advanced persulfate-based oxidation technologies have been developed and received a lot of interest [197]. Advanced oxidation processes (AOPs) are a chemical degradation type preceded by high oxidative reactive organic species (ROS). These species perform the degradation of organic pollutants in soil and water. AOPs include electrochemical oxidation, photocatalytic oxidation, and photochemical oxidation, which successfully apply MPs degradation [17]. The cosmetics sources’ MPs were degraded by the sulfate-based AOPs (SR-AOPs) method via catalytic activation of peroxynitrite by magnetic nanocomposites. Under a hydrothermal environment, activation of peroxynitrite yields reactive radicals with a high potential for MPs degradation [198]. ROS, an AOPs product, directly caused MP degradation, resulting in chain breakdown, product generation, and even total conversion of MPs into subunits [17].
Further studies are required to combine chemical treatment with other strategies. For instance, Mustafa et al. [199] recently concluded that combining degradation with filtration significantly enhanced the removal of emerging organic pollutants from water. Such strategies would be effective for the efficient removal of MPs.

7. Conclusions and Future Prospects

Systematically and explicitly, this review summarized the sampling, pretreatment, identification, health risk, and control of MPs pollution around the globe. In addition, both governmental policies and nongovernmental campaigns are listed to show the efforts to curb MPs worldwide. Undoubtedly, public concerns about MPs pollution, but more collaborative efforts are needed in the future to facilitate solution-searching efforts. For instance, the investigation protocols remain standardized, the fate of MPs in the environment requires more effort, either laboratory studies or modeling of transfer and transportation, and the ecotoxicological effects on living organisms and even humans based on the background concentrations of MPs in the environment. Furthermore, government organizations and public and private sectors will develop more innovative, long-term solutions to deal with plastic litters and MPs production.
Some countermeasures to curb plastic pollution may be achieved by changing industry behavior and influencing regulation, limiting the consumption of disposable plastic products, forbidding the supplement of plastic pellets into products (e.g., skin washing products, shower products, and others) at a much larger scale or worldwide, modifying plastic recycling actions and inventing feasible biodegradable plastics, and others. This article explores further the degradation mechanisms, influencing factors, and environmental interactions of degradable plastics. Public awareness regarding the wise use of plastic items and their reduction and recycling should also be extended to underdeveloped countries and small cities.
Additionally, a link has been demonstrated between plastic degradation and micro/nanoplastics. With additional research, it is possible to transform those non-biodegradable plastics into biodegradable plastics. Simultaneously, studies may conduct into the environmental impact and health risks of microplastic substances, laying the theoretical foundation for the future scientific selection of environmentally benign microplastic degradation procedures. Based on chemical and microbiological-assisted degradation information presented in this research, there are still some knowledge gaps, including (i) studies should focus more on designing methodologies to hinder the formation of hazardous intermediates by degradation, (ii) meanwhile parallel approaches that preferentially break chemical bonds in polymer MPs to produce usable organics at a high rate (for example, CH2O2, C2H4O4, CH4O, and C2H6O) and extract these molecules safely, (iii) following assisted degradation processes, the remaining MPs are detrimental to the environment, and MP biodegradation products have not been collected or recognized in a timely manner in most of the available literature, and (iv) also determine if interactions between various microbial species and distinct enzymes influence the degradation and utilization of MPs by microbial consortiums.

Author Contributions

Conceptualization, L.P. and T.M.; methodology, L.P. and T.M.; software, L.P., and R.B.; validation, L.P., R.B., Z.W. and D.F.; formal analysis, L.P.; investigation, L.P. and T.M.; resources, L.P.; data curation, L.P., T.M., R.B., Z.W. and D.F.; writing—original draft preparation, L.P. and T.M.; writing—review and editing, L.P., T.M. and R.B.; visualization, R.B., Z.W. and D.F.; supervision, L.P.; project administration, L.P.; funding acquisition, L.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [National Natural Science Foundation of China] grant number [41766003] And Start-up funding from Hainan University (kyqd(zr)1719). This work received full APC waiver by MDPI.

Acknowledgments

This study was supported, in part, by the National Natural Science Foundation of China (41766003), Start-up funding from Hainan University (kyqd(zr)1719).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. North, E.J.; Halden, R.U. Plastics and environmental health: The road ahead. Rev. Environ. Health 2013, 28, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bao, R.; Wang, Z.; Qi, H.; Mehmood, T.; Cai, M.; Zhang, Y.; Yang, R.; Peng, L.; Liu, F. Occurrence and distribution of microplastics in wastewater treatment plant in a tropical region of China. J. Clean. Prod. 2022, 349, 131454. [Google Scholar] [CrossRef]
  3. Ivleva, N.P.; Wiesheu, A.C.; Niessner, R. Microplastic in aquatic ecosystems. Angew. Chem. Int. Ed. 2017, 56, 1720–1739. [Google Scholar] [CrossRef] [PubMed]
  4. 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]
  5. Sridharan, S.; Kumar, M.; Bolan, N.S.; Singh, L.; Kumar, S.; Kumar, R.; You, S. Are microplastics destabilizing the global network of terrestrial and aquatic ecosystem services? Environ. Res. 2021, 198, 111243. [Google Scholar] [CrossRef]
  6. Wang, L.; Wu, W.-M.; Bolan, N.S.; Tsang, D.C.; Li, Y.; Qin, M.; Hou, D. Environmental fate, toxicity and risk management strategies of nanoplastics in the environment: Current status and future perspectives. J. Hazard. Mater. 2021, 401, 123415. [Google Scholar] [CrossRef]
  7. Cai, H.; Xu, E.G.; Du, F.; Li, R.; Liu, J.; Shi, H. Analysis of environmental nanoplastics: Progress and challenges. Chem. Eng. J. 2021, 410, 128208. [Google Scholar] [CrossRef]
  8. Fang, C.; Sobhani, Z.; Zhang, X.; McCourt, L.; Routley, B.; Gibson, C.T.; Naidu, R. Identification and visualisation of microplastics/nanoplastics by Raman imaging (iii): Algorithm to cross-check multi-images. Water Res. 2021, 194, 116913. [Google Scholar] [CrossRef]
  9. Peiponen, K.-E.; Räty, J.; Ishaq, U.; Pélisset, S.; Ali, R. Outlook on optical identification of micro- and nanoplastics in aquatic environments. Chemosphere 2019, 214, 424–429. [Google Scholar] [CrossRef]
  10. Lambert, S.; Sinclair, C.; Boxall, A. Occurrence, degradation, and effect of polymer-based materials in the environment. In Reviews of Environmental Contamination and Toxicology; Whitacre, D.M., Ed.; Springer International Publishing: Cham, Switzerland, 2014; Volume 227, pp. 1–53. [Google Scholar] [CrossRef]
  11. Sudhakar, M.; Trishul, A.; Doble, M.; Suresh Kumar, K.; Syed Jahan, S.; Inbakandan, D.; Viduthalai, R.R.; Umadevi, V.R.; Sriyutha Murthy, P.; Venkatesan, R. Biofouling and biodegradation of polyolefins in ocean waters. Polym. Degrad. Stab. 2007, 92, 1743–1752. [Google Scholar] [CrossRef]
  12. Sudhakar, M.; Priyadarshini, C.; Doble, M.; Sriyutha Murthy, P.; Venkatesan, R. Marine bacteria mediated degradation of nylon 66 and 6. Int. Biodeterior. Biodegrad. 2007, 60, 144–151. [Google Scholar] [CrossRef]
  13. Sudhakar, M.; Doble, M.; Murthy, P.S.; Venkatesan, R. Marine microbe-mediated biodegradation of low- and high-density polyethylenes. Int. Biodeterior. Biodegrad. 2008, 61, 203–213. [Google Scholar] [CrossRef]
  14. Krzan, A.; Hemjinda, S.; Miertus, S.; Corti, A.; Chiellini, E. Standardization and certification in the area of environmentally degradable plastics. Polym. Degrad. Stab. 2006, 91, 2819–2833. [Google Scholar] [CrossRef]
  15. Born, M.P.; Brüll, C. From model to nature—A review on the transferability of marine (micro-) plastic fragmentation studies. Sci. Total Environ. 2022, 811, 151389. [Google Scholar] [CrossRef]
  16. Rutkowska, M.; Heimowska, A.; Krasowska, K.; Janik, H. Biodegradability of polyethylene starch blends in sea water. Pol. J. Environ. Stud. 2002, 11, 267–271. [Google Scholar]
  17. Du, H.; Xie, Y.; Wang, J. Microplastic degradation methods and corresponding degradation mechanism: Research status and future perspectives. J. Hazard. Mater. 2021, 418, 126377. [Google Scholar] [CrossRef]
  18. Shan, J.; Zhao, J.; Zhang, Y.; Liu, L.; Wu, F.; Wang, X. Simple and rapid detection of microplastics in seawater using hyperspectral imaging technology. Anal. Chim. Acta 2019, 1050, 161–168. [Google Scholar] [CrossRef]
  19. Cincinelli, A.; Scopetani, C.; Chelazzi, D.; Lombardini, E.; Martellini, T.; Katsoyiannis, A.; Fossi, M.C.; Corsolini, S. Microplastic in the surface waters of the Ross Sea (Antarctica): Occurrence, distribution and characterization by FTIR. Chemosphere 2017, 175, 391–400. [Google Scholar] [CrossRef]
  20. Zheng, Y.; Li, J.; Sun, C.; Cao, W.; Wang, M.; Jiang, F.; Ju, P. Comparative study of three sampling methods for microplastics analysis in seawater. Sci. Total Environ. 2021, 765, 144495. [Google Scholar] [CrossRef]
  21. Liu, H.; Yang, X.; Liu, G.; Liang, C.; Xue, S.; Chen, H.; Ritsema, C.J.; Geissen, V. Response of soil dissolved organic matter to microplastic addition in Chinese loess soil. Chemosphere 2017, 185, 907–917. [Google Scholar] [CrossRef]
  22. Ziajahromi, S.; Drapper, D.; Hornbuckle, A.; Rintoul, L.; Leusch, F.D.L. Microplastic pollution in a stormwater floating treatment wetland: Detection of tyre particles in sediment. Sci. Total Environ. 2020, 713, 136356. [Google Scholar] [CrossRef] [PubMed]
  23. Pitt, J.A.; Kozal, J.S.; Jayasundara, N.; Massarsky, A.; Trevisan, R.; Geitner, N.; Wiesner, M.; Levin, E.D.; Di Giulio, R.T. Uptake, tissue distribution, and toxicity of polystyrene nanoparticles in developing zebrafish (Danio rerio). Aquat. Toxicol. 2018, 194, 185–194. [Google Scholar] [CrossRef] [PubMed]
  24. von Friesen, L.W.; Granberg, M.E.; Hassellöv, M.; Gabrielsen, G.W.; Magnusson, K. An efficient and gentle enzymatic digestion protocol for the extraction of microplastics from bivalve tissue. Mar. Pollut. Bull. 2019, 142, 129–134. [Google Scholar] [CrossRef] [PubMed]
  25. Cai, L.; Wang, J.; Peng, J.; Tan, Z.; Zhan, Z.; Tan, X.; Chen, Q. Characteristic of microplastics in the atmospheric fallout from Dongguan city, China: Preliminary research and first evidence. Environ. Sci. Pollut. Res. 2017, 24, 24928–24935. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, G.; Fu, Z.; Yang, H.; Wang, J. An overview of analytical methods for detecting microplastics in the atmosphere. TrAC Trends Anal. Chem. 2020, 130, 115981. [Google Scholar] [CrossRef]
  27. Fu, Z.; Chen, G.; Wang, W.; Wang, J. Microplastic pollution research methodologies, abundance, characteristics and risk assessments for aquatic biota in China. Environ. Pollut. 2020, 266, 115098. [Google Scholar] [CrossRef] [PubMed]
  28. Mehmood, T.; Peng, L. Polyethylene scaffold net and synthetic grass fragmentation: A source of microplastics in the atmosphere? J. Hazard. Mater. 2022, 429, 128391. [Google Scholar] [CrossRef] [PubMed]
  29. Lin, J.; Xu, X.-P.; Yue, B.-Y.; Li, Y.; Zhou, Q.-Z.; Xu, X.-M.; Liu, J.-Z.; Wang, Q.-Q.; Wang, J.-H. A novel thermoanalytical method for quantifying microplastics in marine sediments. Sci. Total Environ. 2021, 760, 144316. [Google Scholar] [CrossRef] [PubMed]
  30. Mai, L.; Bao, L.-J.; Shi, L.; Wong, C.S.; Zeng, E.Y. A review of methods for measuring microplastics in aquatic environments. Environ. Sci. Pollut. Res. 2018, 25, 11319–11332. [Google Scholar] [CrossRef]
  31. Ceccarini, A.; Corti, A.; Erba, F.; Modugno, F.; La Nasa, J.; Bianchi, S.; Castelvetro, V. The hidden microplastics: New insights and figures from the thorough separation and characterization of microplastics and of their degradation byproducts in coastal sediments. Environ. Sci. Technol. 2018, 52, 5634–5643. [Google Scholar] [CrossRef]
  32. Coppock, R.L.; Cole, M.; Lindeque, P.K.; Queirós, A.M.; Galloway, T.S. A small-scale, portable method for extracting microplastics from marine sediments. Environ. Pollut. 2017, 230, 829–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Song, Y.K.; Hong, S.H.; Eo, S.; Jang, M.; Han, G.M.; Isobe, A.; Shim, W.J. Horizontal and vertical distribution of microplastics in Korean coastal waters. Environ. Sci. Technol. 2018, 52, 12188–12197. [Google Scholar] [CrossRef] [PubMed]
  34. Li, Q.; Wu, J.; Zhao, X.; Gu, X.; Ji, R. Separation and identification of microplastics from soil and sewage sludge. Environ. Pollut. 2019, 254, 113076. [Google Scholar] [CrossRef] [PubMed]
  35. 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] [PubMed] [Green Version]
  36. Liu, M.; Lu, S.; Chen, Y.; Cao, C.; Bigalke, M.; He, D. Analytical methods for microplastics in environments: Current advances and challenges. In Microplastics in Terrestrial Environments: Emerging Contaminants and Major Challenges; He, D., Luo, Y., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 3–24. [Google Scholar] [CrossRef]
  37. Zhang, J.; Gao, D.; Li, Q.; Zhao, Y.; Li, L.; Lin, H.; Bi, Q.; Zhao, Y. Biodegradation of polyethylene microplastic particles by the fungus Aspergillus flavus from the guts of wax moth Galleria mellonella. Sci. Total Environ. 2020, 704, 135931. [Google Scholar] [CrossRef] [PubMed]
  38. Schwaferts, C.; Niessner, R.; Elsner, M.; Ivleva, N.P. Methods for the analysis of submicrometer- and nanoplastic particles in the environment. TrAC Trends Anal. Chem. 2019, 112, 52–65. [Google Scholar] [CrossRef]
  39. Vermeiren, P.; Muñoz, C.; Ikejima, K. Microplastic identification and quantification from organic rich sediments: A validated laboratory protocol. Environ. Pollut. 2020, 262, 114298. [Google Scholar] [CrossRef]
  40. Karlsson, T.M.; Kärrman, A.; Rotander, A.; Hassellöv, M. Comparison between manta trawl and in situ pump filtration methods, and guidance for visual identification of microplastics in surface waters. Environ. Sci. Pollut. Res. 2020, 27, 5559–5571. [Google Scholar] [CrossRef] [Green Version]
  41. Fu, D.; Zhang, Q.; Fan, Z.; Qi, H.; Wang, Z.; Peng, L. Aged microplastics polyvinyl chloride interact with copper and cause oxidative stress towards microalgae Chlorella vulgaris. Aquat. Toxicol. 2019, 216, 105319. [Google Scholar] [CrossRef]
  42. Hahn, A.; Gerdts, G.; Völker, C.; Niebühr, V. Using FTIRS as pre-screening method for detection of microplastic in bulk sediment samples. Sci. Total Environ. 2019, 689, 341–346. [Google Scholar] [CrossRef]
  43. Elkhatib, D.; Oyanedel-Craver, V. A critical review of extraction and identification methods of microplastics in wastewater and drinking water. Environ. Sci. Technol. 2020, 54, 7037–7049. [Google Scholar] [CrossRef] [PubMed]
  44. Sorensen, L.; Groven, A.S.; Hovsbakken, I.A.; Del Puerto, O.; Krause, D.F.; Sarno, A.; Booth, A.M. UV degradation of natural and synthetic microfibers causes fragmentation and release of polymer degradation products and chemical additives. Sci. Total Environ. 2021, 755, 143170. [Google Scholar] [CrossRef] [PubMed]
  45. Reichel, J.; Graßmann, J.; Letzel, T.; Drewes, J.E. Systematic development of a simultaneous determination of plastic particle identity and adsorbed organic compounds by thermodesorption–pyrolysis GC/MS (TD-Pyr-GC/MS). Molecules 2020, 25, 4985. [Google Scholar] [CrossRef] [PubMed]
  46. Li, P.; Li, Q.; Hao, Z.; Yu, S.; Liu, J. Analytical methods and environmental processes of nanoplastics. J. Environ. Sci. 2020, 94, 88–99. [Google Scholar] [CrossRef] [PubMed]
  47. Gewert, B.; Ogonowski, M.; Barth, A.; MacLeod, M. Abundance and composition of near surface microplastics and plastic debris in the Stockholm Archipelago, Baltic Sea. Mar. Pollut. Bull. 2017, 120, 292–302. [Google Scholar] [CrossRef] [PubMed]
  48. Di Mauro, R.; Kupchik, M.J.; Benfield, M.C. Abundant plankton-sized microplastic particles in shelf waters of the northern Gulf of Mexico. Environ. Pollut. 2017, 230, 798–809. [Google Scholar] [CrossRef] [PubMed]
  49. Zhao, S.; Zhu, L.; Li, D. Microplastic in three urban estuaries, China. Environ. Pollut. 2015, 206, 597–604. [Google Scholar] [CrossRef]
  50. Tsang, Y.Y.; Mak, C.W.; Liebich, C.; Lam, S.W.; Sze, E.T.P.; Chan, K.M. Microplastic pollution in the marine waters and sediments of Hong Kong. Mar. Pollut. Bull. 2017, 115, 20–28. [Google Scholar] [CrossRef]
  51. Desforges, J.-P.W.; Galbraith, M.; Dangerfield, N.; Ross, P.S. Widespread distribution of microplastics in subsurface seawater in the NE Pacific Ocean. Mar. Pollut. Bull. 2014, 79, 94–99. [Google Scholar] [CrossRef]
  52. van der Hal, N.; Ariel, A.; Angel, D.L. Exceptionally high abundances of microplastics in the oligotrophic Israeli Mediterranean coastal waters. Mar. Pollut. Bull. 2017, 116, 151–155. [Google Scholar] [CrossRef]
  53. Anderson, P.J.; Warrack, S.; Langen, V.; Challis, J.K.; Hanson, M.L.; Rennie, M.D. Microplastic contamination in Lake Winnipeg, Canada. Environ. Pollut. 2017, 225, 223–231. [Google Scholar] [CrossRef] [PubMed]
  54. Castillo, A.B.; Al-Maslamani, I.; Obbard, J.P. Prevalence of microplastics in the marine waters of Qatar. Mar. Pollut. Bull. 2016, 111, 260–267. [Google Scholar] [CrossRef]
  55. Güven, O.; Gökdağ, K.; Jovanović, B.; Kıdeyş, A.E. Microplastic litter composition of the Turkish territorial waters of the Mediterranean Sea, and its occurrence in the gastrointestinal tract of fish. Environ. Pollut. 2017, 223, 286–294. [Google Scholar] [CrossRef] [PubMed]
  56. Miller, M.E.; Kroon, F.J.; Motti, C.A. Recovering microplastics from marine samples: A review of current practices. Mar. Pollut. Bull. 2017, 123, 6–18. [Google Scholar] [CrossRef] [PubMed]
  57. Masura, J.; Baker, J.; Foster, G.; Arthur, C. Laboratory methods for the analysis of microplastics in the marine environment: Recommendations for quantifying synthetic particles in waters and sediments. NOAA Tech. Memo. NOS-ORR-48 2015. [Google Scholar]
  58. Zhang, W.; Zhang, S.; Wang, J.; Wang, Y.; Mu, J.; Wang, P.; Lin, X.; Ma, D. Microplastic pollution in the surface waters of the Bohai Sea, China. Environ. Pollut. 2017, 231, 541–548. [Google Scholar] [CrossRef]
  59. 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]
  60. Collard, F.; Gilbert, B.; Eppe, G.; Parmentier, E.; Das, K. Detection of anthropogenic particles in fish stomachs: An isolation method adapted to identification by Raman spectroscopy. Arch. Environ. Contam. Toxicol. 2015, 69, 331–339. [Google Scholar] [CrossRef]
  61. Tagg, A.S.; Sapp, M.; Harrison, J.P.; Ojeda, J.J. Identification and quantification of microplastics in wastewater using focal plane array-based reflectance micro-FT-IR imaging. Anal. Chem. 2015, 87, 6032–6040. [Google Scholar] [CrossRef] [Green Version]
  62. Green, D.S.; Kregting, L.; Boots, B.; Blockley, D.J.; Brickle, P.; da Costa, M.; Crowley, Q. A comparison of sampling methods for seawater microplastics and a first report of the microplastic litter in coastal waters of Ascension and Falkland Islands. Mar. Pollut. Bull. 2018, 137, 695–701. [Google Scholar] [CrossRef]
  63. Kedzierski, M.; Falcou-Préfol, M.; Kerros, M.E.; Henry, M.; Pedrotti, M.L.; Bruzaud, S. A machine learning algorithm for high throughput identification of FTIR spectra: Application on microplastics collected in the Mediterranean Sea. Chemosphere 2019, 234, 242–251. [Google Scholar] [CrossRef] [PubMed]
  64. Aytan, Ü.; Valente, A.; Senturk, Y.; Usta, R.; Sahin, F.B.E.; Mazlum, R.E.; Agirbas, E. First evaluation of neustonic microplastics in Black Sea waters. Mar. Environ. Res. 2016, 119, 22–30. [Google Scholar] [CrossRef] [PubMed]
  65. 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]
  66. Karami, A.; Golieskardi, A.; Choo, C.K.; Romano, N.; Ho, Y.B.; Salamatinia, B. A high-performance protocol for extraction of microplastics in fish. Sci. Total Environ. 2017, 578, 485–494. [Google Scholar] [CrossRef] [PubMed]
  67. Kun, W.; Kunde, L.; Dongxing, Y. Research progress on the analysis of microplastics in the environment. Environ. Chem. 2017, 36, 27–36. [Google Scholar] [CrossRef]
  68. Hidalgo-Ruz, V.; Gutow, L.; Thompson, R.C.; Thiel, M. Microplastics in the marine environment: A review of the methods used for identification and quantification. Environ. Sci. Technol. 2012, 46, 3060–3075. [Google Scholar] [CrossRef]
  69. Vianello, A.; Boldrin, A.; Guerriero, P.; Moschino, V.; Rella, R.; Sturaro, A.; Da Ros, L. Microplastic particles in sediments of Lagoon of Venice, Italy: First observations on occurrence, spatial patterns and identification. Estuar. Coast. Shelf Sci. 2013, 130, 54–61. [Google Scholar] [CrossRef]
  70. 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] [Green Version]
  71. Li, J.; Liu, H.; Chen, J.P. Microplastics in freshwater systems: A review on occurrence, environmental effects, and methods for microplastics detection. Water Res. 2018, 137, 362–374. [Google Scholar] [CrossRef]
  72. Yu, X.; Peng, J.; Wang, J.; Wang, K.; Bao, S. Occurrence of microplastics in the beach sand of the Chinese inner sea: The Bohai Sea. Environ. Pollut. 2016, 214, 722–730. [Google Scholar] [CrossRef]
  73. Retama, I.; Jonathan, M.P.; Shruti, V.C.; Velumani, S.; Sarkar, S.K.; Roy, P.D.; Rodríguez-Espinosa, P.F. Microplastics in tourist beaches of Huatulco Bay, Pacific coast of southern Mexico. Mar. Pollut. Bull. 2016, 113, 530–535. [Google Scholar] [CrossRef] [PubMed]
  74. Näkki, P.; Setälä, O.; Lehtiniemi, M. Bioturbation transports secondary microplastics to deeper layers in soft marine sediments of the northern Baltic Sea. Mar. Pollut. Bull. 2017, 119, 255–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Fischer, E.K.; Paglialonga, L.; Czech, E.; Tamminga, M. Microplastic pollution in lakes and lake shoreline sediments–a case study on Lake Bolsena and Lake Chiusi (central Italy). Environ. Pollut. 2016, 213, 648–657. [Google Scholar] [CrossRef]
  76. Nel, H.A.; Dalu, T.; Wasserman, R.J. Sinks and sources: Assessing microplastic abundance in river sediment and deposit feeders in an Austral temperate urban river system. Sci. Total Environ. 2018, 612, 950–956. [Google Scholar] [CrossRef] [PubMed]
  77. Akhbarizadeh, R.; Moore, F.; Keshavarzi, B.; Moeinpour, A. Microplastics and potentially toxic elements in coastal sediments of Iran’s main oil terminal (Khark Island). Environ. Pollut. 2017, 220, 720–731. [Google Scholar] [CrossRef] [PubMed]
  78. Maes, T.; Jessop, R.; Wellner, N.; Haupt, K.; Mayes, A.G. A rapid-screening approach to detect and quantify microplastics based on fluorescent tagging with Nile Red. Sci. Rep. 2017, 7, 44501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Yu, X.; Ladewig, S.; Bao, S.; Toline, C.A.; Whitmire, S.; Chow, A.T. Occurrence and distribution of microplastics at selected coastal sites along the southeastern United States. Sci. Total Environ. 2018, 613–614, 298–305. [Google Scholar] [CrossRef]
  80. Wahl, A.; Le Juge, C.; Davranche, M.; El Hadri, H.; Grassl, B.; Reynaud, S.; Gigault, J. Nanoplastic occurrence in a soil amended with plastic debris. Chemosphere 2021, 262, 127784. [Google Scholar] [CrossRef]
  81. Quinn, B.; Murphy, F.; Ewins, C. Validation of density separation for the rapid recovery of microplastics from sediment. Anal. Methods 2017, 9, 1491–1498. [Google Scholar] [CrossRef] [Green Version]
  82. Zobkov, M.; Esiukova, E. Microplastics in Baltic bottom sediments: Quantification procedures and first results. Mar. Pollut. Bull. 2017, 114, 724–732. [Google Scholar] [CrossRef]
  83. Abayomi, O.A.; Range, P.; Al-Ghouti, M.A.; Obbard, J.P.; Almeer, S.H.; Ben-Hamadou, R. Microplastics in coastal environments of the Arabian Gulf. Mar. Pollut. Bull. 2017, 124, 181–188. [Google Scholar] [CrossRef] [PubMed]
  84. Li, C.; Cui, Q.; Zhang, M.; Vogt, R.D.; Lu, X. 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] [PubMed]
  85. Crichton, E.M.; Noël, 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]
  86. Blumenröder, J.; Sechet, P.; Kakkonen, J.E.; Hartl, M.G.J. Microplastic contamination of intertidal sediments of Scapa Flow, Orkney: A first assessment. Mar. Pollut. Bull. 2017, 124, 112–120. [Google Scholar] [CrossRef] [PubMed]
  87. Sruthy, S.; Ramasamy, E.V. Microplastic pollution in Vembanad Lake, Kerala, India: The first report of microplastics in lake and estuarine sediments in India. Environ. Pollut. 2017, 222, 315–322. [Google Scholar] [CrossRef]
  88. Terepocki, A.K.; Brush, A.T.; Kleine, L.U.; Shugart, G.W.; Hodum, P. Size and dynamics of microplastic in gastrointestinal tracts of Northern Fulmars (Fulmarus glacialis) and Sooty Shearwaters (Ardenna grisea). Mar. Pollut. Bull. 2017, 116, 143–150. [Google Scholar] [CrossRef]
  89. Kolandhasamy, P.; Su, L.; Li, J.; Qu, X.; Jabeen, K.; Shi, H. Adherence of microplastics to soft tissue of mussels: A novel way to uptake microplastics beyond ingestion. Sci. Total Environ. 2018, 610–611, 635–640. [Google Scholar] [CrossRef]
  90. Maxwell, S.H.; Melinda, K.F.; Matthew, G. Counterstaining to separate nile red-stained microplastic particles from terrestrial invertebrate biomass. Environ. Sci. Technol. 2020, 54, 5580–5588. [Google Scholar] [CrossRef]
  91. Wang, J.; Coffin, S.; Sun, C.; Schlenk, D.; Gan, J. Negligible effects of microplastics on animal fitness and HOC bioaccumulation in earthworm Eisenia fetida in soil. Environ. Pollut. 2019, 249, 776–784. [Google Scholar] [CrossRef]
  92. Courtene-Jones, W.; Quinn, B.; Murphy, F.; Gary, S.F.; Narayanaswamy, B.E. Optimisation of enzymatic digestion and validation of specimen preservation methods for the analysis of ingested microplastics. Anal. Methods 2017, 9, 1437–1445. [Google Scholar] [CrossRef]
  93. Taylor, M.L.; Gwinnett, C.; Robinson, L.F.; Woodall, L.C. Plastic microfibre ingestion by deep-sea organisms. Sci. Rep. 2016, 6, 33997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. 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]
  95. Desforges, J.-P.W.; Galbraith, M.; Ross, P.S. Ingestion of microplastics by zooplankton in the northeast Pacific Ocean. Arch. Environ. Contam. Toxicol. 2015, 69, 320–330. [Google Scholar] [CrossRef] [PubMed]
  96. Hermsen, E.; Pompe, R.; Besseling, E.; Koelmans, A.A. Detection of low numbers of microplastics in North Sea fish using strict quality assurance criteria. Mar. Pollut. Bull. 2017, 122, 253–258. [Google Scholar] [CrossRef]
  97. Vendel, A.L.; Bessa, F.; Alves, V.E.N.; Amorim, A.L.A.; Patrício, J.; Palma, A.R.T. Widespread microplastic ingestion by fish assemblages in tropical estuaries subjected to anthropogenic pressures. Mar. Pollut. Bull. 2017, 117, 448–455. [Google Scholar] [CrossRef]
  98. Alomar, C.; Estarellas, F.; Deudero, S. Microplastics in the Mediterranean Sea: Deposition in coastal shallow sediments, spatial variation and preferential grain size. Mar. Environ. Res. 2016, 115, 1–10. [Google Scholar] [CrossRef]
  99. Sussarellu, R.; Suquet, M.; Thomas, Y.; Lambert, C.; Fabioux, C.; Pernet Marie Eve, J.; Le Goïc, N.; Quillien, V.; Mingant, C.; Epelboin, Y.; et al. Oyster reproduction is affected by exposure to polystyrene microplastics. Proc. Natl. Acad. Sci. USA 2016, 113, 2430–2435. [Google Scholar] [CrossRef] [Green Version]
  100. Cole, M.; Lindeque, P.; Halsband, C.; Galloway, T.S. Microplastics as contaminants in the marine environment: A review. Mar. Pollut. Bull. 2011, 62, 2588–2597. [Google Scholar] [CrossRef]
  101. Romeo, T.; Pietro, B.; Pedà, C.; Consoli, P.; Andaloro, F.; Fossi, M.C. First evidence of presence of plastic debris in stomach of large pelagic fish in the Mediterranean Sea. Mar. Pollut. Bull. 2015, 95, 358–361. [Google Scholar] [CrossRef]
  102. English, M.D.; Robertson, G.J.; Avery-Gomm, S.; Pirie-Hay, D.; Roul, S.; Ryan, P.C.; Wilhelm, S.I.; Mallory, M.L. Plastic and metal ingestion in three species of coastal waterfowl wintering in Atlantic Canada. Mar. Pollut. Bull. 2015, 98, 349–353. [Google Scholar] [CrossRef]
  103. Mascarenhas, R.; Santos, R.; Zeppelini, D. Plastic debris ingestion by sea turtle in Paraíba, Brazil. Mar. Pollut. Bull. 2004, 49, 354–355. [Google Scholar] [CrossRef]
  104. Santana, M.F.M.; Ascer, L.G.; Custódio, M.R.; Moreira, F.T.; Turra, A. Microplastic contamination in natural mussel beds from a Brazilian urbanized coastal region: Rapid evaluation through bioassessment. Mar. Pollut. Bull. 2016, 106, 183–189. [Google Scholar] [CrossRef] [PubMed]
  105. Li, J.; Qu, X.; Su, L.; Zhang, W.; Yang, D.; Kolandhasamy, P.; Li, D.; Shi, H. Microplastics in mussels along the coastal waters of China. Environ. Pollut. 2016, 214, 177–184. [Google Scholar] [CrossRef]
  106. Dehaut, A.; Cassone, A.-L.; Frère, L.; Hermabessiere, L.; Himber, C.; Rinnert, E.; Rivière, G.; Lambert, C.; Soudant, P.; Huvet, A.; et al. Microplastics in seafood: Benchmark protocol for their extraction and characterization. Environ. Pollut. 2016, 215, 223–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Mehmood, T.; Ahmad, I.; Bibi, S.; Mustafa, B.; Ali, I. Insight into monsoon for shaping the air quality of Islamabad, Pakistan: Comparing the magnitude of health risk associated with PM10 and PM2. 5 exposure. J. Air Waste Manag. Assoc. 2020, 70, 1340–1355. [Google Scholar] [CrossRef] [PubMed]
  108. Gasperi, J.; Wright, S.L.; Dris, R.; Collard, F.; Mandin, C.; Guerrouache, M.; Langlois, V.; Kelly, F.J.; Tassin, B. Microplastics in air: Are we breathing it in? Curr. Opin. Environ. Sci. Health 2018, 1, 1–5. [Google Scholar] [CrossRef] [Green Version]
  109. Gaston, E.; Woo, M.; Steele, C.; Sukumaran, S.; Anderson, S. Microplastics differ between indoor and outdoor air masses: Insights from multiple microscopy methodologies. Appl. Spectrosc. 2020, 74, 1079–1098. [Google Scholar] [CrossRef]
  110. Wright, S.L.; Ulke, J.; Font, A.; Chan, K.L.A.; Kelly, F.J. Atmospheric microplastic deposition in an urban environment and an evaluation of transport. Environ. Int. 2020, 136, 105411. [Google Scholar] [CrossRef] [PubMed]
  111. Dris, R.; Gasperi, J.; Mirande, C.; Mandin, C.; Guerrouache, M.; Langlois, V.; Tassin, B. A first overview of textile fibers, including microplastics, in indoor and outdoor environments. Environ. Pollut. 2017, 221, 453–458. [Google Scholar] [CrossRef] [Green Version]
  112. Dris, R.; Gasperi, J.; Saad, M.; Mirande, C.; Tassin, B. Synthetic fibers in atmospheric fallout: A source of microplastics in the environment? Mar. Pollut. Bull. 2016, 104, 290–293. [Google Scholar] [CrossRef]
  113. Sarkar, D.J.; Das Sarkar, S.; Das, B.K.; Praharaj, J.K.; Mahajan, D.K.; Purokait, B.; Mohanty, T.R.; Mohanty, D.; Gogoi, P.; Kumar, V.S.; et al. Microplastics removal efficiency of drinking water treatment plant with pulse clarifier. J. Hazard. Mater. 2021, 413, 125347. [Google Scholar] [CrossRef] [PubMed]
  114. Tong, H.; Jiang, Q.; Hu, X.; Zhong, X. Occurrence and identification of microplastics in tap water from China. Chemosphere 2020, 252, 126493. [Google Scholar] [CrossRef] [PubMed]
  115. Bretas Alvim, C.; Bes-Piá, M.A.; Mendoza-Roca, J.A. Separation and identification of microplastics from primary and secondary effluents and activated sludge from wastewater treatment plants. Chem. Eng. J. 2020, 402, 126293. [Google Scholar] [CrossRef]
  116. Pivokonsky, M.; Cermakova, L.; Novotna, K.; Peer, P.; Cajthaml, T.; Janda, V. Occurrence of microplastics in raw and treated drinking water. Sci. Total Environ. 2018, 643, 1644–1651. [Google Scholar] [CrossRef] [PubMed]
  117. Wiesheu, A.C.; Anger, P.M.; Baumann, T.; Niessner, R.; Ivleva, N.P. Raman microspectroscopic analysis of fibers in beverages. Anal. Methods 2016, 8, 5722–5725. [Google Scholar] [CrossRef] [Green Version]
  118. Mason, S.A.; Welch, V.G.; Neratko, J. Synthetic polymer contamination in bottled water. Front. Chem. 2018, 6, 407. [Google Scholar] [CrossRef] [Green Version]
  119. Schymanski, D.; Goldbeck, C.; Humpf, H.-U.; Fürst, 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]
  120. Mintenig, S.M.; Löder, M.G.J.; Primpke, S.; Gerdts, G. Low numbers of microplastics detected in drinking water from ground water sources. Sci. Total Environ. 2019, 648, 631–635. [Google Scholar] [CrossRef]
  121. Mehmood, T.; Hassan, M.A.; Faheem, M.; Shakoor, A. Why is inhalation the most discriminative route of microplastics exposure? Environ. Sci. Pollut. Res. 2022, 29, 49479–49482. [Google Scholar] [CrossRef]
  122. Schwabl, P.; Köppel, S.; Königshofer, P.; Bucsics, T.; Trauner, M.; Reiberger, T.; Liebmann, B. Detection of various microplastics in human stool. Ann. Intern. Med. 2019, 171, 453–457. [Google Scholar] [CrossRef]
  123. Prata, J.C. Airborne microplastics: Consequences to human health? Environ. Pollut. 2018, 234, 115–126. [Google Scholar] [CrossRef] [PubMed]
  124. Pauly, J.L.; Stegmeier, S.J.; Allaart, H.A.; Cheney, R.T.; Zhang, P.J.; Mayer, A.G.; Streck, R.J. Inhaled cellulosic and plastic fibers found in human lung tissue. Cancer Epidemiol. Biomark. Prev. 1998, 7, 419–428. [Google Scholar] [CrossRef]
  125. Law, B.D.; Bunn, W.B.; Hesterberg, T.W. Solubility of polymeric organic fibers and manmade vitreous fibers in gambles solution. Inhal. Toxicol. 1990, 2, 321–339. [Google Scholar] [CrossRef]
  126. Schirinzi, G.F.; Pérez-Pomeda, I.; Sanchís, J.; Rossini, C.; Farré, M.; Barceló, D. Cytotoxic effects of commonly used nanomaterials and microplastics on cerebral and epithelial human cells. Environ. Res. 2017, 159, 579–587. [Google Scholar] [CrossRef]
  127. Hassan, M.A.; Mehmood, T.; Lodhi, E.; Bilal, M.; Dar, A.A.; Liu, J. Lockdown Amid COVID-19 Ascendancy over Ambient Particulate Matter Pollution Anomaly. Int. J. Environ. Res. Public Health 2022, 19, 13540. [Google Scholar] [CrossRef]
  128. Deng, Y.; Zhang, Y.; Lemos, B.; Ren, H. Tissue accumulation of microplastics in mice and biomarker responses suggest widespread health risks of exposure. Sci. Rep. 2017, 7, 46687. [Google Scholar] [CrossRef] [Green Version]
  129. von Moos, N.; Burkhardt-Holm, P.; Köhler, A. Uptake and effects of microplastics on cells and tissue of the blue mussel Mytilus edulis L. after an experimental exposure. Environ. Sci. Technol. 2012, 46, 11327–11335. [Google Scholar] [CrossRef]
  130. Greven, A.-C.; Merk, T.; Karagöz, F.; Mohr, K.; Klapper, M.; Jovanović, B.; Palić, D. Polycarbonate and polystyrene nanoplastic particles act as stressors to the innate immune system of fathead minnow (Pimephales promelas). Environ. Toxicol. Chem. 2016, 35, 3093–3100. [Google Scholar] [CrossRef]
  131. Jones, A.E.; Watts, J.A.; Debelak, J.P.; Thornton, L.R.; Younger, J.G.; Kline, J.A. Inhibition of prostaglandin synthesis during polystyrene microsphere-induced pulmonary embolism in the rat. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2003, 284, L1072–L1081. [Google Scholar] [CrossRef] [Green Version]
  132. Geiser, M.; Rothen-Rutishauser, B.; Kapp, N.; Schürch, S.; Kreyling, W.; Schulz, H.; Semmler, M.; Im Hof, V.; Heyder, J.; Gehr, P. Ultrafine particles cross cellular membranes by nonphagocytic mechanisms in lungs and in cultured cells. Environ. Health Perspect 2005, 113, 1555–1560. [Google Scholar] [CrossRef]
  133. Mehmood, T.; Gaurav, G.K.; Cheng, L.; Klemeš, J.J.; Usman, M.; Bokhari, A.; Lu, J. A review on plant-microbial interactions, functions, mechanisms and emerging trends in bioretention system to improve multi-contaminated stormwater treatment. J. Environ. Manag. 2021, 294, 113108. [Google Scholar] [CrossRef] [PubMed]
  134. Bouwmeester, H.; Hollman, P.C.H.; Peters, R.J.B. Potential health impact of environmentally released micro- and nanoplastics in the human food production chain: Experiences from nanotoxicology. Environ. Sci. Technol. 2015, 49, 8932–8947. [Google Scholar] [CrossRef] [PubMed]
  135. Oberbeckmann, S.; Löder, M.G.J.; Labrenz, M. Marine microplastic-associated biofilms—A review. Environ. Chem. 2015, 12, 551–562. [Google Scholar] [CrossRef]
  136. Mehmood, T.; Liu, C.; Niazi, N.K.; Gaurav, G.K.; Ashraf, A.; Bibi, I. Compost-mediated arsenic phytoremediation, health risk assessment and economic feasibility using Zea mays L. in contrasting textured soils. Int. J. Phytoremediation 2021, 23, 899–910. [Google Scholar] [CrossRef]
  137. Wang, F.; Feng, X.; Liu, Y.; Adams, C.A.; Sun, Y.; Zhang, S. Micro (nano) plastics and terrestrial plants: Up-to-date knowledge on uptake, translocation, and phytotoxicity. Resour. Conserv. Recycl. 2022, 185, 106503. [Google Scholar] [CrossRef]
  138. Heusinkveld, H.J.; Wahle, T.; Campbell, A.; Westerink, R.H.; Tran, L.; Johnston, H.; Stone, V.; Cassee, F.R.; Schins, R.P. Neurodegenerative and neurological disorders by small inhaled particles. Neurotoxicology 2016, 56, 94–106. [Google Scholar] [CrossRef]
  139. Hu, J.; Qin, X.; Zhang, J.; Zhu, Y.; Zeng, W.; Lin, Y.; Liu, X. Polystyrene microplastics disturb maternal-fetal immune balance and cause reproductive toxicity in pregnant mice. Reprod Toxicol. 2021, 106, 42–50. [Google Scholar] [CrossRef]
  140. Dubey, I.; Khan, S.; Kushwaha, S. Developmental and reproductive toxic effects of exposure to microplastics: A review of associated signaling pathways. Front. Toxicol. 2022, 4, 901798. [Google Scholar] [CrossRef]
  141. Plastics Europe. An Analysis of European Plastics Production, Demand and Waste Data. 2019. Available online: https://plasticseurope.org/knowledge-hub/plastics-the-facts-2019/ (accessed on 12 September 2022).
  142. Worm, B.; Lotze, H.K.; Jubinville, I.; Wilcox, C.; Jambeck, J. Plastic as a persistent marine pollutant. Annu. Rev. Environ. Resour. 2017, 42, 1–26. [Google Scholar] [CrossRef]
  143. Awasthi, S.; Srivastava, P.; Singh, P.; Tiwary, D.; Mishra, P.K. Biodegradation of thermally treated high-density polyethylene (HDPE) by Klebsiella pneumoniae CH001. 3 Biotech 2017, 7, 332. [Google Scholar] [CrossRef]
  144. Legislative Assembly of Ontario. Bill 75, Microbead Elimination and Monitoring Act; Legislative Assembly of Ontario: Toronto, ON, Canada, 2015. [Google Scholar]
  145. United States Congress. Microbead-Free Waters Act of 2015; United States Congress: Washington, DC, USA, 2015. [Google Scholar]
  146. Inter IKEA Group. Sustainability Summary Report FY17; Inter IKEA Group: Delft, The Netherlands, 2017. [Google Scholar]
  147. Schnurr, R.E.J.; Alboiu, V.; Chaudhary, M.; Corbett, R.A.; Quanz, M.E.; Sankar, K.; Srain, H.S.; Thavarajah, V.; Xanthos, D.; Walker, T.R. Reducing marine pollution from single-use plastics (SUPs): A review. Mar. Pollut. Bull. 2018, 137, 157–171. [Google Scholar] [CrossRef] [PubMed]
  148. Hartley, B.L.; Thompson, R.C.; Pahl, S. Marine litter education boosts children’s understanding and self-reported actions. Mar. Pollut. Bull. 2015, 90, 209–217. [Google Scholar] [CrossRef] [PubMed]
  149. Wong, C.S.; Green, D.R.; Cretney, W.J. Quantitative tar and plastic waste distributions in the Pacific Ocean. Nature 1974, 247, 30–32. [Google Scholar] [CrossRef]
  150. Yang, D.; Shi, H.; Li, L.; Li, J.; Jabeen, K.; Kolandhasamy, P. Microplastic pollution in table salts from China. Environ. Sci. Technol. 2015, 49, 13622–13627. [Google Scholar] [CrossRef]
  151. Iñiguez, M.E.; Conesa, J.A.; Fullana, A. Microplastics in Spanish table salt. Sci. Rep. 2017, 7, 8620. [Google Scholar] [CrossRef]
  152. Gündoğdu, S. Contamination of table salts from Turkey with microplastics. Food Addit. Contam. Part A Chem. Anal. Control Expo. Risk Assess. 2018, 35, 1006–1014. [Google Scholar] [CrossRef]
  153. Kosuth, M.; Mason, S.A.; Wattenberg, E.V. Anthropogenic contamination of tap water, beer, and sea salt. PLoS ONE 2018, 13. [Google Scholar] [CrossRef]
  154. Bombelli, P.; Howe, C.J.; Bertocchini, F. Polyethylene bio-degradation by caterpillars of the wax moth Galleria mellonella. Curr. Biol. 2017, 27, R292–R293. [Google Scholar] [CrossRef] [Green Version]
  155. Huerta Lwanga, E.; Thapa, B.; Yang, X.; Gertsen, H.; Salánki, T.; Geissen, V.; Garbeva, P. Decay of low-density polyethylene by bacteria extracted from earthworm’s guts: A potential for soil restoration. Sci. Total Environ. 2018, 624, 753–757. [Google Scholar] [CrossRef]
  156. 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]
  157. Maroof, L.; Khan, I.; Yoo, H.S.; Kim, S.; Park, H.-T.; Ahmad, B.; Azam, S. Identification and characterization of low density polyethylene-degrading bacteria isolated from soils of waste disposal sites. Environ. Eng. Res. 2021, 26, 200167. [Google Scholar] [CrossRef]
  158. Taghavi, N.; Singhal, N.; Zhuang, W.-Q.; Baroutian, S. Degradation of plastic waste using stimulated and naturally occurring microbial strains. Chemosphere 2021, 263, 127975. [Google Scholar] [CrossRef] [PubMed]
  159. 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] [PubMed]
  160. Munir, E.; Harefa, R.S.M.; Priyani, N.; Suryanto, D. Plastic degrading fungi Trichoderma viride and Aspergillus nomius isolated from local landfill soil in Medan. IOP Conf. Ser. Earth Environ. Sci. 2018, 126, 012145. [Google Scholar] [CrossRef]
  161. Muhonja, C.N.; Makonde, H.; Magoma, G.; Imbuga, M. Biodegradability of polyethylene by bacteria and fungi from Dandora dumpsite Nairobi-Kenya. PLoS ONE 2018, 13, e0198446. [Google Scholar] [CrossRef] [Green Version]
  162. Thomas, B.T.; Olanrewaju-Kehinde, D.S.K.; Popoola, O.D.; James, E.S. Degradation of plastic and polythene materials by some selected microorganisms isolated from soil. World Appl. Sci. J. 2015, 33, 1888–1891. [Google Scholar] [CrossRef]
  163. Skariyachan, S.; Setlur, A.S.; Naik, S.Y.; Naik, A.A.; Usharani, M.; Vasist, K.S. Enhanced biodegradation of low and high-density polyethylene by novel bacterial consortia formulated from plastic-contaminated cow dung under thermophilic conditions. Environ. Sci. Pollut. Res. Int. 2017, 24, 8443–8457. [Google Scholar] [CrossRef]
  164. Delacuvellerie, A.; Cyriaque, V.; Gobert, S.; Benali, S.; Wattiez, R. The plastisphere in marine ecosystem hosts potential specific microbial degraders including Alcanivorax borkumensis as a key player for the low-density polyethylene degradation. J. Hazard. Mater. 2019, 380, 120899. [Google Scholar] [CrossRef]
  165. Singh, L.; Wahid, Z.A. Methods for enhancing bio-hydrogen production from biological process: A review. J. Ind. Eng. Chem. 2015, 21, 70–80. [Google Scholar] [CrossRef]
  166. Taniguchi, I.; Yoshida, S.; Hiraga, K.; Miyamoto, K.; Kimura, Y.; Oda, K. Biodegradation of PET: Current status and application aspects. ACS Catal. 2019, 9, 4089–4105. [Google Scholar] [CrossRef]
  167. Zhang, Y.; Kang, S.; Allen, S.; Allen, D.; Gao, T.; Sillanpää, M. Atmospheric microplastics: A review on current status and perspectives. Earth-Sci. Rev. 2020, 203, 103118. [Google Scholar] [CrossRef]
  168. Matjašič, T.; Simčič, T.; Medvešček, N.; Bajt, O.; Dreo, T.; Mori, N. Critical evaluation of biodegradation studies on synthetic plastics through a systematic literature review. Sci. Total Environ. 2021, 752, 141959. [Google Scholar] [CrossRef] [PubMed]
  169. Kim, H.R.; Lee, H.M.; Yu, H.C.; Jeon, E.; Lee, S.; Li, J.; Kim, D.-H. Biodegradation of Polystyrene by Pseudomonas sp. Isolated from the Gut of Superworms (Larvae of Zophobas atratus). Environ. Sci. Technol. 2020, 54, 6987–6996. [Google Scholar] [CrossRef] [PubMed]
  170. Auta, H.S.; Emenike, C.U.; Jayanthi, B.; Fauziah, S.H. Growth kinetics and biodeterioration of polypropylene microplastics by Bacillus sp. and Rhodococcus sp. isolated from mangrove sediment. Mar. Pollut. Bull. 2018, 127, 15–21. [Google Scholar] [CrossRef]
  171. Peng, B.Y.; Chen, Z.; Chen, J.; Yu, H.; Zhou, X.; Criddle, C.S.; Wu, W.M.; Zhang, Y. Biodegradation of polyvinyl chloride (PVC) in Tenebrio molitor (Coleoptera: Tenebrionidae) larvae. Environ. Int. 2020, 145, 106106. [Google Scholar] [CrossRef]
  172. Song, Y.; Qiu, R.; Hu, J.; Li, X.; Zhang, X.; Chen, Y.; Wu, W.-M.; He, D. Biodegradation and disintegration of expanded polystyrene by land snails Achatina fulica. Sci. Total Environ. 2020, 746, 141289. [Google Scholar] [CrossRef] [PubMed]
  173. Ibrahim, S.; Gupta, R.K.; War, A.R.; Hussain, B.; Kumar, A.; Sofi, T.; Noureldeen, A.; Darwish, H. Degradation of chlorpyriphos and polyethylene by endosymbiotic bacteria from citrus mealybug. Saudi J. Biol. Sci. 2021, 28, 3214–3224. [Google Scholar] [CrossRef]
  174. Wang, Z.; Xin, X.; Shi, X.; Zhang, Y. A polystyrene-degrading Acinetobacter bacterium isolated from the larvae of Tribolium castaneum. Sci. Total Environ. 2020, 726, 138564. [Google Scholar] [CrossRef]
  175. Paço, 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]
  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]
  177. Ali, S.S.; Elsamahy, T.; Al-Tohamy, R.; Zhu, D.; Mahmoud, Y.A.G.; Koutra, E.; Metwally, M.A.; Kornaros, M.; Sun, J. Plastic wastes biodegradation: Mechanisms, challenges and future prospects. Sci. Total Environ. 2021, 780, 146590. [Google Scholar] [CrossRef] [PubMed]
  178. Chia, W.Y.; Ying Tang, D.Y.; Khoo, K.S.; Kay Lup, A.N.; Chew, K.W. Nature’s fight against plastic pollution: Algae for plastic biodegradation and bioplastics production. Environ. Sci. Ecotechnol. 2020, 4, 100065. [Google Scholar] [CrossRef] [PubMed]
  179. Tang, D.Y.Y.; Khoo, K.S.; Chew, K.W.; Tao, Y.; Ho, S.-H.; Show, P.L. Potential utilization of bioproducts from microalgae for the quality enhancement of natural products. Bioresour. Technol. 2020, 304, 122997. [Google Scholar] [CrossRef] [PubMed]
  180. Sarmah, P.; Rout, J. Efficient biodegradation of low-density polyethylene by cyanobacteria isolated from submerged polyethylene surface in domestic sewage water. Environ. Sci. Pollut. Res. Int. 2018, 25, 33508–33520. [Google Scholar] [CrossRef]
  181. Khoironi, A.; Anggoro, S.; Sudarno, S. Evaluation of the Interaction among microalgae Spirulina sp, plastics polyethylene terephthalate and polypropylene in freshwater environment. J. Ecol. Eng. 2019, 20, 161–173. [Google Scholar] [CrossRef]
  182. Tanasupawat, S.; Takehana, T.; Yoshida, S.; Hiraga, K.; Oda, K. Ideonella sakaiensis sp. nov., isolated from a microbial consortium that degrades poly(ethylene terephthalate). Int. J. Syst. Evol. Microbiol. 2016, 66, 2813–2818. [Google Scholar] [CrossRef] [PubMed]
  183. Moog, D.; Schmitt, J.; Senger, J.; Zarzycki, J.; Rexer, K.-H.; Linne, U.; Erb, T.J.; Maier, U.G. Correction to: Using a marine microalga as a chassis for polyethylene terephthalate (PET) degradation. Microb. Cell Fact. 2020, 19, 1. [Google Scholar] [CrossRef]
  184. Vimal Kumar, R.; Kanna, G.R.; Elumalai, S. Biodegradation of polyethylene by green photosynthetic microalgae. J. Bioremed. Biodegr. 2017, 8, 1–8. [Google Scholar] [CrossRef] [Green Version]
  185. Zurier, H.S.; Goddard, J.M. Biodegradation of microplastics in food and agriculture. Curr. Opin. Food Sci. 2021, 37, 37–44. [Google Scholar] [CrossRef]
  186. Yuan, J.; Ma, J.; Sun, Y.; Zhou, T.; Zhao, Y.; Yu, F. Microbial degradation and other environmental aspects of microplastics/plastics. Sci. Total Environ. 2020, 715, 136968. [Google Scholar] [CrossRef]
  187. Danso, D.; Chow, J.; Streit, W.R. Plastics: Environmental and biotechnological perspectives on microbial degradation. Appl. Environ. Microbiol. 2019, 85, e01095-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Dobretsov, S.; Abed, R.M.; Teplitski, M. Mini-review: Inhibition of biofouling by marine microorganisms. Biofouling 2013, 29, 423–441. [Google Scholar] [CrossRef] [PubMed]
  189. Mehmood, T.; Lu, J.; Liu, C.; Gaurav, G.K. Organics removal and microbial interaction attributes of zeolite and ceramsite assisted bioretention system in copper-contaminated stormwater treatment. J. Environ. Manag. 2021, 292, 112654. [Google Scholar] [CrossRef] [PubMed]
  190. Gao, L.; Xie, Y.; Su, Y.; Mehmood, T.; Bao, R.; Fan, H.; Peng, L. Elucidating the negatively influential and potentially toxic mechanism of single and combined micro-sized polyethylene and petroleum to Chlorella vulgaris at the cellular and molecular levels. Ecotoxicol. Environ. Saf. 2022, 245, 114102. [Google Scholar] [CrossRef]
  191. Miao, F.; Liu, Y.; Gao, M.; Yu, X.; Xiao, P.; Wang, M.; Wang, S.; Wang, X. 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]
  192. da Silva, K.I.M.; Fernandes, J.A.; Kohlrausch, E.C.; Dupont, J.; Santos, M.J.L.; Gil, M.P. Structural stability of photodegradable poly(l-lactic acid)/PE/TiO2 nanocomposites through TiO2 nanospheres and TiO2 nanotubes incorporation. Polym. Bull. 2014, 71, 1205–1217. [Google Scholar] [CrossRef]
  193. Firestone, G.; Huang, H.; Bochinski, J.R.; Clarke, L.I. Photothermally-driven thermo-oxidative degradation of low density polyethylene: Heterogeneous heating plus a complex reaction leads to homogeneous chemistry. Nanotechnology 2019, 30, 475706. [Google Scholar] [CrossRef]
  194. Lucas, N.; Bienaime, C.; Belloy, C.; Queneudec, M.; Silvestre, F.; Nava-Saucedo, J.-E. Polymer biodegradation: Mechanisms and estimation techniques—A review. Chemosphere 2008, 73, 429–442. [Google Scholar] [CrossRef]
  195. Iskander Magued, G.; Hassan, M. Accelerated degradation of recycled plastic piling in aggressive soils. J. Compos. Constr. 2001, 5, 179–187. [Google Scholar] [CrossRef]
  196. Lin, J.; Yan, D.; Fu, J.; Chen, Y.; Ou, H. Ultraviolet-C and vacuum ultraviolet inducing surface degradation of microplastics. Water Res. 2020, 186, 116360. [Google Scholar] [CrossRef]
  197. Zhang, D.; Wu, L.; Yao, J.; Herrmann, H.; Richnow, H.-H. Carbon and hydrogen isotope fractionation of phthalate esters during degradation by sulfate and hydroxyl radicals. Chem. Eng. J. 2018, 347, 111–118. [Google Scholar] [CrossRef] [Green Version]
  198. Kang, J.-H.; Kwon, O.-Y.; Shim, W.J. Potential threat of microplastics to zooplanktivores in the surface waters of the southern sea of Korea. Arch. Environ. Contam. Toxicol. 2015, 69, 340–351. [Google Scholar] [CrossRef] [PubMed]
  199. Mustafa, B.; Mehmood, T.; Wang, Z.; Chofreh, A.G.; Shen, A.; Yang, B.; Yuan, J.; Wu, C.; Liu, Y.; Lu, W.; et al. Next-generation graphene oxide additives composite membranes for emerging organic micropollutants removal: Separation, adsorption and degradation. Chemosphere 2022, 308, 136333. [Google Scholar] [CrossRef] [PubMed]
Figure 3. Procedures of MPs investigation in the environment.
Figure 3. Procedures of MPs investigation in the environment.
Sustainability 14 14338 g003
Figure 4. Sampling tools used in the investigation of MPs.
Figure 4. Sampling tools used in the investigation of MPs.
Sustainability 14 14338 g004
Figure 5. Supporting policies enacted and campaigns launched to curb worldwide MPs pollution.
Figure 5. Supporting policies enacted and campaigns launched to curb worldwide MPs pollution.
Sustainability 14 14338 g005
Table 1. Common methods used in the sampling and identification of MPs in seawater.
Table 1. Common methods used in the sampling and identification of MPs in seawater.
ProceduresCommon MethodsRefs
Sampling ToolPlankton net; Manta trawl; Neuston net; Continue intake; Teflon pump[47,48,49,50,51]
ReservationFormalin solution (4%, 5%); Refrigerated; Ethanol (70%, 90%)[52,53,54]
DigestionHydrogen peroxide (30%); Sodium hydroxide (5M); Nitric acid (65%); Fe (II) solution (0.05 M)[47,55,56]
SeparationSieving; Density separation, filtration[57]
Pore Size (μm)0.45; 0.70; 0.75[58]
Temperature for drying50 ℃; 55 ℃; 60 °C; Air-dried[59]
IdentificationStereomicroscope; SEM; μ-FTIR or Raman[60]
Table 2. Typical sampling and identification of MPs and NPs in sediment or soil.
Table 2. Typical sampling and identification of MPs and NPs in sediment or soil.
ParametersRange/Chemicals/InstrumentsRefs
Sampling depth (cm)1; 2; 3; 5; 10[59,72]
Mesh size (mm)1; 2; 5[59,75]
Dry temperatureFreeze-dried; Air-dried; 40 ℃; 50 ℃; 60 ℃[74,75]
Separation solutionNaCl; NaI; KI; ZnCl2 (1.37~1.7 g/mL); ZnBr2; Canola oil[77,78]
Filter paper mesh size (μm)0.7; 0.8; 1.0; 1.2[50,79]
IdentificationStereomicroscope; SEM; μ-FTIR or μ-Raman; Py-GC-MS[60,61,80]
Table 3. Advantages and disadvantages of density separation solutions.
Table 3. Advantages and disadvantages of density separation solutions.
Solution (Density)AdvantagesDisadvantagesRefs
NaCl (1.2 g/mL)Low cost, environmental soundHigh-density plastic particles are not applicable[75]
NaI (1.57 g/mL)Cheaper, less environmental dangerousOxidizes filter paper, makes filter black, and makes it difficult to distinguish MPs[81]
KI (1.254 g/mL)Applicable to high-density polymerExpensive, oxidize filter paper[83]
ZnCl2 (1.6 g/mL)The high recovery rate of plastic particles (even high-density ones)High cost, hazardous to the environment[82]
ZnBr22 (1.71 g/mL)Good recoveries for larger MPsExpensive, severely hazardous to the environment[81]
Canola oilHigh recovery rateInterfere the chemical analysis of MPs[85]
NaBrLow cost, Environmental friendlyNot applicable to some types of plastics[84]
Table 4. Common methods used in the sampling and identification of MPs in organisms.
Table 4. Common methods used in the sampling and identification of MPs in organisms.
ItemsParametersRefs
Biological speciesBirds; Fish; Mussel; Bivalves; Invertebrates[55,89,93]
Tissue typeDigestive tract; All soft tissue; Whole animal; Gastrointestinal tract[93,94]
PreservationFrozen at −20 °C; Ethanol; Formalin[60,89]
DigestionAcid (HNO3); Enzymatic; H2O2; NaClO; Alkaline[56,66]
Digestion temperature25 ℃; 65 °C; 100 ℃[76,104,105]
SeparationFlotation; Filtration; Sieving[71,104]
IdentificationStereomicroscope; SEM; μ-FTIR or μ-Raman[52,60,61]
Table 5. Advantages and disadvantages of different digestion methods in MPs study.
Table 5. Advantages and disadvantages of different digestion methods in MPs study.
Digestion MethodDigestion SolutionAdvantageDisadvantageRefs
Acid digestionHCl (5% or 37%)Fully digested fish tissuesMay degrade some plastic polymers[66]
HNO3 (65% or 69%)Full digestion and inexpensiveCorrosive acid, may degrade plastics[92]
HClO4 (65% or 68%)Common reagent in labs, easy to obtainCorrosive acid, may degrade plastics[105]
HF (40%)Full digestionmay impact the structural or chemical integrity of plastics[56]
H2O2 (30% or 35%)No effect on any plastic particlesIncomplete digestion[96]
Alkaline digestionKOH (10%)The most suitable solution to digest fish tissueMay hinder FTIR if not cleaned[66]
NaOH (5M)Low chemical harm and low costUnknown effects on polymer types[55]
Enzymatic digestionEnzyme (proteinase and collagenase)High digestion efficiencyIncomplete digestion and may be worse for nylon fibres[60]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peng, L.; Mehmood, T.; Bao, R.; Wang, Z.; Fu, D. An Overview of Micro(Nano)Plastics in the Environment: Sampling, Identification, Risk Assessment and Control. Sustainability 2022, 14, 14338. https://doi.org/10.3390/su142114338

AMA Style

Peng L, Mehmood T, Bao R, Wang Z, Fu D. An Overview of Micro(Nano)Plastics in the Environment: Sampling, Identification, Risk Assessment and Control. Sustainability. 2022; 14(21):14338. https://doi.org/10.3390/su142114338

Chicago/Turabian Style

Peng, Licheng, Tariq Mehmood, Ruiqi Bao, Zezheng Wang, and Dongdong Fu. 2022. "An Overview of Micro(Nano)Plastics in the Environment: Sampling, Identification, Risk Assessment and Control" Sustainability 14, no. 21: 14338. https://doi.org/10.3390/su142114338

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