Next Article in Journal
Dynamic Mechanical Behavior of Rock Specimens with Varying Joint Roughness and Inclination under Impact Load
Next Article in Special Issue
A New Technique for the Passive Monitoring of Particulate Matter: Olive Pollen Grains as Bioindicators of Air Quality in Urban and Industrial Areas
Previous Article in Journal
Bioactive Phenolic Compounds from Apples during Simulated In Vitro Gastrointestinal Digestion: Kinetics of Their Release
Previous Article in Special Issue
Hatchability and Survival of Lamproglena clariae Fryer, 1956 Exposed to Increasing Concentrations of Aqueous Aluminium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microbial-Based Heavy Metal Bioremediation: Toxicity and Eco-Friendly Approaches to Heavy Metal Decontamination

Urban and Rural Construction Institute, Hebei Agricultural University, Baoding 071001, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(14), 8439; https://doi.org/10.3390/app13148439
Submission received: 21 June 2023 / Revised: 14 July 2023 / Accepted: 18 July 2023 / Published: 21 July 2023
(This article belongs to the Special Issue Heavy Metal Toxicity: Environmental and Human Health Risk Assessment)

Abstract

:
There are several industrial processes in which heavy metals are used, including but not limited to chrome plating and tanning. Amongst the most toxic heavy metals to human health are arsenic, cadmium, chromium, lead, copper, nickel, and mercury. The aforementioned toxic metals possess the ability to cause contamination upon their release into the environment. Humans and aquatic and terrestrial animals are at risk from heavy metals in water and soil. Heavy metal toxicity has the potential to result in several health complications, such as renal and hepatic impairment, dermatological afflictions, cognitive lethargy, and potentially oncogenic manifestations. The removal of heavy metals from wastewater and soil can be accomplished using a variety of conventional methods, such as membrane filtration, reverse osmosis, chemical reduction, and adsorption. These methods have several disadvantages, such as generating an abundance of secondary pollutants, and entail significantly higher costs in comparison to biological methods. Conversely, eco-friendly techniques based on microbes have numerous advantages. This review provides a comprehensive overview of biological processes that remove heavy metal ions, both metabolically dependent and metabolically independent. Additionally, we also focused on the source and toxicity of these heavy metals. This study is expected to be particularly beneficial for the development of biological heavy metal treatment systems for soil and water.

1. Introduction

The dissemination of heavy metal contamination on a global scale has emerged as a primary concern due to the potential threats it poses to the environment. Heavy metals occur naturally within the Earth’s crust [1]. Nonetheless, due to their recalcitrant properties, they exhibit resistance towards degradation. These heavy metals have the potential to accumulate in humans and animals from soil and water by entering into the food chain [2]. Numerous natural and anthropogenic processes can introduce heavy metals into the environment. However, contemporary methods of agriculture, which involve an increasing dependence on agrochemicals and inorganic fertilizers, have resulted in the contamination of agricultural lands and have had disastrous consequences for ecosystems and the natural world [3]. In addition, heavy metals are introduced into water and soil by sewage sludge and organic waste manure, industrial byproducts, and wastewater irrigation [4]. Furthermore, the extraction of heavy metals from ores can pose significant environmental risks when certain areas are left unattended, and heavy metals may subsequently be displaced to other locations as a result of flooding or wind [5].
There has been extensive recognition of certain heavy metals that pose a health hazard to humans, such as arsenic (As), copper (Cu), chromium (Cr), lead (Pb), mercury (Hg), and cadmium (Cd). However, gold (Au) has not been found to pose such health hazards [6]. It is noteworthy that heavy metals, such as As, Hg, Cr, Cd, and Pb, typically possess densities of greater than 5 g/cm3. There are four primary pathways through which heavy metal ions are introduced into the human body, including the consumption of food that has been contaminated with heavy metals, the ingestion of water that has been contaminated with heavy metals, skin contact, and the inhalation of air that has been contaminated with heavy metals [6,7]. Heavy metal compounds are accountable for covalent bonds, which are known to be extremely hazardous in metalloid compounds. The propensity of heavy metals to covalently attach to organic groups is well documented, leading to the formation of lipophilic compounds or ions [8]. These metallic compounds are lipophilic, allowing them to effortlessly traverse cell membranes and gain entry into cells [9]. The interaction of these metallic compounds with cell organelles can result in toxic effects. The imperative requirement to eliminate heavy metals from polluted water and soil arises from their elevated toxicity [10].
Several technologies have been created to decrease the existence of detrimental heavy metals in water [11]. For the treatment of heavy metals, the most common approach is physicochemical, including filtration, ion exchange, reverse osmosis, desorption, and precipitation [12]. Precipitation, in particular, is widely acknowledged for its effectiveness in eliminating As, Cu, Cr, Pb, and Cd [12]. This method operates through the alteration of wastewater pH levels. It should be noted, however, that these physicochemical techniques are often associated with high costs and the generation of secondary chemical sludge [12,13]. In light of the limitations outlined above, it is of utmost importance to develop techniques that are both cost-effective and environmentally friendly, as these are essential to the removal of heavy metals from water and soil [14].
As a result of their commendable cost-effectiveness and environmentally sustainable nature, an extensive range of techniques and strategies for bioremediation have been identified, including, but not limited to, biosorption, bioreduction, bioaccumulation, mycoremediation, bacterial bioremediation, and phytoremediation [10]. Specifically, bioremediation has demonstrated an economical and ecologically sound approach to extracting heavy metals from contaminated sites [15]. In addition to rice and wheat husk, activated carbon, lignite, agricultural waste, bananas and citrus peels, and green-synthesized nanoparticles have been used as active biosorbents to effectively remove and eradicate heavy metal pollution [16]. In addition, it has been established through scientific research that living microorganisms, encompassing microalgae as well as bacteria, and plants are significantly efficacious in the role of bioremediation agents [10]. This is primarily attributed to their remarkable attributes of possessing high heavy metal tolerance properties, which have been demonstrated to be key determinants of their effectiveness in this regard [17]. The surface of bacterial cells is replete with a myriad of functional groups that are intimately involved in the binding of heavy metals, which may occur either to the cell membrane or in a reduced state. This intricate interplay effectively serves to mitigate the deleterious effects of heavy metal toxicity [18,19].
This review is focused on heavy metal contamination in water and soil, with natural and anthropogenic sources. Moreover, this review is also focused on the heavy-metal-mediated carcinogenicity and microbial-based techniques for the removal of certain heavy metals such as As, Cr, Cd, Pb, and Cu from contaminated water and soil.

2. Source and Toxicity of Heavy Metal Ions

2.1. Sources of Heavy Metals in the Environment

A vast and diverse array of industrial processes, encompassing a wide range of activities such as leather tanning, the electroplating of materials with Cr, the manufacture of batteries, the operation of glass industries, various agricultural activities, the disposal of domestic waste, and the conduct of pharmaceutical industrial processes, have all been conclusively identified as major contributors to the widespread and pervasive dissemination of heavy metals throughout the environment [14,20]. Toxic heavy metal ions are released into the environment, which poses a significant threat to both human health and the ecosystem. The production of Pb has resulted in the generation of roughly ten million tons of the element [21]. A substantial proportion of the total amount, precisely 85.10%, has been allocated to the battery industry for utilization. An additional 5.5% has been employed in the development of pigments. The remaining 2.1% has been directed towards miscellaneous industrial processes [4]. Furthermore, Pb is commonly utilized in petrol as tetraethyl and tetramethyl agents, which act as anti-knocking compounds. On the contrary, Cr finds its purpose in the production of steel, wood preservation, chrome plating, pigments, and electroplating [22]. Meanwhile, Cd is primarily employed in the electroplating and battery industry. To tackle the issue of heavy metal discharge in industrial effluent, environmental agencies such as the central pollution control board (CPCB), India, and the state environmental protection administration (SEPA), China, have laid down the maximum permissible limit for heavy metals. The maximum permissible limit of heavy metals in industrial effluent should be 0.2 mg/L (As), 2 mg/L (Cd), 0.1–2 mg/L (Pb), 2 mg/L (Cr), and 3 mg/L (Cu) [23]. The maximum permissible limit of heavy metals in industrial effluent according to the SEPA should be 0.1 mg/L (As), 0.01 mg/L (Cd), 0.05 mg/L (Cr6+), 0.1 mg/L (Pb), and 0.5 mg/L (Ni) [4,24]. Similarly, the United States Environmental Protection Agency (US EPA) has also established the maximum allowable limit of heavy metals present in drinking water as 0.05 mg/L (As), 0.05 mg/L (Cr), 0.1 mg/L (Pb), 0.01 mg/L (Cd), and 0.10 mg/L (Ni) [25]. Any transgressions beyond these permissible limits may result in multiple detrimental effects caused by heavy metals.

2.1.1. Natural Arsenic Sources

There exists a multitude of naturally occurring origins that contribute to the contamination of heavy metals in water and soil. The principal natural sources of these toxic heavy metals are found within geological formations, such as sedimentary deposits, volcanic rocks, and soils, as well as the combustion of vegetation and coal and volcanic eruptions [26]. Geothermal water is a significant contributor of heavy metals to groundwater, surface water, and soil. The concentration of As in the Earth’s composition varies, and the average concentration typically ranges from 1.5 to 5 mg/kg. As is commonly found in the sulphide ores of various metals, including silver (Ag), Cu, Au, and Pb. Heavy-metal-containing minerals such as pyrites, realgar, and orpiment have an important role in the groundwater contamination of heavy metals [27]. Phyllite ore has the highest arsenic concentration. Thereafter, some other major natural sources of arsenic are peaty soils and mudstones/marine shale. Higher heavy metal levels in groundwater are probable in sedimentary aquifers [28].

2.1.2. Anthropogenic Arsenic Sources

The anthropogenic sources of heavy metal pollution in soil, surface water, and drinking water have been reported worldwide. This type of heavy metal contamination in the water is related to human intervention, mining and coal processing, and petroleum refineries [29]. Heavy metal contamination through anthropogenic activity may be characterized into several types, such as coal-related and mining-based contamination [30]. The smelting and mining of heavy-metal-containing compounds cause heavy metals to leak into the surface water and groundwater by dumping and discarding these heavy-metal-containing waste materials [30]. The harmful impact of coal-mining-related heavy metal contamination has been extensively recorded and meticulously catalogued in an overwhelming prevalence of approximately 74 countries, spanning the globe [31]. The insidiousness of petroleum-based heavy metal contamination can scarcely be overstated, for it represents one of the most significant sources of heavy metal contamination, with its deleterious effects felt across a swath of approximately 17 countries across the world [32].

2.2. Heavy-Metal-Mediated Toxicity

2.2.1. Exposure to Heavy Metals and Environmental Pollution

ROS are generated in living cells when heavy metals bond to sulfhydryl groups and enter the cell. As a result, macromolecules are inactivated, ROS stress is caused, and glutathione levels are depleted [33]. When toxic metals enter the body, they initiate several processes, including interfering with metabolic pathways or inhibiting them. As a result, both humans and animals suffer from a variety of toxic effects of heavy metals [21]. A variety of problems may result, including organ dysfunction, metabolic abnormalities, hormone changes, congenital disorders, immune system problems, and cancer [34]. Metals in the environment, food, and water have therefore been subject to standards established by several international organizations. Food and water heavy metals are subjected to risk assessment studies [35]. Approximately 20% of 193 Ayurvedic medicines and famous Indian herbal medicines were found to contain Hg, Pb, and As, according to Saper et al. [36]. Pb, Hg, and As were detected in nearly 21% of samples [36]. The level of acceptable daily metal intake was exceeded by all metal-containing products. This would cause Pb and/or Hg levels to exceed acceptable levels by 100 to 10,000 times [37]. The presence of heavy metals in dietary supplements is a matter of considerable concern, and as such, these potentially toxic substances must be subjected to daily dosage limits as established by regulatory authorities [38]. It is important to note that the establishment of such dosage limits is crucial for ensuring the safety and well-being of individuals who consume such supplements and that any deviation from these limits could result in severe health consequences. Thus, manufacturers and regulatory authorities must work together to ensure that the levels of heavy metals in dietary supplements are strictly controlled and monitored to minimize any potential risks. Ultimately, the establishment of and adherence to daily dosage limits for heavy metals in dietary supplements is a critical component of ensuring public health and safety and should be given the utmost priority by all stakeholders involved in the production and regulation of such products [39]. Manufacturers must assess these standards as well [40]. The phenomenon of environmental pollution has a profound impact on areas that possess an unspoiled quality. This can be observed in the case of Mount Everest, a location that is isolated and remote [38,39,40]. There have been reports of heavy metal contamination in this area associated with Pb, Cd, Cr, As, and Hg [41]. As and Cd levels in all snow samples collected from Mount Everest have exceeded the drinking water guidelines outlined by the US EPA, according to Yeo and Langley-Turnbaugh [42]. Moreover, a substantial amount of As contamination was detected in soil samples taken from the heavy-metal-contaminated sites located on Mount Everest.

2.2.2. Heavy Metals Binding to the Thiol Group

Upon entering the body, heavy metals form bonds with proteins, nucleic acids, and lipids with a particular tendency to form bonds with lipids and proteins. In proteins, cysteine residues are modified by thiol (-SH) groups when metals bond with enzymes and proteins [43]. It is important to note that such inhibition of protein function can ultimately lead to the disruption of the intracellular redox equilibrium [44]. Consequently, an impaired antioxidant defense mechanism can contribute to the onset of hepatocellular injury. It is worth mentioning that heavy metals can elicit comparable reactions with proteins [44]. In the course of the intricate process of complexation that involves heavy metals and thiol-containing proteins, it is noteworthy that the ligands which participate are none other than amino acids that are characterized by the presence of the -SH functional group [44,45].

2.2.3. Heavy-Metal-Mediated Carcinogenicity

The complex and ambiguous nature surrounding the etiology of carcinogenicity in heavy metals has been widely postulated [46]. It has been put forth as a proposition, with a high degree of specificity, that the carcinogenic properties of certain heavy metals are attributable to their binding to regulatory proteins that perform an exceedingly crucial function in the regulation of the cell cycle and DNA synthesis and repair mechanisms, as well as the highly intricate process of apoptosis [46,47]. As a result of toxicogenomic investigations, gene expressions following metal toxicity have been characterized. Activator protein 1 (AP-1) and nuclear factor-kappa B (NF-KB) are two transcription factors that are targeted by Cd and As [48]. Uncontrolled cell growth and division are the result of failure to control protective gene expression. Several studies have examined the carcinogenic effects of heavy metals and mutations in Ras proteins [48,49]. Human prostate epithelial cells exposed to As exhibited the overexpression of Ras, as demonstrated by Ngalame et al. [49]. Furthermore, in vitro analysis has uncovered that Cd triggers the augmentation of the extracellular signal-regulated kinase 1/2 (ERK 1/2) level, as well as the jun and fos transcription factors. Correspondingly, Cr (VI) has been discovered to instigate the overexpression of jun in cultured cells [49]. Consequently, the mutated Ras protein loses its ability to undergo inactivation, and the kinase cascade remains unturned. Furthermore, the heightened jun and fos or activated ERK 1/2 sustains the gene expression, resulting in the activation of an incessantly active signaling pathway, which ultimately culminates in the continuous activation of proliferation and the increased formation of tumors [50,51].

3. Heavy Metal Removal Approaches

The notion of utilizing biological means to remove heavy metals is exceedingly appealing when juxtaposed with other conventional techniques because biological procedures are characterized by their cost-effectiveness, eco-friendly nature, and their exceptional efficiency when it comes to dealing with low concentrations of heavy metal ions in wastewater [52]. A wide range of heterogeneous biological entities, including various substances such as plant biomass, agricultural detritus, microbial biomass, nanoparticles synthesized through green methods, residual fruit matter, and biopolymers, have been effectively employed in the removal of heavy metals [53]. These aforementioned agents have demonstrated their efficacy in this role, and their use has been widely documented in the literature [54]. It is significant to acknowledge that the application of these biological agents for heavy metal elimination holds immense importance within the realm of environmental science. This is because it presents a sustainable and environmentally conscious approach towards tackling the problem of heavy metal pollution [52,54]. Thus, further research into the application of these agents in heavy metal removal is warranted and is likely to yield valuable insights and contributions to the field. Living organisms, which are vital constituents of the biosphere, including bacteria, algae, and fungi, have been identified as playing significant and emerging roles in the process of heavy metal removal from the environment [55]. Recent advances in biotechnology have enhanced the understanding of the mechanisms through which these organisms mediate the detoxification of heavy metals. These mechanisms include biosorption, biomineralization, bioaccumulation, and bioremediation and are crucial in the removal of heavy metals from ecosystems [10,56]. Therefore, it is evident that the utilization of living organisms in the removal of heavy metals is a promising approach towards achieving sustainable environmental management. Various types of microorganisms, including fungi, bacteria, and algae, can absorb heavy metals from their surrounding environment and transport them into their cells [57]. It has been documented and reported in numerous studies that a diverse range of microbial species possesses the unique capability to convert the extremely hazardous Cr (VI) to a less toxic form of Cr (III). Microbial cells possess the ability to effortlessly internalize Cr (III) owing to their limited solubility in aqueous solutions and negligible toxicological characteristics [58,59]. Numerous techniques for heavy metal bioremediation have been proposed in the past, including biosorption, phytoremediation, bioreduction, and bioaccumulation [58]. Biosorption involves the use of dead biomass, while phytoremediation necessitates the removal of heavy metals through plant mediation. On the contrary, it is imperative to note that bioreduction is a highly intricate process that encompasses the transformation of the oxidation states of heavy metal ions, which is a highly significant phenomenon in the field of environmental chemistry [60]. Conversely, bioaccumulation is a process that is characterized by the absorption and uptake of heavy metal ions into the intracellular space, which is a crucial mechanism that is fundamental to the understanding of the behavior of heavy metals in biological systems. These techniques have been extensively researched and have shown promise in the bioremediation of heavy metals [61]. The process of remediating heavy metals can be achieved through the utilization of two distinct mechanisms: metabolically independent mechanisms, which entail the use of inorganic, lifeless materials, and metabolically dependent mechanisms, which involve the use of living cells of bacteria, fungi, and algae that are capable of carrying out the necessary metabolic processes required to transform the heavy metal contaminants into less harmful forms [4,62]. In the context of this current evaluation, the esteemed authors have decidedly and intently directed their concentration towards the intricate and multifaceted process of the bioremediation of heavy metals, which is undeniably and inextricably linked to the intricate and complex workings of metabolic pathways and processes.

3.1. Microbial Mechanisms Involved in Heavy Metal Bioremediation

Bioremediation is a biological technique that is employed to extract heavy metal ions from sites that have been contaminated with heavy metals [63]. In bioremediation, absorption/adsorption, bioaccumulation, biotransformation, and bioleaching are included in the remediation of toxic metal ions. There are several natural components, such as bark, coconut husk, rice husk, wheat husk, seaweeds, seeds, aquatic plants, agro-waste, and microorganisms, which are utilized to reduce toxic heavy metal concentrations in polluted sites. In these materials, microorganisms play a significant role in bioremediation [64]. The phenomenon of microorganisms altering the ionic state of heavy metals is a significant process that has far-reaching implications for their solubility, bioavailability, and movement in both soil and aquatic environments [65]. The mobilization and immobilization of heavy metals is a crucial aspect of microbial remediation, a complex process that involves a multitude of intricate mechanisms including, but not limited to, oxidation–reduction, chelation, modification of the metallic complex, and biomethylation [63]. It should be emphasized that microbial enzymatic catalysis plays a pivotal role in the solubilization of metals with higher oxidation states to lower oxidation states. For example, the enzymatic oxidation of uranium by Thiobacillus ferrooxidans and T. thiooxidans serves as a prime illustration of this phenomenon [66]. Although the isolation of microorganisms responsible for the degradation of heavy metals from both aerobic and anaerobic environments is feasible, aerobic microbes are considered to be more suitable for bioremediation when compared to anaerobic microorganisms [67]. Furthermore, microorganisms utilize membrane-linked transport mechanisms to transport heavy metals and convert them into non-hazardous forms, which is of utmost importance for their survival in a metal-polluted environment [68]. In this regard, microorganisms employ a range of processes, such as biosorption, bioaccumulation, biotransformation, and bioleaching, to counteract the detrimental effects of heavy metals and ensure their sustainability [68,69].
It is of great importance to note that the utilization of the aforementioned techniques has proven to be efficacious in facilitating the process of environmental clean-up, thereby emphasizing the substantial capacity possessed by microorganisms in ameliorating the deleterious impact of heavy metals on the ecosystem [70].

3.2. Biosorption

The phenomenon of biosorption, which is frequently defined as the process of adsorption of heavy metal ions by biomass, is considered to be metabolically independent heavy metal adsorption [71]. This captivating process is generally executed by a diverse range of deceased biomass, encompassing but not limited to plant waste, microalgae, agro-waste, microbial biomass, and agriculture residue [72]. In the course of biosorption, it is widely acknowledged that heavy metal ions tend to adhere to the surface of the biomass using a series of functional groups that are present on said surface. It is, however, worth noting that biosorption can also be executed by live cell biomass. Specifically, living microbial cells possess the exceptional capacity to attach heavy metal ions to their cell surface through the process of passive adsorption and complexation [73].
Biosorption, an intriguing phenomenon for researchers in the field, entails a multifaceted process whereby heavy metal ions adhere to surface functional groups, namely amino, amide, imidazole, sulfonate, and carboxyl. It is noteworthy to emphasize that the pKa value of the medium in which biosorption takes place exerts a substantial impact on the adsorption of heavy metal ions and the binding propensity of functional groups. It is believed that hydroxyl groups have a pKa value between 9.5 and 13.21. As a result, carboxylate groups are mostly responsible for the negative charges on biomass between these Pka values. Aqueous media requires the deprotonation of carboxylic groups to generate negative charges and H+. As a result, different metal ions can be adsorbed [74]. One of the most crucial and indelible aspects of biosorption is the judicious selection of a suitable material to employ. This material should be easily obtainable, non-hazardous, and economically viable, which represents an important consideration for researchers in the field [75].
The efficacy evaluation of a biosorbent is contingent upon the array of surface groups that are present. For a biosorbent to be deemed optimal, the concentration of surface functional groups must be augmented [25]. The surface morphology of an adsorbent represents a pivotal characteristic in the heavy metal biosorption process. Surfaces that are rough and porous provide a larger surface area, thereby conferring an advantage for the binding of heavy metal ions onto the biosorbent surface, as per reference [76]. It is crucial to carefully scrutinize the surface morphology and functional groups of a biosorbent. A myriad of techniques, including Fourier Transformation Infra-Red (FTIR), Scanning Electron Microscopy (SEM), Energy Dispersive X-ray Analysis (EDX), Nuclear Magnetic Resonance (NMR), and X-ray Diffraction (XRD) are readily available to characterize a biosorbent, as per reference [77]. As such, it is imperative to employ these aforementioned techniques judiciously to obtain a comprehensive understanding of the intricate details of the biosorbent’s surface morphology and functional groups [76,77].
The biosorption process, which involves the use of biomass to remove metal ions from aqueous solutions, is influenced by various factors that impact its efficacy. These factors encompass the morphology and composition of the biomass, the presence of multiple metal ions in the media, the ambient temperature, and the pH of the medium [78]. It is commonly observed that a decrease in pH results in competition between cationic heavy metal ions. Conversely, an increase in pH leads to the deprotonation of the biosorbent surface, which causes the exposure of surface functional groups [79]. Additionally, desorption, which involves releasing adsorbed heavy metals from the biosorbent, can be employed to regenerate the biosorbent. The recovery of metal ions can be achieved by manipulating the pH of the medium, which is a critical factor in the process [80]. Moreover, the biosorption mechanism of Cr (VI) is a complex phenomenon in which the Cr (VI) ion attaches to the biosorbent surface and is subsequently reduced to less toxic Cr (III) [80].
The process of reducing and adsorbing Cr (VI) proceeds in a sequence of three distinct stages, which have been observed and studied extensively. The initial stage involves the binding of chromate ions, which are negatively charged, to positively charged functional groups [81]. This binding process is an essential step in the overall reduction-cum-adsorption process, as it enables the subsequent stages to take place. In the second stage, Cr (VI) is reduced to Cr (III) by functional groups, which donate electrons to facilitate the reduction process [82]. This stage is crucial in the sense that it marks a significant transformation in the chemical composition of the system, as Cr (VI) is converted to Cr (III). The ultimate stage in the process entails the discharge of Cr (III) ions into the solution, while the residual Cr (III) ions are retained by the adsorbent surface, which is adorned with negatively charged functional groups. This phase is a fundamental requisite in guaranteeing the consummation of the adsorption process and the total occupation of the Cr (III) ions on the adsorbent surface [83]. It is noteworthy that this process represents a complex interplay of various chemical and physical phenomena, which require further investigation and elucidation [84]. Table 1, which is effectively and efficiently utilized as a visual representation of data, illustrates, in a comprehensive manner, the inherent capabilities of different biosorbents to uptake heavy metal ions.
The microbial biomasses, as previously mentioned, have been identified as remarkably efficient biosorbents when it comes to removing heavy metal ions. Additionally, it is noteworthy that the biosorption phenomenon can be conducted on immobilized biomass. This includes bacterial biomass that exists on a solid surface, fungal biomass located on a supportive medium, or deceased biomass that has been immobilized on another supportive surface [92]. Immobilization serves to not only augment the stiffness, potency, and longevity of the biosorbent but also ultimately to enrich the biosorption capacity. The existence of a diverse array of matrices has made immobilization possible, with matrices such as alginate and polyurethane being extensively employed for this purpose [92,93]. Moreover, it is noteworthy to emphasize that a multitude of microbial and plant biomasses have been employed for immobilization on the aforementioned matrices [91,94]. To illustrate, the biomass of Chlorella homosphaera has been successfully immobilized on matrices containing sodium alginate [95].

3.3. Bioaccumulation

Bioaccumulation is an intricate biological phenomenon that involves the accumulation of heavy metal ions within microbial cells. It is a process that occurs at a slower rate in comparison to biosorption, as it entails the active participation of several metabolic pathways [61]. In contrast to the phenomenon of biosorption, characterized by the adsorption of metal ions onto the exterior surface of microbial cells, the process of bioaccumulation entails the uptake and subsequent retention of said metal ions by living microbial cells suspended within a medium [96]. Furthermore, the bioaccumulation process necessitates the persistent monitoring and observation of said cells. Owing to the multifaceted nature of this process, a thorough understanding of the underlying mechanisms is crucial to comprehend the impact of bioaccumulation on the environment and its potential applications in bioremediation [96]. The process of bioaccumulation presents an advantage in reducing various stages of biosorbent preparation, such as biomass collection, drying, preparation through washing and crushing, and storage. However, it is subject to significant influence by experimental conditions, as reported in [75]. The presence of pollutants in wastewater poses a substantial challenge to microbial growth and results in competition with heavy metals selected for bioaccumulation. Furthermore, these pollutants can bind to the bacterial outer surface and disrupt cell heavy metal accumulation, leading to further challenges in the bioaccumulation process [97]. It is noteworthy that numerous species of fungi, bacteria, algae, and plants have been identified as playing an essential role in the elimination of heavy metals, thus reinforcing the importance of considering biotic factors when assessing the effectiveness of bioaccumulation [98,99].
Among the various living systems that exist, it has been observed that microbes exhibit a noteworthy capacity for both effectively removing heavy metals as well as demonstrating resistance towards them [63,100]. Bacteria, which are single-celled microorganisms, have been observed to exhibit the extraordinary ability to selectively assimilate heavy metal ions through the utilization of cell surface receptors, which are specialized proteins that are embedded within the outer membrane of the cell. Following the initial binding of these heavy metals to the receptors, the bacteria have been observed to efficiently accumulate and store them within their cellular structures, thereby enabling them to maintain their physiological functions under adverse environmental conditions [101]. Notably, certain bacterial cells, such as those belonging to the Pseudomonas, Klebsiella, and Microbacterium species, have exhibited a remarkable ability to resist heavy metal ions that have been extracted from diverse contaminated sites. It has been observed that bacteria that have been isolated from areas that have been contaminated with heavy metals have generally developed a resistance to these ions [102]. Furthermore, numerous researchers have successfully identified strains of Microbacterium that exhibit heavy metal resistance within such contaminated regions. Henson and colleagues [103] conducted the isolation of Microbacterium sp. (Cr-K29), which exhibited the remarkable ability to reduce up to 88% of Cr (VI) from water that has been contaminated with 2 mM of Cr (VI). In the interim, Pattanapipitpaisal and colleagues [104] discovered that Microbacterium liquefaciens, when anchored to solid surfaces, has the capability of eliminating 90–95% from a 50 µM Cr (VI) solution. It is of utmost importance to take note that bacteria have various mechanisms at their disposal to rid heavy metals from their surroundings, as these metals are detrimental to living organisms [105]. These microorganisms employ heavy metal ions to facilitate their metabolic activities, or they employ soluble enzymes that are generated by the bacterial cell in order to detoxify heavy metal ions [106]. The complete bioremediation process of Cr (VI) by bacteria is depicted in Figure 1.
Reactive oxygen species (ROS) are known to be generated in bacterial cells upon exposure to toxic heavy metals. Such ROS have the potential to inflict damage upon the cell organelles or cause the disruption of multiple metabolic functions, ultimately leading to aberrations in normal cell functions [107]. As a consequence, the antioxidant system within the microbial cells, including enzymes like superoxide dismutase (SOD), glutathione-S-transferase (GST), and catalase (CAT), plays an active role in combating the stress induced by ROS generation in bacterial cells [108]. Researchers have made a noteworthy discovery that several bacterial species possess the capability to exhibit resistance towards heavy metals [109]. In Table 2, the bioremediation efficacy of heavy-metal-tolerant bacterial strains has been meticulously outlined.

3.4. Bioleaching

Bioleaching is an exceedingly multifarious and intricate process that involves the utilization of an extensive range of microbial organisms, with a special emphasis on acidophiles. These remarkable chemolithotrophic microorganisms exhibit an exceptional ability to oxidize Fe (II) to Fe (III) and/or reduce sulfur to sulfuric acid, which endows them with the capacity to thrive in environments with low pH levels typically ranging at or below 2.0 [112,113]. The solubilization of metal sulfides and oxides from ores, thereby enabling the extraction of metal through the segregation of metals in the solid phase from the more water-soluble phase, is accounted for by the production of ferric ions and protons resulting from sulfuric acid [114]. Bioleaching, which involves the reduction of metal ions by microorganisms, plays an indispensable role in the extraction and recovery of heavy metals [115]. The ability of microorganisms to convert solid chemicals within contaminated soil into a soluble substance that can be removed and recovered significantly affects the efficacy of the recovery process, especially given the finite and non-renewable nature of metal resources [116]. In light of this, scholars have proposed bioremediation as a viable means of recovering raw materials from effluent. For instance, an absorbent based on Annona squamosa was utilized to achieve a substantial recovery rate of up to 98.7% for Cd (II) through 0.1 M HCl [117]. Likewise, it is noteworthy to mention that the application of Pseudomonas aeruginosa biomass, in conjunction with a highly concentrated solution of hydrochloric acid (0.1 M HCl), has been observed to yield a remarkably substantial rate of retrieval for the elusive heavy metal known as cadmium (II), with recovery rates reaching an impressive 82% [118]. Additionally, a noteworthy 100% recovery rate of Cu (II) was reported by utilizing volcanic rock matrix-immobilized P. putida cells that possessed surface-displayed cyanobacterial metallothioneins at a pH of 2.35. Analogously, 100% recovery of Cu (II) was obtained from activated sludge at a pH of 1.0 [119]. It is a matter of significance to highlight that the incorporation of the autochthonous variant Enterobacter sp. J1 brought about a remarkable increase in the recuperation percentage, exceeding 90% for both copper and lead ions when the pH level was at 2 [120].

3.5. Biotransformation

Biotransformation, a crucial process in the field of chemistry, refers to the alteration of a molecule’s chemical structure, ultimately leading to the creation of a comparatively more polar molecule [121]. To put it differently, the interaction between metals and microorganisms induces a change in toxic metals and organic compounds, converting them into a less hazardous form. This mechanism plays a significant role in facilitating heavy metal adaptation by microorganisms [97]. It is worth noting that microbial transformations can be accomplished through a variety of methods, including carbon bond formation, isomerization, functional group introduction, oxidation, reduction, condensation, hydrolysis, methylation, and demethylation, among others [122]. The process of altering metals utilizing microbial organisms has been extensively recorded. The investigation has brought to light the fact that Micrococcus sp. and Acinetobacter sp. possess the remarkable capability to expedite the oxidation of the most toxic As (III) into compounds such as arsenic acid and arsenates that are comparatively less soluble and non-poisonous As (V), consequently mitigating its detrimental impact [123]. According to the findings of the study carried out by Thatoi et al. [116], the deployment of NADH-dependent reductase by Cr (VI)-tolerant Bacillus sp. SFC 500-1E has been observed to be highly effective in reducing the hazardous Cr (VI) to less toxic Cr (III) [124].

3.6. Some Other Advanced Microbial-Based Approaches for Heavy Metal Removal

3.6.1. Rhizoremediation

The process of rhizoremediation, which is a method employed to eliminate heavy metals from the soil, involves the utilization of microorganisms that are found in the rhizosphere of plants. Rhizoremediation is a technique that combines phytoremediation and bioaugmentation, which are two techniques used to cleanse contaminated substrates [125]. Rhizoremediation is a highly innovative and cutting-edge approach that has been developed to address the pressing issue of organic pollutants and heavy metal contaminants that are present in our environment [126]. This method is based on the use of plants and rhizospheric microorganisms, which have proven to be highly effective in the process of restoring polluted environments. Indeed, the success of rhizoremediation cannot be overstated, as it is an eco-friendly, cost-effective, and highly efficient means of dealing with the problem of pollution [126]. However, it is important to note that the success of rhizoremediation at any given site is dependent on a wide range of factors, including the level of soil contamination and the quantity of metal contaminant present in the soil, as well as the ability of plants to aggressively take up metals from the soil. Therefore, these factors must be carefully considered and taken into account when implementing rhizoremediation to ensure that the process is as effective as possible [127]. Ultimately, the use of rhizoremediation represents a crucial step forward in the fight against pollution, and we must continue to explore and develop this innovative approach to safeguard our environment for future generations [68].
Plants that are employed in the process of rhizoremediation can be categorized into two distinct groups, namely the hyperaccumulators and non-hyperaccumulators. The former group comprises plants that possess an exceedingly high ability to accumulate heavy metals, albeit at the expense of biomass efficiency [128]. Conversely, the latter group has a comparatively lower capacity for extraction of heavy metals when compared to hyperaccumulators, but their total biomass yield is considerably higher. Moreover, non-hyperaccumulators exhibit rapid growth characteristics [129]. In the context of removing heavy metals from contaminated soils, certain plants utilize a variety of processes, which can be observed in Figure 2.
Phytostabilization is a process which involves the immobilization of heavy metals in soil by the adsorption or precipitation of heavy metals. The term phytoextraction is related to the translocation of heavy metal ions from the soil and their accumulation in plants (root, stem, and leaves). In phytovolatilization, heavy metals are absorbed by plant roots and pushed up through the xylem stream. The heavy metal ions transfer to the upper parts of the plant, such as the leaves and stem, then convert into a volatile form and are released into the atmosphere [130,131]. This scientific approach has been integrated with several other techniques to create a comprehensive method for soil renewal. Currently, the application of plants and plant-growth-promoting bacteria (PGPB) is being evaluated as an effective and environmentally acceptable method for soil renewal [132]. PGPB in the soil that are close to the roots of plants play a vital role in facilitating the absorption or adsorption of heavy metals by plants. In addition to enhancing remediation abilities, PGPB act as a plant growth stimulant by secreting hormones and metabolites. This stimulates growth, solubilizes minerals, fixes nitrogen, and protects against pathogens. Other biotic and abiotic stresses that plants face can also be relieved by PGPB. It is important to note that the process of rhizoremediation is not only environmentally friendly but also enhances soil quality and promotes the growth of healthy plants [133,134].

3.6.2. Genetically Engineered Organisms

The utilization of microorganisms for bioremediation has proven to be an efficacious technique for alleviating soil and water contamination caused by heavy metals. Microorganisms possess the capacity to decompose intricate substances into simpler chemicals, thereby offering a long-lasting solution for the mitigation of heavy metal contamination [63,66]. Recent advances in the field of genetic engineering have facilitated the creation of genetically engineered microorganisms or biocatalysts, which have been demonstrated to outperform natural microbes in terms of their aptitude for eliminating persistent compounds from natural surroundings [135]. Various genetic and metabolic engineering strategies have been utilized to produce highly efficient engineered microorganisms. These strategies consist of single-gene editing and pathway construction, as well as modifications of both coding and controlling sequences of pre-existing genes [136]. The central emphasis of the aforementioned adjustments is primarily on the stages that restrict the rate of metabolic processes. It is noteworthy that the employment of genetically modified bacteria has facilitated the efficient elimination of heavy metals, including but not limited to Cd, As, Cu, Hg, and Ni [137]. However, it is crucial to acknowledge that the speed at which degradation takes place is contingent upon the catalytic efficacy of enzymes that are either present within the cells or have been stimulated to act on a specific substrate [6].
The utilization of recombinant DNA technology, a genetic engineering tool, has been implemented to amplify the genome of genetically engineered microorganisms (GEMs) by integrating foreign genes from other species [138]. The development of genetically engineered metabolic pathways in microbial cells through bioremediation processes has been enhanced as a result. The advanced technological nature of GEMs has attracted public attention [139]. The incorporation of metal regulatory genes in bacteria has facilitated the conversion of heavy metals from toxic forms to less toxic ones [140]. Additionally, heavy metal accumulation has been observed to increase in GEMs that express metallothioneins (MT) [141]. Azad et al. [142] conducted a comprehensive evaluation of the application of genetically modified bacteria and plants in the bioremediation of heavy metals and other organic pollutant-contaminated environments [142]. Furthermore, Nascimento and Chartone-Souza [143] have demonstrated that the integration of the mer operon from bacterial resistance to mercury, which encodes for the reduction of Hg2+, into the genetically modified bacterium affords the mer genes the capacity to mitigate Hg pollution at high temperatures [143].
The genes that confer resistance to heavy metals, namely alkB, alkB1, alkB2, alkM, xylE, nidA, and ndoB, have been observed to be harbored on plasmids within n-alkane-degrading microorganisms. These genes, which have been frequently employed as markers for identifying microbial biodegradation, are known to undergo horizontal transfer [144].
Within the realm of environmental bioremediation, the genetic modification of microbial membrane transporters serves as a significant means of enhancing the removal of hazardous metals [145]. The crucial roles of transporters and binding mechanisms in the remediation of such metals cannot be overstated. Typically, transporter systems are categorized into three principal categories, namely channels, secondary carriers, and primary active transporters. Channel transporters, which are a subset of the aforementioned categories, are defined by their ability to facilitate the passage of molecules across the inner lipid membrane [62,146]. Examples of channel transporters include Fps, Mer T/P, and GlpF. Additionally, secondary carriers, such as Hxt7, NixA, and Pho84, can be found in the inner lipid membrane. Finally, primary active transporters, including cdtB/Ip_3327, MntA, TcHMA3, and CopA, are also situated in the inner lipid membrane. It is noteworthy that certain transporters, such as the porin channels, may also be found in the outer lipid membrane [66,147]. This is an important consideration to keep in mind when studying the intricate mechanisms of transporter systems. Once hazardous metals infiltrate the cell, a myriad of phytochelatins, metallothioneins, and polyphosphates work in harmony to sequester the metals [148]. Altering the hazardous metal import–storage systems of microorganisms could potentially augment their proficiency in extracting hazardous metals from water and soil. In the ongoing battle against the deleterious compounds that pervade the environment, it has become increasingly apparent that the utilization of genetically engineered microorganisms constitutes an indispensable tool in expediting the restoration process [149]. Such organisms, endowed with the capacity to efficiently degrade and neutralize the toxic substances present in the environment, have demonstrated their efficacy in mitigating the deleterious impact of these compounds on the ecosystem [150].
Recent research has brought to light recently discovered genetic elements that play a crucial role in the degradation of a broad spectrum of hazardous material contaminants. The elucidation of these genes was made possible through the application of high-throughput and next-generation sequencing techniques [151]. In addition, the advent of innovative technologies such as CRISPR-Cas gene-editing tools has provided an avenue for the optimization of bioremediation. By employing CRISPR technology, heavy metal remediation genes can be transported to plants or bacteria and facilitate heavy metal detoxification. Metallothioneins and phytochelatins can be synthesized more efficiently with CRISPR-mediated gene expression [152]. The use of CRISPR-Cas9 gene-editing technology has been demonstrated in numerous studies to enhance metal remediation by transferring certain plant and bacterial genes. In Arabidopsis and tobacco plants, the NAS1 gene (encoding the nicotianamine (NA) synthase enzyme) has been shown to improve tolerance to harmful metals, such as Cu, Cd, Ni, Mn, Fe, Mn, and Zn [153]. By genetically modifying microorganisms with genes that are intimately involved in the breakdown of particularly stubborn substances, the bioremediation process can be vastly improved [152,154].

4. Technology Challenges and Future Prospects

The process of remediating heavy metals presents a significant challenge that requires a considerable amount of time and effort [155]. Furthermore, one of the primary concerns is the successful implementation of microbial-based heavy metal removal techniques at a commercial level, specifically in water treatment plants. There exist diverse strategies that can be employed to enhance the removal of heavy metals from aqueous media. These include the manipulation of the immobilization of bacterial strains onto solid substrates and the utilization of composite nanomaterials or biopolymers. Moreover, genetically engineered microorganisms have better heavy metal removal potential in comparison to wild-type microorganisms. It is better to scale up the implementation of these microbes into wastewater treatment plants. It is noteworthy to mention that biological remediation methods have the potential to eliminate other toxic metal ions, including Ni, V, Hg, and Ti, which can be further explored in future research. Moreover, there is a need to implement CRISPR-mediated gene expression in the microbes and plants used in the heavy metal remediation process and their scale-up for wastewater treatment plants. Growth media, optimum growth conditions, and culture maintenance are the major technological changes in microbial-based heavy metal removal.
In addition, phytoremediation is a good method for the removal of heavy metals from contaminated water but there needs to be further research on the safe disposal of the used plant. In addition, advanced technical assessment tools such as the life cycle assessment (LCA) and techno-economic analysis of large-scale wastewater treatment plants can be beneficial in developing cost-effective, energy-saving, and eco-friendly systems in the future. Furthermore, it is of great value to contemplate the molecular mechanisms of microbes underlying the elimination of heavy metals and conduct a comparative transcriptomic analysis in prospective research agendas. As a result, it is imperative to persistently investigate and devise novel strategies to enhance the efficacy of heavy metal elimination from water, all the while taking into account the long-term viability and ecological implications of the remediation process.

5. Conclusions

Various natural and industrial processes release heavy metals, which result in the contamination of both water and soil. The presence of heavy metal ions in these two mediums has detrimental effects on human and animal health. The toxicity of these metals has been linked to a wide range of health issues, including but not limited to liver and kidney damage, skin conditions, mental impairment, and even cancer. This highlights the severity of the matter at hand. Given the observed toxic effects of heavy metals, it is imperative to implement eco-friendly biological methods for their removal. The microbial-based techniques outlined in this review have the potential to effectively decontaminate both soil and water. These methods, including biosorption, bioaccumulation, and rhizoremediation, are deemed to be eco-friendly, efficient, and cost-effective. This study is expected to be particularly beneficial for the development of microbial-based heavy metal removal methods that can significantly contribute to the protection of human and environmental health.

Author Contributions

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

Funding

This research was funded by the Key R&D Project of Hebei Province (R&D and Demonstration of Key Technologies for “Zero Carbon” Disposal of Rural Habitat Manure), grant number 22323802D, and the APC was funded by 22323802D.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are mentioned in this manuscript.

Acknowledgments

The authors are grateful to the Hebei Agricultural University for necessary support in this research work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, F.; Dong, Z.Q.; Wang, C.C.; Wei, X.H.; Hu, Y.; Zhang, L.J. Heavy Metal Contamination of Soils and Crops near a Zinc Smelter. Huan Jing Ke Xue 2017, 38, 4360–4369. [Google Scholar]
  2. Okereafor, U.; Makhatha, M.; Mekuto, L.; Uche-Okereafor, N.; Sebola, T.; Mavumengwana, V. Toxic Metal Implications on Agricultural Soils, Plants, Animals, Aquatic life and Human Health. Int. J. Environ. Res. Public Health 2020, 17, 2204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Angulo-Bejarano, P.I.; Puente-Rivera, J.; Cruz-Ortega, R. Metal and Metalloid Toxicity in Plants: An Overview on Molecular Aspects. Plants 2021, 10, 635. [Google Scholar] [CrossRef] [PubMed]
  4. Nunes, N.; Ragonezi, C.; Gouveia, C.S.S.; Pinheiro de Carvalho, M.Â.A. Review of Sewage Sludge as a Soil Amendment in Relation to Current International Guidelines: A Heavy Metal Perspective. Sustainability 2021, 13, 2317. [Google Scholar] [CrossRef]
  5. Zgłobicki, W. Special Issue on Heavy Metals in the Environment-Causes and Consequences. Appl. Sci. 2022, 12, 835. [Google Scholar] [CrossRef]
  6. Jaishankar, M.; Tseten, T.; Anbalagan, N.; Mathew, B.B.; Beeregowda, K.N. Toxicity, mechanism and health effects of some heavy metals. Interdiscip. Toxicol. 2014, 7, 60–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Ali, S.; Abbas, Z.; Rizwan, M.; Zaheer, I.E.; Yavaş, İ.; Ünay, A.; Abdel-DAIM, M.M.; Bin-Jumah, M.; Hasanuzzaman, M.; Kalderis, D. Application of Floating Aquatic Plants in Phytoremediation of Heavy Metals Polluted Water: A Review. Sustainability 2020, 12, 1927. [Google Scholar] [CrossRef] [Green Version]
  8. Limo, M.J.; Sola-Rabada, A.; Boix, E.; Thota, V.; Westcott, Z.C.; Puddu, V.; Perry, C.C. Interactions between Metal Oxides and Biomolecules: From Fundamental Understanding to Applications. Chem. Rev. 2018, 118, 11118–11193. [Google Scholar] [CrossRef] [Green Version]
  9. Singh, V.; Singh, J.; Singh, N.; Rai, S.N.; Verma, M.K.; Verma, M.; Singh, V.; Chivate, M.S.; Bilal, M.; Mishra, V. Simultaneous removal of ternary heavy metal ions by a newly isolated Microbacterium paraoxydans strain VSVM IIT(BHU) from coal washery effluent. BioMetals 2022. [Google Scholar] [CrossRef]
  10. Abd Elnabi, M.K.; Elkaliny, N.E.; Elyazied, M.M.; Azab, S.H.; Elkhalifa, S.A.; Elmasry, S.; Mouhamed, M.S.; Shalamesh, E.M.; Alhorieny, N.A.; Abd Elaty, A.E.; et al. Toxicity of Heavy Metals and Recent Advances in Their Removal: A Review. Toxics 2023, 11, 580. [Google Scholar] [CrossRef]
  11. Yang, J.; Hou, B.; Wang, J.; Tian, B.; Bi, J.; Wang, N.; Li, X.; Huang, X. Nanomaterials for the Removal of Heavy Metals from Wastewater. Nanomaterials 2019, 9, 424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Wołowiec, M.; Komorowska-Kaufman, M.; Pruss, A.; Rzepa, G.; Bajda, T. Removal of Heavy Metals and Metalloids from Water Using Drinking Water Treatment Residuals as Adsorbents: A Review. Minerals 2019, 9, 487. [Google Scholar] [CrossRef] [Green Version]
  13. Li, M.; Kuang, S.; Kang, Y.; Ma, H.; Dong, J.; Guo, Z. Recent advances in application of iron-manganese oxide nanomaterials for removal of heavy metals in the aquatic environment. Sci. Total Environ. 2022, 819, 153157. [Google Scholar] [CrossRef] [PubMed]
  14. Singh, V.; Singh, N.; Rai, S.N.; Kumar, A.; Singh, A.K.; Singh, M.P.; Sahoo, A.; Shekhar, S.; Vamanu, E.; Mishra, V. Heavy Metal Contamination in the Aquatic Ecosystem: Toxicity and Its Remediation Using Eco-Friendly Approaches. Toxics 2023, 11, 147. [Google Scholar] [CrossRef] [PubMed]
  15. Elgarahy, A.M.; Elwakeel, K.Z.; Mohammad, S.H.; Elshoubaky, G.A. A critical review of biosorption of dyes, heavy metals and metalloids from wastewater as an efficient and green process. Clean. Eng. Technol. 2021, 4, 100209. [Google Scholar] [CrossRef]
  16. Ungureanu, E.L.; Mocanu, A.L.; Stroe, C.A.; Panciu, C.M.; Berca, L.; Sionel, R.M.; Mustatea, G. Agricultural Byproducts Used as Low-Cost Adsorbents for Removal of Potentially Toxic Elements from Wastewater: A Comprehensive Review. Sustainability 2023, 15, 5999. [Google Scholar] [CrossRef]
  17. Lima de Silva, A.A.; de Carvalho, M.A.; de Souza, S.A.; Dias, P.M.; da Silva Filho, R.G.; de Meirelles Saramago, C.S.; de Melo Bento, C.A.; Hofer, E. Heavy metal tolerance (Cr, Ag AND Hg) in bacteria isolated from sewage. Braz. J. Microbiol. 2012, 43, 1620–1631. [Google Scholar] [CrossRef] [Green Version]
  18. Monachese, M.; Burton, J.P.; Reid, G. Bioremediation and tolerance of humans to heavy metals through microbial processes: A potential role for probiotics? Appl. Environ. Microbiol. 2012, 78, 6397–6404. [Google Scholar] [CrossRef] [Green Version]
  19. Pratush, A.; Kumar, A.; Hu, Z. Adverse effect of heavy metals (As, Pb, Hg, and Cr) on health and their bioremediation strategies: A review. Int. Microbiol. 2018, 21, 97–106. [Google Scholar] [CrossRef]
  20. Luhar, I.; Luhar, S.; Abdullah, M.M.A.B.; Razak, R.A.; Vizureanu, P.; Sandu, A.V.; Matasaru, P.-D. A State-of-the-Art Review on Innovative Geopolymer Composites Designed for Water and Wastewater Treatment. Materials 2021, 14, 7456. [Google Scholar] [CrossRef]
  21. Balali-Mood, M.; Naseri, K.; Tahergorabi, Z.; Khazdair, M.R.; Sadeghi, M. Toxic Mechanisms of Five Heavy Metals: Mercury, Lead, Chromium, Cadmium, and Arsenic. Front. Pharmacol. 2021, 12, 643972. [Google Scholar] [CrossRef]
  22. Wuana, R.A.; Okieimen, F.E. Heavy Metals in Contaminated Soils: A Review of Sources, Chemistry, Risks and Best Available Strategies for Remediation. Int. Sch. Res. Not. 2011, 2011, 402647. [Google Scholar] [CrossRef] [Green Version]
  23. CPCB. General Standards for Discharge of Environmental Pollutants Part-A: Effluents. SCHEDULE—V. Available online: https://cpcb.nic.in/GeneralStandards.pdf (accessed on 9 July 2023).
  24. Shifaw, E. Review of Heavy Metals Pollution in China in Agricultural and Urban Soils. J. Health Pollut. 2018, 8, 180607. [Google Scholar] [CrossRef] [Green Version]
  25. Singh, V.; Singh, M.P.; Mishra, V. Bioremediation of toxic metal ions from coal washery effluent. Desalin. Water Treat. 2020, 197, 300–318. [Google Scholar] [CrossRef]
  26. Bundschuh, J.; Schneider, J.; Alam, M.A.; Niazi, N.K.; Herath, I.; Parvez, F.; Tomaszewska, B.; Guilherme, L.R.G.; Maity, J.P.; López, D.L.; et al. Seven potential sources of arsenic pollution in Latin America and their environmental and health impacts. Sci. Total Environ. 2021, 780, 146274. [Google Scholar] [CrossRef] [PubMed]
  27. Herath, I.; Vithanage, M.; Bundschuh, J.; Maity, J.P.; Bhattacharya, P. Natural Arsenic in Global Groundwaters: Distribution and Geochemical Triggers for Mobilization. Curr. Pollut. Rep. 2016, 2, 68–89. [Google Scholar] [CrossRef] [Green Version]
  28. Missimer, T.M.; Teaf, C.M.; Beeson, W.T.; Maliva, R.G.; Woolschlager, J.; Covert, D.J. Natural background and anthropogenic arsenic enrichment in florida soils, surface water, and groundwater: A review with a discussion on public health risk. Int. J. Environ. Res. Public Health 2018, 15, 2278. [Google Scholar] [CrossRef] [Green Version]
  29. Bidone, E.; Castilhos, Z.; Cesar, R.; Santos, M.C.; Sierpe, R.; Ferreira, M. Hydrogeochemistry of arsenic pollution in watersheds influenced by gold mining activities in Paracatu (Minas Gerais State, Brazil). Environ. Sci. Pollut. Res. Int. 2016, 23, 8546–8555. [Google Scholar] [CrossRef]
  30. Akhtar, N.; Syakir Ishak, M.I.; Bhawani, S.A.; Umar, K. Various Natural and Anthropogenic Factors Responsible for Water Quality Degradation: A Review. Water 2021, 13, 2660. [Google Scholar] [CrossRef]
  31. Masindi, V.; Muedi, K.L. Environmental Contamination by Heavy Metals. In Heavy Metals; InTech: London, UK, 2018. [Google Scholar] [CrossRef] [Green Version]
  32. Mondal, P.; Bhowmick, S.; Chatterjee, D.; Figoli, A.; Van der Bruggen, B. Remediation of inorganic arsenic in groundwater for safe water supply: A critical assessment of technological solutions. Chemosphere 2013, 92, 157–170. [Google Scholar] [CrossRef]
  33. Nita, M.; Grzybowski, A. The Role of the Reactive Oxygen Species and Oxidative Stress in the Pathomechanism of the Age-Related Ocular Diseases and Other Pathologies of the Anterior and Posterior Eye Segments in Adults. Oxidative Med. Cell. Longev. 2016, 2016, 3164734. [Google Scholar] [CrossRef] [Green Version]
  34. Parida, L.; Patel, T.N. Systemic impact of heavy metals and their role in cancer development: A review. Environ. Monit. Assess. 2023, 195, 766. [Google Scholar] [CrossRef]
  35. Liang, G.; Gong, W.; Li, B.; Zuo, J.; Pan, L.; Liu, X. Analysis of Heavy Metals in Foodstuffs and an Assessment of the Health Risks to the General Public via Consumption in Beijing, China. Int. J. Environ. Res. Public Health 2019, 16, 909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Saper, R.B.; Kales, S.N.; Paquin, J.; Burns, M.J.; Eisenberg, D.M.; Davis, R.B.; Phillips, R.S. Heavy Metal Content of Ayurvedic Herbal Medicine Products. JAMA 2004, 292, 2868–2873. [Google Scholar] [CrossRef]
  37. Gnonsoro, U.P.; Ake Assi, Y.E.D.; Sangare, N.S.; Kouakou, Y.U.; Trokourey, A. Health Risk Assessment of Heavy Metals (Pb, Cd, Hg) in Hydroalcoholic Gels of Abidjan, Côte d’Ivoire. Biol. Trace Elem. Res. 2022, 200, 2510–2518. [Google Scholar] [CrossRef]
  38. Jairoun, A.A.; Shahwan, M.; Zyoud, S.H. Heavy Metal contamination of Dietary Supplements products available in the UAE markets and the associated risk. Sci. Rep. 2020, 10, 18824. [Google Scholar] [CrossRef]
  39. Real, M.I.H.; Azam, H.M.; Majed, N. Consumption of heavy metal contaminated foods and associated risks in Bangladesh. Environ. Monit. Assess. 2017, 189, 651. [Google Scholar] [CrossRef] [PubMed]
  40. Nasreddine, L.; Parent-Massin, D. Food contamination by metals and pesticides in the European Union. Should we worry? Toxicol. Lett. 2002, 127, 29–41. [Google Scholar] [CrossRef] [PubMed]
  41. Kepel, B.J.; Bodhi, W.; Fatimawali Tallei, T.E. Heavy metal (As, Cd, Cr, Hg, Pb) analysis and identification of heavy metal resistant bacteria in sediments from Manado Bay. IOP Conf. Ser. Earth Environ. Sci. 2021, 926, 012096. [Google Scholar] [CrossRef]
  42. Yeo, B.; Langley-Turnbaugh, S. Trace Element Deposition on Mount Everest. Soil Surv. Horiz. 2010, 51, 72–78. [Google Scholar] [CrossRef] [Green Version]
  43. Jan, A.T.; Azam, M.; Siddiqui, K.; Ali, A.; Choi, I.; Haq, Q.M. Heavy Metals and Human Health: Mechanistic Insight into Toxicity and Counter Defense System of Antioxidants. Int. J. Mol. Sci. 2015, 16, 29592–29630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Permyakov, E.A. Metal Binding Proteins. Encyclopedia 2021, 1, 261–292. [Google Scholar] [CrossRef]
  45. Wu, X.; Cobbina, S.J.; Mao, G.; Xu, H.; Zhang, Z.; Yang, L. A review of toxicity and mechanisms of individual and mixtures of heavy metals in the environment. Environ. Sci. Pollut. Res. Int. 2016, 23, 8244–8259. [Google Scholar] [CrossRef]
  46. Kim, H.S.; Kim, Y.J.; Seo, Y.R. An Overview of Carcinogenic Heavy Metal: Molecular Toxicity Mechanism and Prevention. J. Cancer Prev. 2015, 20, 232–240. [Google Scholar] [CrossRef]
  47. Chen, Q.Y.; DesMarais, T.; Costa, M. Metals and Mechanisms of Carcinogenesis. Annu. Rev. Pharmacol. Toxicol. 2019, 59, 537–554. [Google Scholar] [CrossRef]
  48. Zhu, Y.; Costa, M. Metals and molecular carcinogenesis. Carcinogenesis 2020, 41, 1161–1172. [Google Scholar] [CrossRef] [PubMed]
  49. Ngalame, N.N.; Makia, N.L.; Waalkes, M.P.; Tokar, E.J. Mitigation of arsenic-induced acquired cancer phenotype in prostate cancer stem cells by miR-143 restoration. Toxicol. Appl. Pharmacol. 2016, 312, 11–18. [Google Scholar] [CrossRef] [Green Version]
  50. Manić, L.; Wallace, D.; Onganer, P.U.; Taalab, Y.M.; Farooqi, A.A.; Antonijević, B.; Buha Djordjevic, A. Epigenetic mechanisms in metal carcinogenesis. Toxicol. Rep. 2022, 9, 778–787. [Google Scholar] [CrossRef]
  51. Zhou, Q.; Xi, S. A review on arsenic carcinogenesis: Epidemiology, metabolism, genotoxicity and epigenetic changes. Regul. Toxicol. Pharmacol. 2018, 99, 78–88. [Google Scholar] [CrossRef]
  52. Singh, V.; Singh, J.; Mishra, V. Development of a cost-effective, recyclable and viable metal ion doped adsorbent for simultaneous adsorption and reduction of toxic Cr (VI) ions. J. Environ. Chem. Eng. 2021, 9, 105124. [Google Scholar] [CrossRef]
  53. YG, T.G.; Ballupete Nagaraju, S.; Puttegowda, M.; Verma, A.; Rangappa, S.M.; Siengchin, S. Biopolymer-Based Composites: An Eco-Friendly Alternative from Agricultural Waste Biomass. J. Compos. Sci. 2023, 7, 242. [Google Scholar] [CrossRef]
  54. Singh, V.; Mishra, V. Sustainable reduction of Cr (VI) and its elemental mapping on chitosan coated citrus limetta peels biomass in synthetic wastewater. Sep. Sci. Technol. 2021, 7, 1609–1626. [Google Scholar] [CrossRef]
  55. Bilal, M.; Rasheed, T.; Sosa-Hernández, J.E.; Raza, A.; Nabeel, F.; Iqbal, H.M.N. Biosorption: An Interplay between Marine Algae and Potentially Toxic Elements-A Review. Mar. Drugs 2018, 16, 65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Deshmukh, R.; Khardenavis, A.A.; Purohit, H.J. Diverse Metabolic Capacities of Fungi for Bioremediation. Indian J. Microbiol. 2016, 56, 247–264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Manikant, T.; Pankaj, S.; Singh, R.; Bala, S.; Pathak, N.; Singh, S.; Chauhan, R.S.; Singh, P.K. Microbial biosorbent for remediation of dyes and heavy metals pollution: A green strategy for sustainable environment. Front. Microbiol. 2023, 14, 1168954. [Google Scholar] [CrossRef]
  58. Sharma, P.; Singh, S.P.; Parakh, S.K.; Tong, Y.W. Health hazards of hexavalent chromium (Cr(VI)) and its microbial reduction. Bioengineered 2022, 13, 4923–4938. [Google Scholar] [CrossRef]
  59. Hossan, S.; Hossain, S.; Islam, M.R.; Kabir, M.H.; Ali, S.; Islam, M.S.; Imran, K.M.; Moniruzzaman, M.; Mou, T.J.; Parvez, A.K.; et al. Bioremediation of Hexavalent Chromium by Chromium Resistant Bacteria Reduces Phytotoxicity. Int. J. Environ. Res. Public Health 2020, 17, 6013. [Google Scholar] [CrossRef]
  60. Akkurt, Ş.; Oguz, M.; Alkan Uçkun, A. Bioreduction and bioremoval of hexavalent chromium by genetically engineered strains (Escherichia coli MT2A and Escherichia coli MT3). World J. Microbiol. Biotechnol. 2022, 38, 45. [Google Scholar] [CrossRef]
  61. Diep, P.; Mahadevan, R.; Yakunin, A.F. Heavy Metal Removal by Bioaccumulation Using Genetically Engineered Microorganisms. Front. Bioeng. Biotechnol. 2018, 6, 157. [Google Scholar] [CrossRef] [Green Version]
  62. Banerjee, A.; Jhariya, M.K.; Yadav, D.K.; Raj, A. Micro-remediation of Metals: A New Frontier in Bioremediation. In Handbook of Environmental Materials Management; Hussain, C., Ed.; Springer: Cham, Switzerland, 2018. [Google Scholar] [CrossRef]
  63. Kapahi, M.; Sachdeva, S. Bioremediation Options for Heavy Metal Pollution. J. Health Pollut. 2019, 9, 191203. [Google Scholar] [CrossRef] [Green Version]
  64. Sharma, P.; Dutta, D.; Udayan, A.; Nadda, A.K.; Lam, S.S.; Kumar, S. Role of microbes in bioaccumulation of heavy metals in municipal solid waste: Impacts on plant and human being. Environ. Pollut. 2022, 305, 119248. [Google Scholar] [CrossRef] [PubMed]
  65. Gupta, R.; Mohapatra, H. Microbial biomass: An economical alternative for removal of heavy metals from waste water. Indian J. Exp. Biol. 2003, 41, 945–966. [Google Scholar] [PubMed]
  66. Pande, V.; Pandey, S.C.; Sati, D.; Bhatt, P.; Samant, M. Microbial Interventions in Bioremediation of Heavy Metal Contaminants in Agroecosystem. Front. Microbiol. 2022, 13, 824084. [Google Scholar] [CrossRef]
  67. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Ojuederie, O.B.; Babalola, O.O. Microbial and Plant-Assisted Bioremediation of Heavy Metal Polluted Environments: A Review. Int. J. Environ. Res. Public Health 2017, 14, 1504. [Google Scholar] [CrossRef] [Green Version]
  69. Jacob, J.M.; Karthik, C.; Saratale, R.G.; Kumar, S.S.; Prabakar, D.; Kadirvelu, K.; Pugazhendhi, A. Biological approaches to tackle heavy metal pollution: A survey of literature. J. Environ. Manag. 2018, 217, 56–70. [Google Scholar] [CrossRef]
  70. Gaur, N.; Flora, G.; Yadav, M.; Tiwari, A. A review with recent advancements on bioremediation-based abolition of heavy metals. Environ. Sci. Process. Impacts 2014, 16, 180–193. [Google Scholar] [CrossRef]
  71. Gunjyal, N.; Rani, S.; Asgari Lajayer, B.; Senapathi, V.; Astatkie, T. A review of the effects of environmental hazards on humans, their remediation for sustainable development, and risk assessment. Environ. Monit. Assess. 2023, 195, 795. [Google Scholar] [CrossRef]
  72. Agarwal, P.; Vibhandik, R.; Agrahari, R.; Daverey, A.; Rani, R. Role of Root Exudates on the Soil Microbial Diversity and Biogeochemistry of Heavy Metals. Appl. Biochem. Biotechnol. 2023. [Google Scholar] [CrossRef]
  73. Kanamarlapudi, S.L.R.K.; Chintalpudi, V.K.; Muddada, S. Application of Biosorption for Removal of Heavy Metals from Wastewater. In Biosorption; InTech: London, UK, 2018. [Google Scholar] [CrossRef] [Green Version]
  74. Henry, H.F.; Suk, W.A. Sustainable exposure prevention through innovative detection and remediation technologies from the NIEHS Superfund Research Program. Rev. Environ. Health 2017, 32, 35–44. [Google Scholar] [CrossRef]
  75. Torres, E. Biosorption: A Review of the Latest Advances. Processes 2020, 8, 1584. [Google Scholar] [CrossRef]
  76. Dey, S.; Basha, S.R.; Babu, G.V.; Nagendra, T. Characteristic and biosorption capacities of orange peels biosorbents for removal of ammonia and nitrate from contaminated water. Clean. Mater. 2021, 1, 100001. [Google Scholar] [CrossRef]
  77. Singh, V.; Singh, N.; Yadav, P.; Mishra, V. Removal of Hexavalent Chromium from Aqueous Media Using Eco-Friendly and Cost-Effective Biological Methods. In Biosorption for Wastewater Contaminants; Selvasembian, R., Singh, P., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2021; pp. 246–268. [Google Scholar] [CrossRef]
  78. El-Naggar, N.E.; Hamouda, R.A.; Mousa, I.E.; Abdel-Hamid, M.S.; Rabei, N.H. Biosorption optimization, characterization, immobilization and application of Gelidium amansii biomass for complete Pb2+ removal from aqueous solutions. Sci. Rep. 2018, 8, 13456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Michalak, I.; Chojnacka, K.; Witek-Krowiak, A. State of the art for the biosorption process—A review. Appl. Biochem. Biotechnol. 2013, 170, 1389–1416. [Google Scholar] [CrossRef] [Green Version]
  80. Qasem, N.A.A.; Mohammed, R.H.; Lawal, D.U. Removal of heavy metal ions from wastewater: A comprehensive and critical review. NPJ Clean Water 2021, 4, 36. [Google Scholar] [CrossRef]
  81. de Freitas, G.R.; da Silva, M.G.C.; Vieira, M.G.A. Biosorption technology for removal of toxic metals: A review of commercial biosorbents and patents. Environ. Sci. Pollut. Res. Int. 2019, 26, 19097–19118. [Google Scholar] [CrossRef] [PubMed]
  82. Nur-E-Alam, M.; Mia, M.A.S.; Ahmad, F.; Rahman, M.M. An overview of chromium removal techniques from tannery effluent. Appl. Water Sci. 2020, 10, 205. [Google Scholar] [CrossRef]
  83. Dai, J.; Ren, F.; Tao, C. Adsorption of Cr(VI) and speciation of Cr(VI) and Cr(III) in aqueous solutions using chemically modified chitosan. Int. J. Environ. Res. Public Health 2012, 9, 1757–1770. [Google Scholar] [CrossRef] [Green Version]
  84. Gopal, B.; Gupta, A. Integrated Approach for Hazardous Cr(VI) Removal: Reduction, Extraction, and Conversion into a Photoactive Composite, CuO/CuCr2O4. ACS Omega 2019, 4, 20443–20449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Wang, G.; Zhang, S.; Yao, P.; Chen, Y.; Xu, X.; Li, T.; Gong, G. Removal of Pb(II) from aqueous solutions by Phytolacca americana L. biomass as a low cost biosorbent. Arab. J. Chem. 2018, 11, 99–110. [Google Scholar] [CrossRef] [Green Version]
  86. Chang, J.S.; Law, R.; Chang, C.C. Biosorption of lead, copper and cadmium by biomass of Pseudomonas aeruginosa PU21. Water Res. 1997, 31, 1651–1658. [Google Scholar] [CrossRef]
  87. Manasi; Rajesh, V.; Kumar, A.S.K.; Rajesh, N. Biosorption of cadmium using a novel bacterium isolated from an electronic industry effluent. Chem. Eng. J. 2014, 235, 176–185. [Google Scholar] [CrossRef]
  88. Black, R.; Sartaj, M.; Mohammadian, A.; Qiblawey, H.A.M. Biosorption of Pb and Cu using fixed and suspended bacteria. J. Environ. Chem. Eng. 2014, 2, 1663–1671. [Google Scholar] [CrossRef]
  89. Aryal, M.; Ziagova, M.; Liakopoulou-Kyriakides, M. Study on arsenic biosorption using Fe(III)-treated biomass of Staphylococcus xylosus. Chem. Eng. J. 2010, 162, 178–185. [Google Scholar] [CrossRef]
  90. Hoyle-Gardner, J.; Jones, W.; Badisa, V.L.D.; Mwashote, B.; Ibeanusi, V.; Gaines, T.; Lowenthal, H.; Tucker, L. Lead metal biosorption and isotherms studies by metal-resistant Bacillus strain MRS-2 bacterium. J. Basic Microbiol. 2021, 61, 697–708. [Google Scholar] [CrossRef] [PubMed]
  91. Bai, J.; Yang, X.; Du, R.; Chen, Y.; Wang, S.; Qiu, R. Biosorption mechanisms involved in immobilization of soil Pb by Bacillus subtilis DBM in a multi-metal-contaminated soil. J. Environ. Sci. 2014, 26, 2056–2064. [Google Scholar] [CrossRef]
  92. Rizvi, A.; Ahmed, B.; Zaidi, A.; Khan, M.S. Biosorption of heavy metals by dry biomass of metal tolerant bacterial biosorbents: An efficient metal clean-up strategy. Environ. Monit. Assess. 2020, 192, 801. [Google Scholar] [CrossRef]
  93. Bié, J.; Sepodes, B.; Fernandes, P.C.B.; Ribeiro, M.H.L. Enzyme Immobilization and Co-Immobilization: Main Framework, Advances and Some Applications. Processes 2022, 10, 494. [Google Scholar] [CrossRef]
  94. Zdarta, J.; Meyer, A.S.; Jesionowski, T.; Pinelo, M. A General Overview of Support Materials for Enzyme Immobilization: Characteristics, Properties, Practical Utility. Catalysts 2018, 8, 92. [Google Scholar] [CrossRef] [Green Version]
  95. da Costa, A.C.A.; Leite, S.G.F. Metals biosorption by sodium alginate immobilizedChlorella homosphaera cells. Biotechnol. Lett. 1991, 13, 559–562. [Google Scholar] [CrossRef]
  96. Timková, I.; Sedláková-Kaduková, J.; Pristaš, P. Biosorption and Bioaccumulation Abilities of Actinomycetes/Streptomycetes Isolated from Metal Contaminated Sites. Separations 2018, 5, 54. [Google Scholar] [CrossRef] [Green Version]
  97. Igiri, B.E.; Okoduwa, S.I.R.; Idoko, G.O.; Akabuogu, E.P.; Adeyi, A.O.; Ejiogu, I.K. Toxicity and Bioremediation of Heavy Metals Contaminated Ecosystem from Tannery Wastewater: A Review. J. Toxicol. 2018, 2018, 2568038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Alengebawy, A.; Abdelkhalek, S.T.; Qureshi, S.R.; Wang, M.-Q. Heavy Metals and Pesticides Toxicity in Agricultural Soil and Plants: Ecological Risks and Human Health Implications. Toxics 2021, 9, 42. [Google Scholar] [CrossRef] [PubMed]
  99. Genchi, G.; Carocci, A.; Lauria, G.; Sinicropi, M.S.; Catalano, A. Thallium Use, Toxicity, and Detoxification Therapy: An Overview. Appl. Sci. 2021, 11, 8322. [Google Scholar] [CrossRef]
  100. Genchi, G.; Lauria, G.; Catalano, A.; Carocci, A.; Sinicropi, M.S. The Double Face of Metals: The Intriguing Case of Chromium. Appl. Sci. 2021, 11, 638. [Google Scholar] [CrossRef]
  101. Gupta, P.; Diwan, B. Bacterial Exopolysaccharide mediated heavy metal removal: A Review on biosynthesis, mechanism and remediation strategies. Biotechnol. Rep. 2016, 23, 58–71. [Google Scholar] [CrossRef]
  102. Li, C.; Yu, Y.; Fang, A.; Feng, D.; Du, M.; Tang, A.; Chen, S.; Li, A. Insight into biosorption of heavy metals by extracellular polymer substances and the improvement of the efficacy: A review. Lett. Appl. Microbiol. 2022, 75, 1064–1073. [Google Scholar] [CrossRef]
  103. Henson, M.W.; Santo Domingo, J.W.; Kourtev, P.S.; Jensen, R.V.; Dunn, J.A.; Learman, D.R. Metabolic and genomic analysis elucidates strain-level variation in Microbacterium spp. isolated from chromate contaminated sediment. PeerJ 2015, 3, 1395–1412. [Google Scholar] [CrossRef] [Green Version]
  104. Pattanapipitpaisal, P.; Brown, N.L.; Macaskie, L.E. Chromate reduction and 16S rRNA identification of bacteria isolated from a Cr (VI)-contaminated site. Appl. Microbiol. Biotechnol. 2001, 57, 257–261. [Google Scholar]
  105. Ahemad, M. Enhancing phytoremediation of chromium-stressed soils through plant-growth-promoting bacteria. J. Genet. Eng. Biotechnol. 2015, 13, 51–58. [Google Scholar] [CrossRef] [Green Version]
  106. Kubrak, O.I.; Lushchak, O.V.; Lushchak, J.V.; Torous, I.M.; Storey, J.M.; Storey, K.B.; Lushchak, V.I. Chromium effects on free radical processes in goldfish tissues: Comparison of Cr (III) and Cr (VI) exposures on oxidative stress markers, glutathione status and antioxidant enzymes. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2010, 152, 360–370. [Google Scholar] [CrossRef] [PubMed]
  107. Sachdev, S.; Ansari, S.A.; Ansari, M.I.; Fujita, M.; Hasanuzzaman, M. Abiotic Stress and Reactive Oxygen Species: Generation, Signaling, and Defense Mechanisms. Antioxidants 2021, 10, 277. [Google Scholar] [CrossRef]
  108. Bhattacharyya, A.; Chattopadhyay, R.; Mitra, S.; Crowe, S.E. Oxidative stress: An essential factor in the pathogenesis of gastrointestinal mucosal diseases. Physiol. Rev. 2014, 94, 329–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Turner, R.J.; Huang, L.N.; Viti, C.; Mengoni, A. Metal-Resistance in Bacteria: Why Care? Genes 2020, 11, 1470. [Google Scholar] [CrossRef] [PubMed]
  110. Tariq, A.; Ullah, U.; Asif, M.; Sadiq, I. Biosorption of arsenic through bacteria isolated from Pakistan. Int. Microbiol. 2019, 22, 59–68. [Google Scholar] [CrossRef]
  111. Jebelli, M.A.; Maleki, A.; Amoozegar, M.A.; Kalantar, E.; Shahmoradi, B.; Gharibi, F. Isolation and identification of indigenous prokaryotic bacteria from arsenic-contaminated water resources and their impact on arsenic transformation. Ecotoxicol. Environ. Saf. 2017, 140, 170–176. [Google Scholar] [CrossRef]
  112. Parashar, D.; Satyanarayana, T. An Insight Into Ameliorating Production, Catalytic Efficiency, Thermostability and Starch Saccharification of Acid-Stable α-Amylases From Acidophiles. Front. Bioeng. Biotechnol. 2018, 6, 125. [Google Scholar] [CrossRef] [Green Version]
  113. Wang, Y.S.; Pan, Z.Y.; Lang, J.M.; Xu, J.M.; Zheng, Y.G. Bioleaching of chromium from tannery sludge by indigenous Acidithiobacillus thiooxidans. J. Hazard. Mater. 2007, 147, 319–324. [Google Scholar] [CrossRef]
  114. Branca, T.A.; Colla, V.; Algermissen, D.; Granbom, H.; Martini, U.; Morillon, A.; Pietruck, R.; Rosendahl, S. Reuse and Recycling of By-Products in the Steel Sector: Recent Achievements Paving the Way to Circular Economy and Industrial Symbiosis in Europe. Metals 2020, 10, 345. [Google Scholar] [CrossRef] [Green Version]
  115. Álvarez, M.L.; Fidalgo, J.M.; Gascó, G.; Méndez, A. Hydrometallurgical Recovery of Cu and Zn from a Complex Sulfide Mineral by Fe3+/H2SO4 Leaching in the Presence of Carbon-Based Materials. Metals 2021, 11, 286. [Google Scholar] [CrossRef]
  116. Sarkodie, E.K.; Jiang, L.; Li, K.; Yang, J.; Guo, Z.; Shi, J.; Deng, Y.; Liu, H.; Jiang, H.; Liang, Y.; et al. A review on the bioleaching of toxic metal(loid)s from contaminated soil: Insight into the mechanism of action and the role of influencing factors. Front. Microbiol. 2022, 13, 1049277. [Google Scholar] [CrossRef]
  117. Ummalyma, S.B.; Sahoo, D.; Pandey, A. Resource recovery through bioremediation of wastewaters and waste carbon by microalgae: A circular bioeconomy approach. Environ. Sci. Pollut. Res. Int. 2021, 28, 58837–58856. [Google Scholar] [CrossRef] [PubMed]
  118. Dickerhof, N.; Isles, V.; Pattemore, P.; Hampton, M.B.; Kettle, A.J. Exposure of Pseudomonas aeruginosa to bactericidal hypochlorous acid during neutrophil phagocytosis is compromised in cystic fibrosis. J. Biol. Chem. 2019, 294, 13502–13514. [Google Scholar] [CrossRef] [PubMed]
  119. Cha, M.; Lee, N.; Kim, M.; Kim, M.; Lee, S. Heterologous production of Pseudomonas aeruginosa EMS1 biosurfactant in Pseudomonas putida. Bioresour. Technol. 2008, 99, 2192–2199. [Google Scholar] [CrossRef] [PubMed]
  120. Bayramoglu, G.; Arica, M.Y. Preparation of a Composite Biosorbent Using Scenedesmus quadricauda Biomass and Alginate/Polyvinyl Alcohol for Removal of Cu(II) and Cd(II) Ions: Isotherms, Kinetics, and Thermodynamic Studies. Water Air Soil Pollut. 2011, 221, 391–403. [Google Scholar] [CrossRef]
  121. Shanu-Wilson, J.; Evans, L.; Wrigley, S.; Steele, J.; Atherton, J.; Boer, J. Biotransformation: Impact and Application of Metabolism in Drug Discovery. ACS Med. Chem. Lett. 2020, 11, 2087–2107. [Google Scholar] [CrossRef]
  122. Smitha, M.S.; Singh, S.; Singh, R. Microbial bio transformation: A process for chemical alterations. J. Bacteriol. Mycol. Open Access 2017, 4, 47–51. [Google Scholar]
  123. Irawati, W.; Parhusip, A.J.N.; Sopiah, N.; Tnunay, J.A. The Role of Heavy Metals-Resistant Bacteria Acinetobacter sp. in Copper Phytoremediation using Eichhornia crasippes [(Mart.) Solms]. KnE Life Sci. 2017, 3, 208–220. [Google Scholar] [CrossRef] [Green Version]
  124. Ilias, M.; Rafiqullah, I.M.; Debnath, B.C.; Mannan, K.S.; Mozammel Hoq, M. Isolation and Characterization of Chromium(VI)-Reducing Bacteria from Tannery Effluents. Indian J. Microbiol. 2011, 51, 76–81. [Google Scholar] [CrossRef] [Green Version]
  125. Correa-García, S.; Pande, P.; Séguin, A.; St-Arnaud, M.; Yergeau, E. Rhizoremediation of petroleum hydrocarbons: A model system for plant microbiome manipulation. Microb. Biotechnol. 2018, 11, 819–832. [Google Scholar] [CrossRef] [Green Version]
  126. Dixit, R.; Wasiullah; Malaviya, D.; Pandiyan, K.; Singh, U.B.; Sahu, A.; Shukla, R.; Singh, B.P.; Rai, J.P.; Sharma, P.K.; et al. Bioremediation of Heavy Metals from Soil and Aquatic Environment: An Overview of Principles and Criteria of Fundamental Processes. Sustainability 2015, 7, 2189–2212. [Google Scholar] [CrossRef] [Green Version]
  127. Singh, H.; Pant, G. Phytoremediation: Low input-based ecological approach for sustainable environment. Appl. Water Sci. 2023, 13, 85. [Google Scholar] [CrossRef]
  128. Skuza, L.; Szućko-Kociuba, I.; Filip, E.; Bożek, I. Natural Molecular Mechanisms of Plant Hyperaccumulation and Hypertolerance towards Heavy Metals. Int. J. Mol. Sci. 2022, 23, 9335. [Google Scholar] [CrossRef] [PubMed]
  129. Bian, F.; Zhong, Z.; Zhang, X.; Yang, C.; Gai, X. Bamboo—An untapped plant resource for the phytoremediation of heavy metal contaminated soils. Chemosphere 2020, 246, 125750. [Google Scholar] [CrossRef]
  130. Bian, F.; Zhong, Z.; Wu, S.; Zhang, X.; Yang, C.; Xiong, X. Comparison of heavy metal phytoremediation in monoculture and intercropping systems of Phyllostachys praecox and Sedum plumbizincicola in polluted soil. Int. J. Phytoremediation 2018, 20, 490–498. [Google Scholar] [CrossRef] [PubMed]
  131. Verma, P.; Rawat, S. Rhizoremediation of Heavy Metal- and Xenobiotic-Contaminated Soil: An Eco-Friendly Approach. In Removal of Emerging Contaminants Through Microbial Processes; Shah, M.P., Ed.; Springer: Singapore, 2021. [Google Scholar] [CrossRef]
  132. Poria, V.; Dębiec-Andrzejewska, K.; Fiodor, A.; Lyzohub, M.; Ajijah, N.; Singh, S.; Pranaw, K. Plant Growth-Promoting Bacteria (PGPB) integrated phytotechnology: A sustainable approach for remediation of marginal lands. Front. Plant Sci. 2022, 13, 999866. [Google Scholar] [CrossRef]
  133. Ma, Y.; Wang, Y.; Shi, X.J.; Chen, X.P.; Li, Z.L. Mechanism and Application of Plant Growth-Promoting Bacteria in Heavy Metal Bioremediation. Huan Jing Ke Xue 2022, 43, 4911–4922. [Google Scholar]
  134. Alves, A.R.A.; Yin, Q.; Oliveira, R.S.; Silva, E.F.; Novo, L.A.B. Plant growth-promoting bacteria in phytoremediation of metal-polluted soils: Current knowledge and future directions. Sci. Total Environ. 2022, 838, 156435. [Google Scholar] [CrossRef] [PubMed]
  135. Liu, Y.; Feng, J.; Pan, H.; Zhang, X.; Zhang, Y. Genetically engineered bacterium: Principles, practices, and prospects. Front. Microbiol. 2022, 13, 997587. [Google Scholar] [CrossRef] [PubMed]
  136. Saravanan, A.; Kumar, P.S.; Ramesh, B. Srinivasan, Removal of toxic heavy metals using genetically engineered microbes: Molecular tools, risk assessment and management strategies. Chemosphere 2022, 298, 134341. [Google Scholar] [CrossRef]
  137. Aquino, E.; Barbieri, C.; Oller Nascimento, C.A. Engineering Bacteria for Bioremediation. In Progress in Molecular and Environmental Bioengineering; InTech: London, UK, 2011. [Google Scholar] [CrossRef] [Green Version]
  138. Khan, S.; Ullah, M.W.; Siddique, R.; Nabi, G.; Manan, S.; Yousaf, M.; Hou, H. Role of Recombinant DNA Technology to Improve Life. Int. J. Genom. 2016, 2016, 2405954. [Google Scholar] [CrossRef] [Green Version]
  139. Liu, S.; Zhang, F.; Chen, J.; Sun, G. Arsenic removal from contaminated soil via biovolatilization by genetically engineered bacteria under laboratory conditions. J. Environ. Sci. 2011, 23, 1544–1550. [Google Scholar] [CrossRef] [PubMed]
  140. Bondarenko, O.; Rolova, T.; Kahru, A.; Ivask, A. Bioavailability of Cd, Zn and Hg in soil to nine recombinant luminescent metal sensor bacteria. Sensors 2008, 8, 6899–6923. [Google Scholar] [CrossRef] [Green Version]
  141. Pazirandeh, M.; Chrisey, L.A.; Mauro, J.M.; Campbell, J.R.; Gaber, B.P. Expression of the Neurospora crassa metallothionein gene in Escherichia coli and its effect on heavy-metal uptake. Appl. Microbiol. Biotechnol. 1995, 43, 1112–1117. [Google Scholar] [CrossRef] [PubMed]
  142. Azad, M.A.K.; Amin, L.; Sidik, N.M. Genetically engineered organisms for bioremediation of pollutants in contaminated sites. Chin. Sci. Bull. 2014, 59, 703–714. [Google Scholar] [CrossRef]
  143. Nascimento, A.M.; Chartone-Souza, E. Operon mer: Bacterial resistance to mercury and potential for bioremediation of contaminated environments. Genet. Mol. Res. 2003, 2, 92–101. [Google Scholar]
  144. De Groote, V.N.; Verstraeten, N.; Fauvart, M.; Kint, C.I.; Verbeeck, A.M.; Beullens, S.; Cornelis, P.; Michiels, J. Novel persistence genes in Pseudomonas aeruginosa identified by high-throughput screening. FEMS Microbiol. Lett. 2009, 297, 73–79. [Google Scholar] [CrossRef] [Green Version]
  145. Chen, S.; Wilson, D.B. Genetic engineering of bacteria and their potential for Hg2+ bioremediation. Biodegradation 1997, 8, 97–103. [Google Scholar] [CrossRef] [PubMed]
  146. Fasani, E.; Manara, A.; Martini, F.; Furini, A.; DalCorso, G. The potential of genetic engineering of plants for the remediation of soils contaminated with heavy metals. Plant Cell Environ. 2018, 41, 1201–1232. [Google Scholar] [CrossRef]
  147. Kotrba, P.; Najmanova, J.; Macek, T.; Ruml, T.; Mackova, M. Genetically modified plants in phytoremediation of heavy metal and metalloid soil and sediment pollution. Biotechnol. Adv. 2009, 27, 799–810. [Google Scholar] [CrossRef]
  148. Boechat, C.L.; Pistóia, V.C.; Gianelo, C.; Camargo, F.A. Accumulation and translocation of heavy metal by spontaneous plants growing on multi-metal-contaminated site in the Southeast of Rio Grande do Sul state, Brazil. Environ. Sci. Pollut. Res. Int. 2016, 23, 2371–2380. [Google Scholar] [CrossRef] [PubMed]
  149. Zhu, N.; Zhang, B.; Yu, Q. Genetic Engineering-Facilitated Coassembly of Synthetic Bacterial Cells and Magnetic Nanoparticles for Efficient Heavy Metal Removal. ACS Appl. Mater. Interfaces 2020, 12, 22948–22957. [Google Scholar] [CrossRef] [PubMed]
  150. Sharma, P.; Sirohi, R.; Tong, Y.W.; Kim, S.H.; Pandey, A. Metal and metal(loids) removal efficiency using genetically engineered microbes: Applications and challenges. J. Hazard. Mater. 2021, 416, 125855. [Google Scholar] [CrossRef] [PubMed]
  151. Qin, D. Next-generation sequencing and its clinical application. Cancer Biol. Med. 2019, 16, 4–10. [Google Scholar]
  152. Basharat, Z.; Novo, L.A.B.; Yasmin, A. Genome Editing Weds CRISPR: What Is in It for Phytoremediation? Plants 2018, 7, 51. [Google Scholar] [CrossRef] [Green Version]
  153. Kim, S.; Takahashi, M.; Higuchi, K.; Tsunoda, K.; Nakanishi, H.; Yoshimura, E.; Nishizawa, N.K. Increased nicotianamine biosynthesis confers enhanced tolerance of high levels of metals, in particular nickel, to plants. Plant Cell Physiol 2005, 46, 1809–1818. [Google Scholar] [CrossRef] [Green Version]
  154. Thai, T.D.; Lim, W.; Na, D. Synthetic bacteria for the detection and bioremediation of heavy metals. Front. Bioeng. Biotechnol. 2023, 11, 1178680. [Google Scholar] [CrossRef]
  155. Mishra, G.K. Microbes in Heavy Metal Remediation: A Review on Current Trends and Patents. Recent Pat. Biotechnol. 2017, 11, 188–196. [Google Scholar] [CrossRef]
Figure 1. This study focuses on the remediation of heavy metals using bacteria and the mechanism of heavy metal reduction in bacterial cells [14].
Figure 1. This study focuses on the remediation of heavy metals using bacteria and the mechanism of heavy metal reduction in bacterial cells [14].
Applsci 13 08439 g001
Figure 2. Processes involved in rhizoremediation for heavy metals.
Figure 2. Processes involved in rhizoremediation for heavy metals.
Applsci 13 08439 g002
Table 1. The heavy metals removal by microbial-based biomasses and heavy metal biosorption capacity.
Table 1. The heavy metals removal by microbial-based biomasses and heavy metal biosorption capacity.
BiosorbentHeavy Metal IonHeavy Metal Biosorption Capacity (mg/g)References
Phytolacca americana L. biomassPb (II)10.83[85]
Pseudomonas aeruginosa PU21 (Rip64)Cu (II)23[86]
Pseudomonas aeruginosa PU21 (Rip64)Pb (II)70[86]
Pseudomonas aeruginosa PU21 (Rip64)Cd (II)58 [86]
Halomonas BVR 1Cd (II)12.023[87]
Mix bacterial biomass of Enterobacter ludwigii, Zoogloea ramigerais, and Comamonas testosteroniCu (II)6.52[88]
Staphylococus xylosus biomassAs (III)54.35 [89]
Staphylococus xylosus biomassAs (V)61.34[89]
Bacillus strain MRS-2 (ATCC 55674)Pb (II)206.5[90]
Bacillus subtilisPb (II)57[91]
Table 2. Heavy metal uptake capacity of bacterial strains.
Table 2. Heavy metal uptake capacity of bacterial strains.
BacteriaHeavy MetalsInitial Heavy Metal Concentration (mg/L)Heavy Metal Removal Capacity (%)References
Pseudomonas aeruginosa PU21 (Rip64)Pb (II)5098[86]
Pseudomonas aeruginosa PU21 (Rip64)Cu (II)5098[86]
Pseudomonas aeruginosa PU21 (Rip64)Cd (II)5098[86]
Enterobacter ludwigii, Zoogloea ramigerais, and Comamonas testosterone mix culturePb (II)1072.6[88]
Enterobacter ludwigii, Zoogloea ramigerais, and Comamonas testosterone mix cultureCu (II)2.583.9[88]
Pseudomonas aeruginosaAs (III and V)1098[110]
Pseudomonas strain As-11As (III)13048[111]
Pseudomonas strain As-11As (V)31278[111]
Microbacterium paraoxydans VSVM IIT (BHU)Cr (VI)5091.62[9]
Microbacterium paraoxydans VSVM IIT (BHU)Cd (II)5089.29[9]
Microbacterium paraoxydans VSVM IIT (BHU)Pb (II)5083.29[9]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, B.; Zhang, T.; Wang, F. Microbial-Based Heavy Metal Bioremediation: Toxicity and Eco-Friendly Approaches to Heavy Metal Decontamination. Appl. Sci. 2023, 13, 8439. https://doi.org/10.3390/app13148439

AMA Style

Zhou B, Zhang T, Wang F. Microbial-Based Heavy Metal Bioremediation: Toxicity and Eco-Friendly Approaches to Heavy Metal Decontamination. Applied Sciences. 2023; 13(14):8439. https://doi.org/10.3390/app13148439

Chicago/Turabian Style

Zhou, Biao, Tiejian Zhang, and Fei Wang. 2023. "Microbial-Based Heavy Metal Bioremediation: Toxicity and Eco-Friendly Approaches to Heavy Metal Decontamination" Applied Sciences 13, no. 14: 8439. https://doi.org/10.3390/app13148439

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