Nano-Microbial Remediation of Polluted Soil: Comparison
Please note this is a comparison between Version 2 by Peter Tang and Version 1 by Behnam Asgari Lajayer.

Nanobiotechnology has been used to bio-remediate or reclaim soil contaminated with organic and inorganic pollutants. The removal of pollutants from industrial wastes is a major challenge. The utilization of nanomaterials is gaining popularity, which might be accredited to their enhanced physical, chemical, and mechanical qualities. The development of advanced nanobiotechnological techniques involving the use of nanomaterials for the reclamation of polluted soils has indicated promising results and future hope for sustainable agriculture. By manufacturing environment-friendly nanomaterials, the industrial expenditure on decreasing the load of pollution might be reduced. A potential emerging domain of nanotechnology for eco-friendly production and cost reduction is “green biotechnology”, alongside the utilization of microorganisms in nanoparticle synthesis.

  • bioremediation
  • environmental pollution
  • nanomaterials
  • remediation

1. Introduction

Agriculture is considered one of the most important human activities, as it is the main source of food, feed, fuel, and fiber [1]. This activity can cause many environmental problems, especially when insecticides and mineral fertilizers are used in excess [2,3][2][3]. Therefore, agricultural contamination might refer to several activities that lead to the destruction or pollution of agroecosystems and affect human well-being [4]. In other words, agricultural soil, soil health, and fertility have been drastically impacted by many different types and classes of pollutants [5]. Some contaminants have a longer lifespan and are recalcitrant. They persist in the soil for many years, disrupting the food chain and causing biological imbalances in the soil, ultimately endangering human health [6]. Pesticides, fertilizers, household and industrial wastewater, industrial activities, and automobile pollution are the major anthropogenic sources of hazardous toxic metals and/or metalloids in the soil [7].
The pressing need is environmental remediation, which must be addressed as a priority [8,9,10][8][9][10]. In recent decades, various techniques have been used for this purpose, such as mycoremediation [11], phytoremediation [12[12][13],13], vermiremediation [4], bioremediation [6], remediation by using biosorbent materials such as biochar [14], fly ash and organic fertilizers [15], humic substances [16], and nanomaterials (NMs) via green remediation; or combined remediation [17].
The notion of sustainable remediation has recently gained much attention [13], as it essentially aims to reduce the concentration of contamination to risk-free levels while avoiding additional environmental impacts [6]. Several recent developments in this field have combined multiple technologies into a system that provides a cost-effective and time-saving way to disinfect a site while being able to restore the site’s quality. As an economical and environmentally sound means of remediating polluted areas, bioremediation is one of the solutions to problems of pollution [18]. The use of microorganisms to remove contaminants from the soil is the key principle of bioremediation [19]. As defined by the Environmental Protection Agency, bioremediation involves the biodegradation of hazardous pollutants to reduce their toxicity or intensity. It offers a number of benefits over physicochemical approaches, including high selectivity, specificity, cost and energy performance, and low demand. However, bioremediation has the disadvantage that it takes longer to degrade toxic compounds, usually several months to a year. It also limits the use of sites that are heavily contaminated with toxic pollutants, resulting in a loss in terms of resource utilization [4,15][4][15]. Nanoparticles are used in many scientific fields including automobiles, cosmetics, agriculture, foods, textiles, aviation, defense, engineering, medicine, and the environment [20,21,22][20][21][22]. According to the National Nanotechnology Initiative of the United States (NNI), there are relatively few studies on using nanotechnology in the analysis and manipulation of materials up to 100 nanometers in size, where unique phenomena enable novel applications of nanotechnology [23]. As an integrated field of nanoscale science, technology, and engineering, nanotechnology consists of viewing, analyzing, modeling, and manipulating materials within this size range. In recent years, nanotechnology has been increasingly used to remove contaminants due to its smaller particulate matter, high surface-to-volume ratio, ease of deployment at impact sites, flexibility, and other advantages [24]. The utilization of nanotechnology for environmental remediation has attracted much attention [23]. Ongoing research and many publications show how nanotechnology can tackle remediation duties and challenges [25,26][25][26]. Nanoremediation is a technology that has been recognized as environmentally beneficial by the Environmental Protection Agency. It is acknowledged as a viable strategy for traditional site cleaning [27]. Various techniques for using NMs for soil and water reclamation (nanoremediation) have been reported, such as nano-phytoremediation [28[28][29],29], nano-bioremediation [30], nano-Fe3O4 [31], nano zero-valent iron [32], nano-hydroxyapatite [33], nano zeolite [34], nano zero-valent iron [35], ZnO-nanoparticles (NPs) [36], nano-TiO2 [37], stabilized NPs [38], and nano-silica [39].

2. An Evaluation of Nanobioremediation-Based Pollutant Reduction with a Focus on Microbe-Mediated Remediation

Previous techniques for removing heavy metals (HMs) from contaminated soils include biosorption and bioaccumulation utilizing crops and bacteria. However, recent evidence has revealed that the use of NPs in the remediation of HMs has produced impressive results [40]. It has been found that the use of NPs in conjunction with specific microbes, either sequentially or simultaneously, has provided promising results [41]. Not only can they aid in the removal of HMs, but they can also act as nanocarriers for microbial populations or microbial adsorbents [42]. The integration of NPs with microbes for bioremediation is a two-phase procedure that combines biotic and abiotic factors [43]. After entering the system, the contaminants encounter a series of physical methods and revisions that include abiotic mechanisms such as uptake, adsorption, and dissolution, as well as synthetic catalyst supports for photocatalysis during the first stage [44]. Biocides, bioaccumulation, biostimulation, and biotransformation are examples of biotic systems in the second stage [45]. These biotic systems are essential for removing pollutants from the mechanism. Table 1 provides an overview of various NP-mediated pollutants removed from contaminated media.
Table 1.
Summary of the various nanoparticle-mediated pollutants removed from contaminated media.

Nanoparticles

Contaminant Remediated

Factors of Performance and Removal Efficiency

References

Iron oxide nanoparticles with a polyvinyl pyrrolidone coating

Cd, Pb

The use of nanoparticles was combined with a bioremediation process driven by Halomonas sp. Halomonas sp. was inoculated for 48 h at 180 rpm and 28 °C in the Cd and Pb removal system. After 24 h, 100% removal was detected, whereas it took 48 h for Cd.

[46]

Industrial suspension of zero-valent iron (nZVI) at two dosages (1% and 10%)

As

The pH of the nZVI suspension was adjusted to 12.2 ± 0.1. Polyacrylic acid was utilized as a stabilizer to prevent the accumulation of nZVI in the suspension. The maximum amount of As immobilized in brownfield soil was 10% of nZVI.

[47]

Graphene oxide nanoparticles (nGOx) and nZVI

Metals such as Cd, Pb, Zn, Cu, and As were found in As- and metal-contaminated soil.

The application of nZVI and nGOx to contaminated soils had a significant influence on the availability of As and metals. nGOx immobilized Cu, Pb, and Cd while mobilizing As and P. In the case of nZVI, it successfully immobilized As and Pb (but not Cd) while increasing Cu’s availability. This study discovered that both NPs may work as techniques for immobilization and stabilization, which can then be used for phytoremediation.

[48]

Titanium oxide nanoparticles bonded to a chitosan nanolayer (NTiO2–NCh)

Cd and Cu

During the experiment, the pH was adjusted at 7.0. The elimination was aided by a microwave-enforced sorption technique that lasted 60–70 s. Cu and Cd were eliminated at a rate of 88.01% and 70.67%, respectively, when NTiO2–NCh was used.

[49]

Palladium (Pd), Pd NPs

Cr

Pd NPs were investigated as a bionanocatalyst. Pd NPs were shown to decrease Cr6+ completely in 12 h. To decrease 5.0 mol of Cr6+, 6.3 mg of Pd NPs was utilized.

[50]

Magnetic iron oxide nanoparticles (Fe3O4 NPs) were treated with Staphylococcus aureus and had their surfaces encapsulated in phthalic acid (n-Fe3O4–Phth–S. aureus)

Cu, Ni, Pb

The remediation efficiency of n-Fe3O4–Phth–S. aureus was reported to be 83.0–89.5% for Cu2+, 99.4–100% for Pb2+, and 92.6–7.5% for Ni2+. The researchers also discovered that n-Fe3O4–Phth–S. aureus was an effective biosorbent for removing pollutants.

[51]

ZnO NPs

Cu, Cd, Cr, and Pb

The maximum removal of Cr, Cu, and Pb by ZnO-NPs at 5 mg L−1 with Bacillus cereus and Lysinibacillus macroides was 60%, 70%, and 85%, respectively. The ideal pH for effective removal was 8.0. The elimination was reduced in the case of bacteria-mediated remediation, which was determined to be 83% and 70% with B. cereus, and 60% and 65% for L. macroides.

[52]

3. Soil Nanoremediation

Nanoremediation is a virtually new application of nanotechnology for addressing environmental pollution issues [75][53]. Recently, this technique has been used to treat hazardous waste. Although it is a new technical sector, the application of nanotechnologies for environmental remediation has recently attracted a lot of attention from the scientific community [76][54]. The use of zero-valent iron (ZVI) as a permeability barrier was the first research idea developed by Gillham [77][55] based on their experience with NPs in decontamination of water-halogenated contaminants [78][56]. Many researchers utilized chemical synthesis, whereas others use green leaf extracts similar to those used to remove pollutants in aqueous solutions to create zero-valent iron NPs. The use of NPs can effectively degrade numerous pollutants such as organic halocarbons [79][57], nitrates, HMs [80][58], pesticides, and dyes [81][59]. There have been very few studies that have applied NP technology for the remediation of contaminants in soil; studies in this field have instead used for the decontamination of water or aqueous solutions [82][60].
Studies have reported that NPs can adsorb pollutants and facilitate their destruction through redox reactions, surface reactions, ion exchange, surface complexation, electrostatic contact, and adsorption [83][61]. A bentonite matrix was used by Shi et al. [82][60] to remove Cr (VI) from water and soil solutions using ZVI nanoparticles (nZVI) and Fe NPs with zero valency (B-nZVI). They discovered that the use of bentonite (B-nZVI) as a carrier material increased the effectiveness of nZVI nanoparticles, resulting in reduced aggregation and improvements in the active surface area. Likewise, the temperature was directly proportional to the amount of Cr (VI) removed, as were pH and total B-nZVI, which decreased with an increase in pH [84][62]. B-nZVI NPs have large surface areas and are highly reactive, enabling them to work as excellent adsorbents of Cr (VI) [82][60]. A wide variety of contaminants have been studied using NPs, including chlorinated organic compounds, insecticides, phenols and amines, organic acids, and chlorinated organic compounds [84][62]. Two decades ago, experiments showed that NPs, when injected into the soil, could remain effective for up to 56 days and could travel up to 20 m through the groundwater [85][63]. Zhang [85][63] reported that it was possible to remove over 99% of trichloroethene (TCE) from polluted locations within a few days.
Studies have shown that zero valent iron NPs trapped in silica microspheres can decompose polybrominated diphenyl ethers, a type of environmental pollutant that can readily accumulate in the soil [86,87][64][65]. Tetrahydrofuran (THF) was used by Qiu et al. [86][64] to degrade decabromodiphenyl ether from an aqueous solution. The researchers discovered that it was efficient in a solution of THF and water when exposed to environmental and temperature stress. Moreover, the study of Xie et al. [87][65] suggested that the removal efficiency or elimination efficiency for decabromodiphenyl ether in soil achieved by this degradation process was 78%. It was more significant than the biomass of plants treated with NPs. Additionally, Cr (VI) phytotoxicity was investigated, and iron NPs supported by bio-carbon were tested on cabbage mustard, which showed increased growth and lowered Cr (VI) levels. With the injection of 8 g per kg of soil, the immobilization efficiency for Cr (VI) and overall chromium (Cr) was 100% and 91.94%, correspondingly, in remediation experiments [88][66]. A lipid derivative of choline-coated silica NPs was used for bioremediation of polycyclic aromatic hydrocarbons (PAHs). Other NMs that have been used include iron sulfide stabilized by carboxymethylcellulose, which was tested for the consolidation of mercury in soils that were heavily contaminated with this metal [89][67].
According to various publications in 2016, the experimental processes and parameters for the synthesis of NPs differ, making it difficult to compare the efficiency gains due to variability in their structures, compositions, and morphologies, all of which impact the adsorption capacities for comparable pollutants. Currently, there is a lack of information on how they break down various types of toxins. The need for comprehensive studies on NMs is underlined by the lack of knowledge on their mechanisms of recovery and reuse, as well as their widespread application and effectiveness for the remediation of industrial effluents and polluted soils. Nevertheless, the reported results have indicated that this remediation technique is valuable compared with conventional techniques.
The effects of nanomaterials on various ecosystems, and their function, life cycle, and release of metal ions are still largely unexplored. Nanoremediation offers several advantages, including lower costs and shorter clean-up times for polluted areas, as well as the possibility to apply it on a large scale. However, to avoid negative impacts on the environment, detailed studies are needed to examine the effects of nanoremediation at the ecosystem level.

4. Microorganism-Assisted Nanoremediation

The use of nanoremediation is more sustainable and environmentally friendly if the NPs are biologically produced and microbes are used at the same time. Chemically produced NMs may have many disadvantages in terms of chemical consumption and self-agglomeration in aqueous solutions. In this regard, the utilization of plant extracts, and fungal and bacterial enzymes for green NP production might be a promising option. In this process, metallic NPs are created due to their reducing effect on the metal complex salts. Co-precipitation, or the addition of proteins and bioactive components to the outer surfaces of the NPs, greatly increased their strength in an aqueous environment. Mahanty et al. [92][68] found Aspergillus tubingensis (STSP 25) biofabricated iron oxide NPs from the rhizosphere of Avicennia officinalis in Sundarbans, India. About 90% of the HMs (Ni (II), Cu (II), Pb (II), and Zn (II)) in wastewater were eliminated or removed by the synthesized NPs, which had a regeneration potential of up to five cycles. The metal ions were chemically bound to the surface of the NPs by an endothermic reaction [92][68]. The co-precipitation of iron oxide NPs and exopolysaccharides (EPS) from Chlorella vulgaris has been described in other studies. The effective alteration of NPs by EPS functional groups was demonstrated by FT-IR spectroscopy. It was also demonstrated that the nanocomposite could remove 85% of NH4+ ions and 91% of PO43− ions [93][69].
It has been claimed that using bacteria to produce NPs is a practical and beneficial method for the environment. A copper-resistant Escherichia species, SINT7, was used to synthesize copper NPs. Biogenic NPs were observed to degrade azo dyes and textile effluents. Consequently, at a lower concentration of 25 mg/L, 83.6%, 90.6%, 97.1%, and 88.4% of reaction black-5, malachite green, Congo red, and direct blue-1 were lowered, respectively. When the concentration was increased to 100 mg/L, they reduced by 76.84%, 31.1%, 83.90%, and 62.32%, respectively. Additionally, treated samples of industrial sewage contained less phosphate and chloride ions, along with the suspended particles. The performance of biogenic NPs such as these may boost cost-effective and long-term industrial manufacturing [94][70]. Cheng et al. [95][71] used no additional sulfur to make iron-sulfur NPs. These NPs had the ability to annihilate Naphthol Green B dye through the extracellular transfer of electrons. The utilization of Pseudoalteromonas sp. CF10-13 in manufacturing NPs offers an environmentally acceptable biodegradation method. The manufacturing of toxic gases and metal complexes was constrained by the endogenous creation of NPs.
The use of biological particles is a more effective way to remediate industrial wastewater. As well as the direct production of NPs from microbes, there are several other ways in which microorganisms can contribute to the advancement of nanotechnology. In addition to NPs, microorganisms can also provide catalytic enzymes that help in wastewater treatment.

5. Utilization of Nanomaterials for Micro-Remediation of Polluted Soils

Bioremediation using of microorganisms has been proposed as a supposedly efficient approach to remediating contaminated sites [105][72]. Microorganisms that are capable of modifying soils contaminated with HMs and organic pollutants have attracted much attention. Volatilization, metal-binding, alteration, and chemical precipitation are some of the techniques used for the remediation of HMs using microorganisms [106,107][73][74]. According to Xu et al. [108][75], the following elements bind metals to microbial cells: CrO42−, Cu2+, Hg2+, Au3+, Cd2+, Ni2+, Pd2+, and Zn2+. The mobility of these metals and their harmful consequences were diminished by this metal-binding. Furthermore, Polti et al. [109][76] studied the use of microorganisms for the bioremediation of Cr (VI)-contaminated soils. Soil samples showed that the Streptomyces species MC1 was able to reduce Cr (VI) to Cr (III), the latter being more stable and less hazardous than the former. Metals that are volatile, such as Hg, can be volatilized by microbes [110][77]. On the other hand, organic pollutants can be destroyed by a variety of microorganisms or enzymes. Certain microorganisms can use the nitrogen and carbon in organic contaminants, leading to soil decontamination.
For example, four microbes were isolated from soils planted with bamboo, pine, and rice to treat polluted soils, including Rhodotorula glutinis 4CD4, Pseudomonas nitroreducens 4CD2, Pseudomonas putida 4CD1, and Pseudomonas putida 4CD3. All the isolated microorganisms effectively broke down p-hydroxybenzoic acid, ferulic acid, p-coumaric acid, and p-hydroxybenzaldehyde using phenols as a carbon source [111][78]. Pseudomonas stutzeri OX1 has also been shown to be able to break down tetrachloroethylene. Due to the production of toluene-xylene monooxygenase, which induces the aerobic breakdown of pollutants in bacteria, researchers ascribed this degradation to Pseudomonas stutzeri OX1 [112][79]. The effectiveness of micro-remediation for remediating pollutants is impacted by the potential impact of NMs on microorganisms. Shrestha et al. [113][80] investigated the influence of NMs on the architecture and function of soil microbial communities using MWCNTs. According to pyrosequencing research, applying 10 g kg−1 MWCNTs increased the abundance of many bacterial taxa such as Cellulomonas, Pseudomonas, Nocardioides, and Rhodococcus, which are thought to be potential degraders of resistant pollutants. NMs were also observed to affect the level of microbial assembly in the soil in favor of species that were more resilient to NMs or were capable of rapid degradation, which was advantageous for soil micro-remediation. Research on the breakdown of 2,4-dichlorophenoxyacetic acid in soils was carried out utilizing Fe3O4 NPs in combination with soil microorganisms. The addition of Fe3O4 NPs to the soil increased the microbial diversity and enzymatic activity (e.g., acid phosphatase, amylase, urease, and catalase), resulting in greater organic waste degradation efficiency than when the soil was treated with microorganisms alone [114][81]. Tilston et al. [115][82] discovered that the use of nZVI coated with polyacrylic acid (PAA) altered the composition of the bacterial community in the contaminated soil and reduced the efficiency of chloroaromatic mineralizing microorganisms. Populations of Dehalococcoides, a bacterium capable of dechlorinating chlorinated organic pollutants, were similarly reduced with 0.1 g L−1 nZVI [111][78].
The adverse upper effect of NMs on microorganisms prevented the biodegradation of pollutants in polluted soils. Furthermore, the impact of NMs on micro-remediation of polluted soils differs depending on the type and concentration of NMs. With increased CNT concentrations, extractability and microbial degradation of PAHs were assumed to decrease with an increasing CNT concentration. Compared with MWCNTs, SWCNTs had a stronger influence on the mineralization and extraction of PAHs [116][83]. Furthermore, a high concentration of MWCNTs significantly inhibited the development of phenanthrene-catabolizing bacteria as well as the growth of phenanthrene-degrading bacteria in the soil, while fullerene and low levels of CNTs had no negative influence on microbial activity [117][84]. When discussing bioremediation, it is important to remember that the pollutants that NMs ingest affect their bioavailability to microorganisms. The reduced bioavailability of contaminants affects the microbial remediation power in the polluted soils. MWCNTs adsorbed on phenanthrene were studied for their biodegradation and mineralization by Agrobacterium. It was found that the use of MWCNTs as contaminants significantly reduced the bioavailability of hydrophobic organic compounds in the environment [118][85]. Other research used 14C-2,4-DCP as the target contaminant to explore the mineralization, breakdown, and residual distribution of radioactively tagged 2,4-dichlorophenol (14C-2,4-DCP) in conjunction with SWCNTs and MWCNTs. In contaminated soils, SWCNTs at a concentration of 2 g kg−1 significantly reduced microbial mineralization and the breakdown of 14C-2,4-DCP. The reduced bioavailability of 2,4-DCP, the potential microbial toxicity of CNTs, and the reduced activity of native soil microorganisms all had inhibitory effects [119][86]. Overall, NMs have both beneficial and detrimental effects on the micro-remediation of contaminated soils (Figure 1).
Figure 1.
Beneficial and negative impacts of nanomaterials on micro-remediation of soil. Green arrows indicate upregulation. Red arrows indicate downregulation.

References

  1. Tian, H.; Qin, Y.; Niu, Z.; Wang, L.; Ge, S. Summer Maize Mapping by Compositing Time Series Sentinel-1A Imagery Based on Crop Growth Cycles. J. Indian Soc. Remote Sens. 2021, 49, 2863–2874.
  2. Yang, Y.; Chen, X.; Liu, L.; Li, T.; Dou, Y.; Qiao, J.; Wang, Y.; An, S.; Chang, S.X. Nitrogen fertilization weakens the linkage between soil carbon and microbial diversity: A global meta-analysis. Glob. Chang. Biol. 2022, 28, 6446–6461.
  3. Yang, Y.; Li, T.; Pokharel, P.; Liu, L.; Qiao, J.; Wang, Y.; An, S.; Chang, S.X. Global effects on soil respiration and its temperature sensitivity depend on nitrogen addition rate. Soil Biol. Biochem. 2022, 174, 108814.
  4. Khan, M.I.; Cheema, S.A.; Anum, S.; Niazi, N.K.; Azam, M.; Bashir, S.; Ashraf, I.; Qadri, R. Phytoremediation of agricultural pollutants. In Phytoremediation; Springer: Cham, Switzerland, 2020; pp. 27–81.
  5. Ghodszad, L.; Reyhanitabar, A.; Maghsoodi, M.R.; Lajayer, B.A.; Chang, S.X. Biochar affects the fate of phosphorus in soil and water: A critical review. Chemosphere 2021, 283, 131176.
  6. Karimi, H.; Mahdavi, S.; Asgari Lajayer, B.; Moghiseh, E.; Rajput, V.D.; Minkina, T.; Astatkie, T. Insights on the bioremediation technologies for pesticide-contaminated soils. Environ. Geochem. Health 2022, 44, 1329–1354.
  7. Souza, L.R.R.; Pomarolli, L.C.; da Veiga, M.A.M.S. From classic methodologies to application of nanomaterials for soil remediation: An integrated view of methods for decontamination of toxic metal (oid) s. Environ. Sci. Pollut. Res. 2020, 27, 10205–10227.
  8. Dai, J.; Feng, H.; Shi, K.; Ma, X.; Yan, Y.; Ye, L.; Xia, Y. Electrochemical degradation of antibiotic enoxacin using a novel PbO2 electrode with a graphene nanoplatelets inter-layer: Characteristics, efficiency and mechanism. Chemosphere 2022, 307, 135833.
  9. Keshavarzi, A.; Kumar, V. Spatial distribution and potential ecological risk assessment of heavy metals in agricultural soils of Northeastern Iran. Geol. Ecol. Landsc. 2020, 4, 87–103.
  10. Tan, Z.; Dong, B.; Xing, M.; Sun, X.; Xi, B.; Dai, W.; He, C.; Luo, Y.; Huang, Y. Electric field applications enhance the electron transfer capacity of dissolved organic matter in sludge compost. Environ. Technol. 2022, 1–11.
  11. Butu, M.; Stef, R.; Corneanu, M.; Butnariu, M. Mycoremediation: A Sustainable Approach for Pesticide Pollution Abatement. In Bioremediation and Biotechnology; Springer: Cham, Switzerland, 2020; Volume 2, pp. 73–96.
  12. Shahi Khalaf Ansar, B.; Kavusi, E.; Dehghanian, Z.; Pandey, J.; Asgari Lajayer, B.; Price, G.W.; Astatkie, T. Removal of organic and inorganic contaminants from the air, soil, and water by algae. Environ. Sci. Pollut. Res. 2022, 1–29.
  13. Kavusi, E.; Ansar, B.S.K.; Ebrahimi, S.; Sharma, R.; Ghoreishi, S.S.; Nobaharan, K.; Abdoli, S.; Dehghanian, Z.; Lajayer, B.A.; Senapathi, V. Critical review on phytoremediation of polyfluoroalkyl substances from environmental matrices: Need for global concern. Environ. Res. 2022, 217, 114844.
  14. Wang, J.; Shi, L.; Zhai, L.; Zhang, H.; Wang, S.; Zou, J.; Shen, Z.; Lian, C.; Chen, Y. Analysis of the long-term effectiveness of biochar immobilization remediation on heavy metal contaminated soil and the potential environmental factors weakening the remediation effect: A review. Ecotoxicol. Environ. Saf. 2021, 207, 111261.
  15. Hu, X.; Huang, X.; Zhao, H.; Liu, F.; Wang, L.; Zhao, X.; Gao, P.; Li, X.; Ji, P. Possibility of using modified fly ash and organic fertilizers for remediation of heavy-metal-contaminated soils. J. Clean. Prod. 2021, 284, 124713.
  16. Dong, Y.; Lin, H.; Zhao, Y.; Menzembere, E.R.G.Y. Remediation of vanadium-contaminated soils by the combination of natural clay mineral and humic acid. J. Clean. Prod. 2021, 279, 123874.
  17. Gu, W.; Li, X.; Li, Q.; Hou, Y.; Zheng, M.; Li, Y. Combined remediation of polychlorinated naphthalene-contaminated soil under multiple scenarios: An integrated method of genetic engineering and environmental remediation technology. J. Hazard. Mater. 2021, 405, 124139.
  18. Delangiz, N.; Aliyar, S.; Pashapoor, N.; Nobaharan, K.; Lajayer, B.A.; Rodríguez-Couto, S. Can polymer-degrading microorganisms solve the bottleneck of plastics’ environmental challenges? Chemosphere 2022, 294, 133709.
  19. Delangiz, N.; Varjovi, M.B.; Lajayer, B.A.; Ghorbanpour, M. Beneficial microorganisms in the remediation of heavy metals. In Molecular Aspects of Plant Beneficial Microbes in Agriculture; Elsevier: Amsterdam, The Netherlands, 2020; pp. 417–423.
  20. Atigh, Z.B.Q.; Sardari, P.; Moghiseh, E.; Lajayer, B.A.; Hursthouse, A.S. Purified montmorillonite as a nano-adsorbent of potentially toxic elements from environment: An overview. Nanotechnol. Environ. Eng. 2021, 6, 12.
  21. Yousefi, S.R.; Alshamsi, H.A.; Amiri, O.; Salavati-Niasari, M. Synthesis, characterization and application of Co/Co3O4 nanocomposites as an effective photocatalyst for discoloration of organic dye contaminants in wastewater and antibacterial properties. J. Mol. Liq. 2021, 337, 116405.
  22. Mahdi, M.A.; Yousefi, S.R.; Jasim, L.S.; Salavati-Niasari, M. Green synthesis of DyBa2Fe3O7. 988/DyFeO3 nanocomposites using almond extract with dual eco-friendly applications: Photocatalytic and antibacterial activities. Int. J. Hydrogen Energy 2022, 47, 14319–14330.
  23. Maghsoodi, M.R.; Lajayer, B.A.; Hatami, M.; Mirjalili, M.H. Challenges and opportunities of nanotechnology in plant-soil mediated systems: Beneficial role, phytotoxicity, and phytoextraction. Adv. Phytonanotechnol. 2019, 2019, 379–404.
  24. Khalkhal, K.; Asgari Lajayer, B.; Ghorbanpour, M. An overview on the effect of soil physicochemical properties on the immobilization of biogenic nanoparticles. In Biogenic Nano-Particles and their Use in Agro-Ecosystems; Springer: Singapore, 2020; pp. 133–160.
  25. Tratnyek, P.G.; Johnson, R.L. Nanotechnologies for environmental cleanup. Nano Today 2006, 1, 44–48.
  26. Patil, S.S.; Shedbalkar, U.U.; Truskewycz, A.; Chopade, B.A.; Ball, A.S. Nanoparticles for environmental clean-up: A review of potential risks and emerging solutions. Environ. Technol. Innov. 2016, 5, 10–21.
  27. Rajput, V.D.; Minkina, T.; Upadhyay, S.K.; Kumari, A.; Ranjan, A.; Mandzhieva, S.; Sushkova, S.; Singh, R.K.; Verma, K.K. Nanotechnology in the Restoration of Polluted Soil. Nanomaterials 2022, 12, 769.
  28. Srivastav, A.; Yadav, K.K.; Yadav, S.; Gupta, N.; Singh, J.K.; Katiyar, R.; Kumar, V. Nano-phytoremediation of pollutants from contaminated soil environment: Current scenario and future prospects. In Phytoremediation; Springer: Cham, Switzerland, 2018; pp. 383–401.
  29. Romeh, A.A.; Saber, R.A.I. Green nano-phytoremediation and solubility improving agents for the remediation of chlorfenapyr contaminated soil and water. J. Environ. Manag. 2020, 260, 110104.
  30. Singh, R.; Behera, M.; Kumar, S. Nano-bioremediation: An innovative remediation technology for treatment and management of contaminated sites. In Bioremediation of Industrial Waste for Environmental Safety; Springer: Singapore, 2020; pp. 165–182.
  31. Tran, T.D.; Dao, N.T.; Sasaki, R.; Tu, M.B.; Dang, G.H.M.; Nguyen, H.G.; Dang, H.M.; Vo, C.H.; Inigaki, Y.; van Nguyen, N. Accelerated remediation of organochlorine pesticide-contaminated soils with phyto-Fenton approach: A field study. Environ. Geochem. Health 2020, 42, 3597–3608.
  32. Gamallo, M.; Fernández, L.; Feijoo, G.; Moreira, M. Nano-based technologies for environmental soil remediation. In Nanomaterials for Sustainable Energy and Environmental Remediation; Elsevier: Amsterdam, The Netherlands, 2020; pp. 307–331.
  33. Liao, Y.; Yang, J. Remediation of vanadium contaminated soil by nano-hydroxyapatite. J. Soils Sediments 2020, 20, 1534–1544.
  34. Liu, Y.; Xu, K.; Cheng, J. Different nanomaterials for soil remediation affect avoidance response and toxicity response in earthworm (Eisenia fetida). Bull. Environ. Contam. Toxicol. 2020, 104, 477–483.
  35. Lv, Y.; Huang, S.; Huang, G.; Liu, Y.; Yang, G.; Lin, C.; Xiao, G.; Wang, Y.; Liu, M. Remediation of organic arsenic contaminants with heterogeneous Fenton process mediated by SiO2-coated nano zero-valent iron. Environ. Sci. Pollut. Res. 2020, 27, 12017–12029.
  36. Ahmad, A.; Ghufran, R.; Al-Hosni, T.K. Bioavailability of zinc oxide nano particle with fly ash soil for the remediation of metals by Parthenium hysterophorus. J. Environ. Health Sci. Eng. 2019, 17, 1195–1203.
  37. Sundararaghavan, A.; Mukherjee, A.; Suraishkumar, G.K. Investigating the potential use of an oleaginous bacterium, Rhodococcus opacus PD630, for nano-TiO2 remediation. Environ. Sci. Pollut. Res. 2020, 27, 27394–27406.
  38. Sarkar, A.; Sengupta, S.; Sen, S. Nanoparticles for soil remediation. In Nanoscience and Biotechnology for Environmental Applications; Springer: Cham, Switzerland, 2019; pp. 249–262.
  39. Wang, Y.; Liu, Y.; Zhan, W.; Zheng, K.; Lian, M.; Zhang, C.; Ruan, X.; Li, T. Long-term stabilization of Cd in agricultural soil using mercapto-functionalized nano-silica (MPTS/nano-silica): A three-year field study. Ecotoxicol. Environ. Saf. 2020, 197, 110600.
  40. Misra, M.; Ghosh Sachan, S. Nanobioremediation of heavy metals: Perspectives and challenges. J. Basic Microbiol. 2022, 62, 428–443.
  41. Abdi, O.; Kazemi, M. A review study of biosorption of heavy metals and comparison between different biosorbents. J. Mater Env. Sci. 2015, 6, 1386–1399.
  42. Ayangbenro, A.S.; Babalola, O.O. A new strategy for heavy metal polluted environments: A review of microbial biosorbents. Int. J. Environ. Res. Public Health 2017, 14, 94.
  43. Usman, M.; Farooq, M.; Wakeel, A.; Nawaz, A.; Cheema, S.A.; ur Rehman, H.; Ashraf, I.; Sanaullah, M. Nanotechnology in agriculture: Current status, challenges and future opportunities. Sci. Total Environ. 2020, 721, 137778.
  44. Abebe, B.; Murthy, H.A.; Amare, E. Summary on adsorption and photocatalysis for pollutant remediation: Mini review. J. Encapsulation Adsorpt. Sci. 2018, 8, 225–255.
  45. Desiante, W.L.; Minas, N.S.; Fenner, K. Micropollutant biotransformation and bioaccumulation in natural stream biofilms. Water Res. 2021, 193, 116846.
  46. Cao, X.; Alabresm, A.; Chen, Y.P.; Decho, A.W.; Lead, J. Improved metal remediation using a combined bacterial and nanoscience approach. Sci. Total Environ. 2020, 704, 135378.
  47. Gil-Díaz, M.; Diez-Pascual, S.; González, A.; Alonso, J.; Rodríguez-Valdés, E.; Gallego, J.; Lobo, M.C. A nanoremediation strategy for the recovery of an As-polluted soil. Chemosphere 2016, 149, 137–145.
  48. Baragaño, D.; Forján, R.; Welte, L.; Gallego, J.L.R. Nanoremediation of As and metals polluted soils by means of graphene oxide nanoparticles. Sci. Rep. 2020, 10, 1896.
  49. Mahmoud, M.E.; Abou Ali, S.A.; Elweshahy, S.M. Microwave functionalization of titanium oxide nanoparticles with chitosan nanolayer for instantaneous microwave sorption of Cu (II) and Cd (II) from water. Int. J. Biol. Macromol. 2018, 111, 393–399.
  50. Ha, C.; Zhu, N.; Shang, R.; Shi, C.; Cui, J.; Sohoo, I.; Wu, P.; Cao, Y. Biorecovery of palladium as nanoparticles by Enterococcus faecalis and its catalysis for chromate reduction. Chem. Eng. J. 2016, 288, 246–254.
  51. Mahmoud, M.E.; Abdou, A.E.; Mohamed, S.M.; Osman, M.M. Engineered staphylococcus aureus via immobilization on magnetic Fe3O4-phthalate nanoparticles for biosorption of divalent ions from aqueous solutions. J. Environ. Chem. Eng. 2016, 4, 3810–3824.
  52. Akhtar, N.; Khan, S.; Rehman, S.U.; Rehman, Z.U.; Khatoon, A.; Rha, E.S.; Jamil, M. Synergistic effects of zinc oxide nanoparticles and bacteria reduce heavy metals toxicity in rice (Oryza sativa L.) plant. Toxics 2021, 9, 113.
  53. Machado, S.; Pinto, S.; Grosso, J.; Nouws, H.; Albergaria, J.T.; Delerue-Matos, C. Green production of zero-valent iron nanoparticles using tree leaf extracts. Sci. Total Environ. 2013, 445, 1–8.
  54. Karn, B.; Kuiken, T.; Otto, M. Nanotechnology and in situ remediation: A review of the benefits and potential risks. Environ. Health Perspect. 2009, 117, 1813–1831.
  55. Gillham, R.W.; O’Hannesin, S.F. Enhanced degradation of halogenated aliphatics by zero-valent iron. Groundwater 1994, 32, 958–967.
  56. Roehl, K.E.; Meggyes, T.; Simon, F.; Stewart, D. Long-Term Performance of Permeable Reactive Barriers; Gulf Professional Publishing: Houston, TX, USA, 2005.
  57. Shih, Y.-h.; Tai, Y.-t. Reaction of decabrominated diphenyl ether by zerovalent iron nanoparticles. Chemosphere 2010, 78, 1200–1206.
  58. Elliott, D.W.; Lien, H.-L.; Zhang, W.-X. Degradation of lindane by zero-valent iron nanoparticles. J. Environ. Eng. 2009, 135, 317–324.
  59. Machado, S.; Pacheco, J.; Nouws, H.; Albergaria, J.T.; Delerue-Matos, C. Characterization of green zero-valent iron nanoparticles produced with tree leaf extracts. Sci. Total Environ. 2015, 533, 76–81.
  60. Shi, L.-n.; Zhang, X.; Chen, Z.-l. Removal of chromium (VI) from wastewater using bentonite-supported nanoscale zero-valent iron. Water Res. 2011, 45, 886–892.
  61. Trujillo-Reyes, J.; Peralta-Videa, J.; Gardea-Torresdey, J. Supported and unsupported nanomaterials for water and soil remediation: Are they a useful solution for worldwide pollution? J. Hazard. Mater. 2014, 280, 487–503.
  62. Wang, S.; Sun, H.; Ang, H.-M.; Tadé, M.O. Adsorptive remediation of environmental pollutants using novel graphene-based nanomaterials. Chem. Eng. J. 2013, 226, 336–347.
  63. Zhang, W. Nanoscale Iron Particles for Environmental Remediation: An Overview. J. Nanoparticle Res. 2003, 5, 323–332.
  64. Qiu, X.; Fang, Z.; Liang, B.; Gu, F.; Xu, Z. Degradation of decabromodiphenyl ether by nano zero-valent iron immobilized in mesoporous silica microspheres. J. Hazard. Mater. 2011, 193, 70–81.
  65. Xie, Y.; Cheng, W.; Tsang, P.E.; Fang, Z. Remediation and phytotoxicity of decabromodiphenyl ether contaminated soil by zero valent iron nanoparticles immobilized in mesoporous silica microspheres. J. Environ. Manag. 2016, 166, 478–483.
  66. Su, H.; Fang, Z.; Tsang, P.E.; Zheng, L.; Cheng, W.; Fang, J.; Zhao, D. Remediation of hexavalent chromium contaminated soil by biochar-supported zero-valent iron nanoparticles. J. Hazard. Mater. 2016, 318, 533–540.
  67. Gong, Y.; Liu, Y.; Xiong, Z.; Kaback, D.; Zhao, D. Immobilization of mercury in field soil and sediment using carboxymethyl cellulose stabilized iron sulfide nanoparticles. Nanotechnology 2012, 23, 294007.
  68. Mahanty, S.; Chatterjee, S.; Ghosh, S.; Tudu, P.; Gaine, T.; Bakshi, M.; Das, S.; Das, P.; Bhattacharyya, S.; Bandyopadhyay, S. Synergistic approach towards the sustainable management of heavy metals in wastewater using mycosynthesized iron oxide nanoparticles: Biofabrication, adsorptive dynamics and chemometric modeling study. J. Water Process Eng. 2020, 37, 101426.
  69. Govarthanan, M.; Jeon, C.-H.; Jeon, Y.-H.; Kwon, J.-H.; Bae, H.; Kim, W. Non-toxic nano approach for wastewater treatment using Chlorella vulgaris exopolysaccharides immobilized in iron-magnetic nanoparticles. Int. J. Biol. Macromol. 2020, 162, 1241–1249.
  70. Noman, M.; Shahid, M.; Ahmed, T.; Niazi, M.B.K.; Hussain, S.; Song, F.; Manzoor, I. Use of biogenic copper nanoparticles synthesized from a native Escherichia sp. as photocatalysts for azo dye degradation and treatment of textile effluents. Environ. Pollut. 2020, 257, 113514.
  71. Cheng, Y.; Dong, H.; Lu, Y.; Hou, K.; Wang, Y.; Ning, Q.; Li, L.; Wang, B.; Zhang, L.; Zeng, G. Toxicity of sulfide-modified nanoscale zero-valent iron to Escherichia coli in aqueous solutions. Chemosphere 2019, 220, 523–530.
  72. Cornu, J.-Y.; Huguenot, D.; Jézéquel, K.; Lollier, M.; Lebeau, T. Bioremediation of copper-contaminated soils by bacteria. World J. Microbiol. Biotechnol. 2017, 33, 26.
  73. Wu, G.; Kang, H.; Zhang, X.; Shao, H.; Chu, L.; Ruan, C. A critical review on the bio-removal of hazardous heavy metals from contaminated soils: Issues, progress, eco-environmental concerns and opportunities. J. Hazard. Mater. 2010, 174, 1–8.
  74. Huang, D.-L.; Zeng, G.-M.; Feng, C.-L.; Hu, S.; Jiang, X.-Y.; Tang, L.; Su, F.-F.; Zhang, Y.; Zeng, W.; Liu, H.-L. Degradation of lead-contaminated lignocellulosic waste by Phanerochaete chrysosporium and the reduction of lead toxicity. Environ. Sci. Technol. 2008, 42, 4946–4951.
  75. Xu, P.; Zeng, G.M.; Huang, D.L.; Lai, C.; Zhao, M.H.; Wei, Z.; Li, N.J.; Huang, C.; Xie, G.X. Adsorption of Pb (II) by iron oxide nanoparticles immobilized Phanerochaete chrysosporium: Equilibrium, kinetic, thermodynamic and mechanisms analysis. Chem. Eng. J. 2012, 203, 423–431.
  76. Polti, M.A.; García, R.O.; Amoroso, M.J.; Abate, C.M. Bioremediation of chromium (VI) contaminated soil by Streptomyces sp. MC1. J. Basic Microbiol. 2009, 49, 285–292.
  77. He, Y.K.; Sun, J.G.; Feng, X.Z.; Czakó, M.; Márton, L. Differential mercury volatilization by tobacco organs expressing a modified bacterial merA gene. Cell Res. 2001, 11, 231–236.
  78. Zhang, Z.Y.; Pan, L.P.; Li, H.H. Isolation, identification and characterization of soil microbes which degrade phenolic allelochemicals. J. Appl. Microbiol. 2010, 108, 1839–1849.
  79. Ryoo, D.; Shim, H.; Canada, K.; Barbieri, P.; Wood, T.K. Aerobic degradation of tetrachloroethylene by toluene-o-xylene monooxygenase of Pseudomonas stutzeri OX1. Nat. Biotechnol. 2000, 18, 775–778.
  80. Shrestha, B.; Acosta-Martinez, V.; Cox, S.B.; Green, M.J.; Li, S.; Cañas-Carrell, J.E. An evaluation of the impact of multiwalled carbon nanotubes on soil microbial community structure and functioning. J. Hazard. Mater. 2013, 261, 188–197.
  81. Fang, G.; Si, Y.; Tian, C.; Zhang, G.; Zhou, D. Degradation of 2, 4-D in soils by Fe3O4 nanoparticles combined with stimulating indigenous microbes. Environ. Sci. Pollut. Res. 2012, 19, 784–793.
  82. Tilston, E.L.; Collins, C.D.; Mitchell, G.R.; Princivalle, J.; Shaw, L.J. Nanoscale zerovalent iron alters soil bacterial community structure and inhibits chloroaromatic biodegradation potential in Aroclor 1242-contaminated soil. Environ. Pollut. 2013, 173, 38–46.
  83. Singh, J.; Lee, B.-K. Influence of nano-TiO2 particles on the bioaccumulation of Cd in soybean plants (Glycine max): A possible mechanism for the removal of Cd from the contaminated soil. J. Environ. Manag. 2016, 170, 88–96.
  84. Oyelami, A.O.; Semple, K.T. The impact of carbon nanomaterials on the development of phenanthrene catabolism in soil. Environ. Sci. Process. Impacts 2015, 17, 1302–1310.
  85. Xia, X.; Li, Y.; Zhou, Z.; Feng, C. Bioavailability of adsorbed phenanthrene by black carbon and multi-walled carbon nanotubes to Agrobacterium. Chemosphere 2010, 78, 1329–1336.
  86. Zhou, W.; Shan, J.; Jiang, B.; Wang, L.; Feng, J.; Guo, H.; Ji, R. Inhibitory effects of carbon nanotubes on the degradation of 14C-2, 4-dichlorophenol in soil. Chemosphere 2013, 90, 527–534.
More