Next Article in Journal
Attributes of Culture Bacteria as Influenced by Ingredients That Help Treat Leaky Gut
Next Article in Special Issue
An In Silico Bioremediation Study to Identify Essential Residues of Metallothionein Enhancing the Bioaccumulation of Heavy Metals in Pseudomonas aeruginosa
Previous Article in Journal
Monitoring Growth and Removal of Pseudomonas Biofilms on Cellulose-Based Fabrics
Previous Article in Special Issue
The Transformation of Hg2+ during Anaerobic S0 Reduction by an AMD Environmental Enrichment Culture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Use of Microbial Consortia in Bioremediation of Metalloid Polluted Environments

by
Elham Lashani
1,
Mohammad Ali Amoozegar
1,*,
Raymond J. Turner
2 and
Hamid Moghimi
3,*
1
Extremophiles Laboratory, Department of Microbiology, School of Biology and Center of Excellence in Phylogeny of Living Organisms, College of Science, University of Tehran, Tehran 14178-64411, Iran
2
Microbial Biochemistry Laboratory, Department of Biological Sciences, University of Calgary, 2500 University Dr. NW, Calgary, AB T2N 1N4, Canada
3
Department of Microbiology, School of Biology, College of Science, University of Tehran, Tehran 14178-64411, Iran
*
Authors to whom correspondence should be addressed.
Microorganisms 2023, 11(4), 891; https://doi.org/10.3390/microorganisms11040891
Submission received: 18 February 2023 / Revised: 8 March 2023 / Accepted: 13 March 2023 / Published: 30 March 2023

Abstract

:
Metalloids are released into the environment due to the erosion of the rocks or anthropogenic activities, causing problems for human health in different world regions. Meanwhile, microorganisms with different mechanisms to tolerate and detoxify metalloid contaminants have an essential role in reducing risks. In this review, we first define metalloids and bioremediation methods and examine the ecology and biodiversity of microorganisms in areas contaminated with these metalloids. Then we studied the genes and proteins involved in the tolerance, transport, uptake, and reduction of these metalloids. Most of these studies focused on a single metalloid and co-contamination of multiple pollutants were poorly discussed in the literature. Furthermore, microbial communication within consortia was rarely explored. Finally, we summarized the microbial relationships between microorganisms in consortia and biofilms to remove one or more contaminants. Therefore, this review article contains valuable information about microbial consortia and their mechanisms in the bioremediation of metalloids.

1. Introduction

1.1. Metalloids

Metalloid elements have some physical and chemical characteristics between metal and nonmetal elements. Metalloids included arsenic (As), antimony (Sb), boron (B), germanium (Ge), silicon (Si), tellurium (Te), and polonium (Po)—although we think less of this element in biology due to its radioactivity. Astatine (At) and selenium (Se) are also sometimes considered as they have metalloid characteristics under certain conditions. Metalloids are exploited in medicine [1,2,3,4], semiconductor devices [5,6,7], glasses and ceramics [8,9], solar cells/batteries [10,11], specific polymers [12], construction material, and agricultural applications [11,13]. B, Si, and Se have essential functions in most organisms, including humans, as essential trace elements. However, high levels of these elements are toxic. While As, Sb, Te, and Ge are more toxic [14].

1.2. Prevalence and Toxicity of Metalloids

1.2.1. Selenium

Selenium was discovered by Jons Jacob Berzelius in 1808 and derived from the Greek word selene which means the moon. This element is the 69th most abundant element on the earth and belongs to the group V periodic table [15]. It exists in different oxidation states such as selenite (IV; SeO32−), selenate (VI; SeO42−), elemental Se (0), and organic forms such as dimethyl selenide, methyl selenide, selenomethionine, selenocysteine. According to Paul and Saha (2019), California (38%), Ireland (32%), Punjab (8%), Jaipur (9%), and China (2%) are the most selenium-polluted regions in the world. Selenium is released to the environment by natural and anthropogenic activities such as volcanic eruption, mining, weathering of rocks, coal mining (so-called selenium curse of the eastern range of the north American Rocky mountains) and combustion, and effluent waste by some industries and agriculture leading to polluted areas [16,17,18,19,20,21,22].
There is a narrow limit between selenium toxicity and the amount required for human health. Paradoxically, less than 40 µg/day of selenium is essential for the body and more than 400 µg/day is toxic to the human body [23]. Daily intake of selenium in food varies from 0.055 to 0.4 mg per day, required for crucial body functions such as antioxidant defense, protein folding, and cell signaling [24,25]. Some primary selenoprotein genes in mammals have central roles in Redox signaling (GPX1, GPX3, GPX4, TRXRD1, TRXRD2), Protein folding and degradation (SEP15, SELS), and metabolism (SEP1, SPS1, SPGS) [26,27]. Selenium deficiency is associated with Keshan disease [28,28,29,30,31], muscle weakness [32], Kashin Beck [33,34,35,36,37], cardiomyopathy [38,39,40,41], and redox dysregulation [42]. In contrast, exposure to excessive amounts of selenium can lead to disorders such as selenosis, loss of hair and nails, redox dysregulation, mitochondrial dysfunction, and cell growth inhibition [43,44,45,46,47,48].

1.2.2. Arsenic

Arsenic (As, group V periodic table) is the 20th most abundant elements in the earth’s crust with a terrestrial abundance of about 1.5–3 mg/kg and an average abundance of about 5 mg/kg [15]. Anthropogenic and natural activities are the main sources of arsenic pollution in the world. The groundwater of different regions of Asia (Bangladesh [49,50,51,52], India [53,54,55,56,57,58], China [59,60,61,62,63,64], Nepal [65,66], Cambodia [67], Vietnam [68], Myanmar [69], Pakistan [70,71,72,73,74,75], and Indonesia [76]), North and South America (USA [77,78], Canada [79,80], Argentina [81], Chile [82], and Mexico [83]), Europe (Hungary [84]) and Africa (South Africa [85]) are contaminated by arsenic [86]. Arsenic occurs in different oxidation states in nature, including arsenate (+5, AsO43−), arsenite (+3, AsO33−), elemental arsenic (0), and arsenide (−3). Akin to SeO32−, AsO33− is the most toxic form among arsenic oxyanions in the environment [87,88]. Arsenic-containing compounds were applied in the manufacture of glass [89], semiconductors [90] and alloys [91], herbicides [92,93], wood preservatives [94], pesticides [95], animal feed additives [96], and medicine [97]. Disruption in cell signaling [98], reactive oxygen species (ROS) generation [99], high affinity to protein thiols or vicinal sulfhydryl groups [100], interruption in the binding of some hormones to their receptor [101], and prevention of oxidative phosphorylation [102] are the main effects of arsenic oxyanion on cells [88].
Depending on the concentration, oxidation state, and exposure time, arsenic can cause health problems such as cancer (skin [103], lung [104,105], bladder [106], and liver [107]), skin lesions [108], PNS (Peripheral Nervous System) disorder [109], liver failure [110], leukopenia [111], circulatory disease [112,113], anemia [114,115], and death [116,117].

1.2.3. Boron

Boron, the fifth element in the periodic table (group III), is a ubiquitous element in the environment that comprised an average concentration of around 10 mg/kg of the earth’s crust. Na2B4O5(OH)4·8H2O is a common form of boron in ores widely distributed in California and Turkey [118]. Argentina, Russia, Chile, Peru, China, Libya, Egypt, Iraq, Morocco, and Syria are other countries containing many boron deposits [119,120]. This element forms approximately 230 compounds, and natural and anthropogenic activities such as volcanoes, commercial uses, fertilizers, wastewater treatment plants, forest fires, and coal combustion can release it into the atmosphere [121]. Boron has very useful applications in some industries such as the manufacture of glass and ceramics, fertilizer and detergent [122,123,124,125].
This essential element plays a pivotal role in immune response [126,127], mineral metabolisms [128], and the endocrine system [129]. It can also inhibit osteoporosis in postmenopausal women [130] and decrease cardiovascular disease [131]. Similar to other metalloids, it can be toxic in higher concentrations and causes many problems such as an increase in the oxidative state of a cell, DNA damage, impairment of DNA repair systems and membrane functions, or the inhibition of protein folding, protein function, and activities in living organisms [119,132,133]. Boron is found in some antibiotics such as boromycin [134], tartrolons [135], aplasmomycin [136], and biomolecules such as the bacterial autoinducer 2 (AI-2] vibrioferrin (a siderophore) [137] and borolithochromes (pigment in algae Solenopora jurassica) [138]. Additionally, some nitrogen-fixing bacteria need boron as a cofactor for growth and nitrogen fixation [119].

1.2.4. Antimony

Antimony belongs to subgroup 15 of the element periodic table (atomic number 51) with average concentrations less than 1 μg/L in nature. Most of the world’s antimony reserves are located in South Africa, China, Russia, Bolivia, Tajikistan, and Mexico [139]. Due to the chalcophilic nature of antimony and its presence in ores containing chalcogen, smelting, and mining of ores containing these compounds, especially sulfide ores [140] are among the polluting sources of this element [141].
Antimonate (Sb (V); Sb(OH)6) and antimonite (Sb (III); Sb(OH)3) are the two common inorganic forms of antimony present in natural waters, and Sb(OH)3 is more toxic than the other one [142]. Antimony’s other toxic compound is antimony trichloride (SbCl3), used in alloys, as a constituent of paint pigments, and in rubber compounding. Other major applications of antimony are included in various industries and use such as semiconductor, alloys, batteries, catalyst, and medicine [143,144,145]. In ancient times, antimony was used to purify precious metals such as gold and silver. Furthermore, antimony is used with other compounds to make textiles, paper, and plastics as a fire retardant agent [140].
Antimony is not present in living systems and, such as arsenic, is highly toxic to humans and living organisms. Eye, skin, lung, mucous membrane irritation, oxidative DNA damage, pneumoconiosis, and increased lung, heart, and gastrointestinal diseases are some problems caused by long-term exposure to antimony [146,147]. Antimony can affect the nitrogen cycle in soil by influencing urease function under pH 7 [148].

1.2.5. Tellurium

Tellurium oxyanions are another highly toxic metalloid that studies in this review. Tellurium is an element that belongs to the 16th group of the periodic table with atomic number 52 and has two allotropic forms, including white crystalline metal and black amorphous powder [149,150]. The concentration of tellurium in the earth’s crust is very low and about 1–5 µg/kg [151]. Tellurium is found as oxyanions tellurite (IV; TeO32−) and tellurate (VI; TeO42−). Tellurium is found in industries such as petroleum refining plants, glass, electronic and photoelectronic industries, optics, and sensors [152]. Tellurium can be found in a variety of ores as well as coal. Another tellurium application is in medicine and has traditionally been exploited as an antimicrobial agent in treating some infectious diseases, including leprosy, tuberculosis, dermatitis, cystitis, and severe eye infections [153]. Tellurium is also used in labeling, imaging, and targeted drug delivery systems and has some anti-inflammatory, anti-fungal, anti-leishmaniasis, and immunomodulatory activities [154,155,156,157,158,159]. Exposure to a high amount of tellurium can cause several health issues, such as respiratory irritation, headache, drowsiness, weakness, malaise, lassitude, gastrointestinal symptoms, dizziness, and dermatitis [160]. Both Se and Te are mixed with Cadmium to make quantum dots (used in phone and TV screens) and photoreceptors in solar cells leading to concerns of their disposal and subsequent release into water systems [161].

1.2.6. Other Metalloids

Germanium is another metalloid that belongs to group 14 and period 4 of the periodic element table. This metalloid is ranked 54th among the most abundant elements in the earth’s crust and has two stable oxidation states +2 and +4 in nature. Only a few compounds of germanium such as GeO2, GeH4, GeCl4, and GeF4 have toxic properties and their organic forms have no effect on human health. Due to similar outer electron structure and properties, Germanium is also called a pseudo isotope of silicon [162,163]. Germanium is used in small quantities in some fields such as fiber optics [164,165], micro- and nano electronics [166], infrared detectors [167], and polymerization catalysts [168].
After oxygen, silicon (Si) with 27% is the most abundant element in the earth’s crust. This element is found in many human organs and its deficiency is related to infection and bone weakness. Si is mostly biologically inert, and it can be used as a drug carrier in ointments and hydrogel coatings in medical devices [169]. There are no specific reports on the devastating effects of astatine and silicon in the literature.

1.3. Methods of Bioremediation

Several methods have been suggested to eradicate metalloid contamination in various environments and locations including physical, chemical, and biological methods. All methods remediate and detoxify; however, bioremediation is beneficial for removing the toxic effects of metalloids and recovering them for industrial applications. This method gains increasing interest in recent years due to its cost-effectiveness, eco-friendly, and simple operable manner. Furthermore, this method needs no special technical equipment/skills and typically does not result in any additional waste and toxic compounds. Bacteria have the primary role in microbial bioremediation because of their small size, high numbers, fast growth, suitable adaptation to different environments, and manageable operating and culture condition. The main factors affecting bioremediation are included, organometallic ions, soil type, type and concentration of contaminants, nutrient availability, redox potentials (Eh), pH, moisture, solubility, microbial community and interaction, electron acceptors, and temperature [170,171,172].
Bioremediation is categorized into in-situ and ex-situ methods (for further information on this process please see [171,173]). Using fungi in a process called mycoremediation, is another approach for the effective eradication of metalloids. Bioremediation with algae and microalgae is called phytoremediation. Phytoremediation refers to bioremediation based on plants, a category that utilizes pollutant hyperaccumulator plants. In combination with plants (phytobial remediation), endophytic bacteria include another biological approach used for the bioremediation of polluted regions [174,175,176,177].
The main aim of bioremediation is to decrease the bioavailability of contaminants in environments and minimize their toxic and adverse effects by using living organisms, specifically bacteria and fungi [178,179]. There are different mechanisms for removing metal/metalloids from contaminated areas, such as bioaccumulation, biosorption, biotransformation, bioleaching, biomineralization, and participating microorganisms in metal/metalloid chemisorption. Biosorption is an energy-independent and reversible process in which pollutant ions interact with specific biochemical functional groups exposed on the surface of the cell wall and outer membrane of microorganisms [180]. Bioaccumulation is an energy-dependent process, which can be irreversible, that ions enter into the cell and accumulate intracellularly after binding to the cell via some particular transferring proteins and channels [181]. Nowadays, interest focuses on biotransformation, a process based on oxidation-reduction reactions. Since the toxicity of metalloids depends on their oxidation states and can be changed by losing or adding electrons, metalloid-reducing and oxidizing microorganisms gain much attention in the recent investigation on bioremediation of contaminants [182]. The main mechanisms of metal resistance in microorganisms are summarized in Figure 1 [183].
Bioleaching refers to a process using microorganisms to alter the solubility of ions and make them soluble forms to facilitate their removal from an ore [184]. Another process based on precipitation called biomineralization can be applied for increasing the effectiveness of other mechanisms for eradicating metal/metalloid toxicants [185]. The latter mechanism refers to facilitating and accelerating the rate of chemisorption and precipitation with microorganisms. A review article by Shahid Sher and Abdul Rehman (2019) described these bioremediation strategies for arsenic [177].
Here in this review, we examine the biodiversity and ecology of bacteria and fungi involved in metalloid bioremediation. A few similar articles are studying both the biodiversity of microorganisms present in metalloid-contaminated areas and the genes, transporters, and mechanisms involved in metalloid bioremediation. Most of these studies usually investigated only one metalloid, such as arsenic or selenium. Furthermore, microbial communication in consortia and microbial aggregates is an important issue that plays a crucial role in the superiority of a consortium over culturing one microorganism for removing metalloids, which received little attention in papers. Therefore, genes and proteins involved in the various bioremediation processes using bacteria and fungi are overviewed. Finally, the review summaries the progress in metalloids’ bioremediation and mechanisms participating in this process by microbial aggregates, biofilm, and consortia.

2. Ecology and Diversity of Metalloid-Resistant Microorganisms

2.1. Microbial Diversity of Metalloid-Contaminated Regions

Microorganisms adapt to the microenvironment, live together in consortia and biofilms, and interact with each other and abiotic parts of the ecosystems [186]. By studying microbial diversity and the basic scientific report, we can better understand microbial community structure and its function. Some anthropogenic activities such as mining, agriculture, and industrial wastewater, the release of organic and inorganic pollutants, pesticides, urban and rural affluent, city development, and so on, can affect microbial diversity and shift the community structure into specific bacterial and fungal taxa. Microbial inhabitants of these areas can tolerate high levels of pollutants, and some of them have a crucial role in nutrient cycling and remediation of contaminated regions [187]. Below follows examples of microbial diversity in relationship to oxyanion pollutants and bioremediation.

2.1.1. Microbial Communities in Selenium Oxyanion-Contaminated Sites

Due to explosive activities in mines, Se oxyanions can be associated with nitrate (NO3). Subedi et al. (2017) used a bacterial consortium from two mine sites in Canada to investigate simultaneous denitrification and SeO42− reduction. Metagenomics analysis of inoculum showed that Pseudomonas, Lysinibacillus, and Thauera were dominant operational taxonomic units (OTU) in biomass, respectively. The minor OTUs, such as Exiguobacterium, Tissierella, and Clostridium, were increased after the addition of NO3 in culture media. Results indicated that a consortium including SeO42− and NO3-reducing bacteria could be effectively applied for removing both SeO42− and NO3 pollution at the same time [188].
Granular sludge from brewery wastewater containing microbial aggregates can be used to transform selenium oxyanion into selenium nanoparticles (SeNPs). The granular sludge’s microbial structure has been revealed by the high-throughput method, and the results showed that members of the families Veillonellaceae and Pseudomonadaceae, and Clostridiaceae are frequent in these aggregates. Both Veillonellaceae and Clostridiaceae can use lactate as a carbon source and electron donor and transform it into acetate, while Pseudomonadaceae members consume this acetate in a syntrophic relationship. Another interesting issue is the significant increase in the abundance of Pseudomonas species (from 0.03% of inoculum to 10% in culture medium) after three weeks of incubation in the presence of SeO32−. Although the genus Pseudomonas is known primarily as an aerobic bacterium, its presence and increase in granular sludge as an anaerobic condition is a curious outcome. This is an example of metabolic cooperation between species present in a consortium. Outer membrane porin and cytoplasmic elongation factor Tu were associated with the selenium nanosphere, indicating intracellular nanoparticle biosynthesis and subsequent secretion [189].
In another study, SeO32−-reducing bacteria were isolated from wastewater sludge in the presence of 100 mM SeO32− and lactate as an electron donor. Citrobacter sp. and Providencia were identified by 16S rRNA sequencing after spreading on BSM culture (basal salt medium) plates. Further studies on isolates revealed that acetate and lactate were better electron donors and carbon sources than propionate, butyrate, and glucose [190].
Thirty-eight selenium-resistant bacteria were isolated from different SeO32− and SeO42− polluted areas in India, ranging from 2.25–3 mg/kg in soil and 0.08–1.10 mg/L in water samples. Eight morphologically different strains were chosen for polyphasic taxonomy analysis, including biochemical tests, 16S rRNA sequencing, and FAME (Fatty acid methyl ester) methods. Two and three strains phylogenetically belong to the genera Delftia, and Bacillus, respectively. Other strains were related to the genera Achromobacter, Pseudomonas, and Enterobacter. Minimum inhibitory concentrations (MIC) of these strains were determined by 300–750 mM which showed the enormous potential of these microorganisms in bioremediation of selenium oxyanion polluted sites [191].
Pseudomonas is one of the most predominant genera present in the selenium oxyanion-contaminated regions. This bacterium can reduce SeO32− to SeNPs and is well known for its potential ability for a high amount of selenium oxyanion resistance. Because of its specific reductase enzymes and capability in NADPH equivalent production, Pseudomonas is a suitable candidate for the bioremediation of metalloids [192,193]. Another group is Rhodococcus, which has particular capabilities in the biotransformation of metalloids such as changes in the sensitivity of cellular targets, tolerance of high concentrations of metalloids, biodegradation of hydrocarbons, and the production of siderophores to bind to metalloids such as AsO33−. In addition, Rhodococcus contains operon and genetic elements, mycothiol, phosphate transporters, which play an essential role in reducing the toxicity and biotransformation of metalloids, reducing oxidative stress, and the entry of metalloids into the cell, respectively. Other notable abilities of this bacterium include the bioconversion of the oxyanions into Se and Tellurium nanoparticles (TeNPs) and nanorods [25,149,194,195,196,197,198].
Sediment samples were collected from New Jersey (USA) and Chennai (India) and were enriched with 10 mM sodium SeO42− in MSM medium and pyruvate or 4-hydroxy benzoate as a carbon source. SeO42− was the sole electron acceptor in culture media. After culturing on solid media, isolated strains were selected considering different sizes, shapes, and morphology. A 16S rRNA analysis was applied for the molecular identification of selected strains. Pseudomonas stutzeri and three new species were characterized that phylogenetically related to Geovibrio ferrireducens, Pelobacter acidigallici, and Chrysiogenes arsenatis with 99, 98, 97, and 94% similarity, respectively. All strains prefer pyruvate as an electron donor in comparison with 3-hydroxy benzoate, benzoate, and 4-hydroxybenzoate. NO3 and SeO42− also are electron acceptors in the respiration process. This study indicated that SeO42−-respiring bacteria are phylogenetically and physiologically diverse and ubiquitous in aquatic sediments [199]. In another study conducted to evaluate selenium oxyanion pollutants, two phosphate mines in Idaho with different Se concentrations targeted high throughput sequencing to understand these reclaimed sites’ bacterial and fungal diversity. Ascomycota, Basidiomycota, Zygomycota, Chytridomycota, Glomeromycota, and Rozellomycota were abundant fungal phyla isolated from different Se-contaminated sites, respectively. In contrast, Actinomycetota, Alphaproteobacteria, Bacteroidota, Acidobacteriota, Chloroflexota, Verrucomicrobiota, Planctomycetota, Deltaproteobacteria, Gemmatimonadota, Betaproteobacteria, Firmicutes, and Gammaproteobacteria were dominant in bacterial strains. Microorganisms involved in the nitrogen cycle are dominant in sequences that indicate that there could be a connection between nitrogen and selenium cycles in this ecosystem [200].

2.1.2. Bacterial Communities in Arsenic Sites

The contaminant concentration is a crucial factor that affects the microbial community structure of the polluted area [173]. In this regard, Valverde et al. (2011) studied the microbial community shift in response to different concentrations of arsenic in the Terrubias mine in Spain by the culture-dependent method. The results showed that Bacillota and Gammaproteobacteria are the most frequent taxa, and Bacillus and Pseudomonas are the dominant genera in this region, respectively. In the presence of a high concentration of arsenic in soil samples, the bacterial diversity of samples shifts to the preponderance of Firmicutes [201].
The microbial diversity of the As-contaminated groundwater of Assam (India) was studied by the culture-dependent method. In this study, 89 bacterial strains were isolated, and 16S rRNA sequencing analysis was applied for microbial community characterization. The results were shown in Table 1. Brevundimonas (35%) and Acidovorax (23%), Acinetobacter (10%), Pseudomonas (9%) were the dominant genera isolated from this region. The other genera include Undibacterium, Herbaspirillum, Rhodococcus, Staphylococcus, Bosea, Bacillus, Ralstonia, Caulobacter, and the member of the order Hyphomicrobiales [202].
Due to wastewater release in marine environments, the estuary is prone to contamination, including metalloids. In this regard, Jackson et al. (2005) studied AsO43−-resistant bacteria isolated from Lake Pontchartrain estuary, Louisiana, USA. The amount of arsenic in samples was 7.02 µg/g sediment. Thirty-seven isolated strains were selected for molecular characterization. These isolates belong to the phyla or classes Gammaproteobacteria, Betaproteobacteria, Alphaproteobacteria, Bacillota, Actinomycetota, and Bacteroidota. The results are shown in Table 1 [203].
In the case of bioremediation, there are many studies investigating arsenic contamination in aquifer samples. Three groundwater samples from As-rich sites of Chianan plain, Taiwan were exploited for the detection diversity of dissimilatory AsO43− reducing microorganisms with the pyrosequencing method. In this study, Bacillota were dominant at Arsenic high concentration areas, and Proteobacteria was frequently observed in other groundwater samples with lower arsenic concentrations. The genera Acinetobacter and Bacillus are dominant in all three samples. The results showed that lactate is the favorable carbon and electron source and arsenic reducing bacteria constitute the leading performers, and sulfate-reducing bacteria (SRB) play a negligible role in the bioremediation of arsenic [204].
Molecular identification of arsenic-rich sites of the Latium region of Italy showed that Gammaproteobacteria are the dominant taxon in arsenic-tolerant microorganisms in thermal waters, while Alpha- and Beta- proteobacteria are prevailed in surface and ground waters [206]. In another study of arsenic-rich areas in Italy, the abundance of bacterial phyla and classes was as follows: Gammaproteobacteria > Betaproteobacteria > Bacillota > Alphaproteobacteria > Bacteroidota. Pseudomonas and Comamonas were the dominant genera in sequenced strains [206]. Most taxa in deep-sea sediments of the Indian Ocean Ridge were mainly related to Proteobacteria and Actinomycetota, especially Microbacterium esteraromaticum [207].
According to Table 1, different classes of Proteobacteria are dominant in arsenic-contaminated regions. Among them, Alphaproteobacteria and Betaproteobacteria are predominant in all samples. Acinetobacter, Undibacterium, Brevundimonas, Janthinobacterium, and Bacillus were the remarkable genera in all samples.

2.1.3. Communities of Antimonite Sites

Samples were taken from Sb-contaminated areas of Guizhou province (China) to determine microbial diversity. The results indicated that the phyla Pseudomonadota (60.5% of all sequences) and Actinomycetota (10.3% of all sequences) were prevalent in all samples. Chloroflexota, Cyanobacteria, Acidobacteriota, and Gemmatimonadota were the other phyla presented in the samples. The most dominant taxa included Thiobacillus, Limnobacter, Nocardioides, Lysobacter, Phormidium, and Kaistobacter. Besides, archaeal phyla such as Euryarchaeota and Crenarchaeota comprised only 0.4% of all sequences [208]. In another study, Wang et al. (2018) investigated the microbial diversity of the Xikuangshang Sb mine in China which is known as the capital of antimony around the world. Proteobacteria, Acidobacteriota, Chloroflexota, Bacteroidota, Actinomycetota, Gemmatimonadota, and Cyanobacteria were abundant phyla detected from this region [209].

2.2. Metalloid-Resistant Bacteria from Metalloid-Contaminated Sites

Here we will overview specific species and strains isolated from metalloid-polluted sites that display strong resistance to this class of elements. Maltman et al. (2016) investigated the biodiversity of metalloid respiring bacteria associated with Riftia pachyptila inhabitants in hydrothermal vents. They isolated 107 bacteria, and 106 of them can use at least one metalloid oxyanion, such as TeO42−, TeO32−, SeO42−, and SeO32−, as a last electron acceptor. Epibiont diversity of different locations of Juan de Fuca Ridge Black Smoker Field showed the predominance of Curvibacter, Shewanella, Pseudomonas, and Pseudoalteromonas in Explorer Ridge and Vibrio, Pseudoalteromonas, Curvibacter, and Shewanella in Axial Volcano, respectively [210].

3. Metalloid Tolerance Genes and Mechanisms

3.1. Arsenic

The arsABC operon is a set of genes that give tolerance to arsenic oxyanion in both Gram-negative and Gram-positive bacteria such as Staphylococcus, Bacillus, Rhodococcus, and E. coli [196,211,212,213]. This operon is located in plasmids, as well as chromosomes, and encodes an arsenate reductase asrC, an enzyme that transforms AsO43− into the AsO33− form. Other protein encoded by this operon in Rhodococcus includes asrB (As (III) specific efflux integral membrane protein), asrA (Anion translocating ATPase protein, asrA + asrB constitute an efflux pump), asrD (Regulatory protein, a chaperon protein), and asrR (Regulatory protein, repressor) [196]. Entrance to the cell has been discussed [214]. GlpF is an aquaglyceroporin transporter specific for organic solutes such as urea and glycerol and is used to transport AsO33− across the membrane. Due to the structural similarity to phosphate, AsO43− uptake to the cell occurs through phosphate transporters, including pit (phosphate inorganic transport) and pst (phosphate-specific transport). The pit is the primary transporter protein that gains energy from the proton motive force. In contrast, pst is an ABC-type periplasmic transporter that facilitates the uptake of AsO43− less efficiently [215]. Down-regulation or mutations in these transporters can give rise to As oxyanion resistance. AsO43− can be an electron acceptor in dissimilatory AsO43− reduction and energy generated by coupling reduction of AsO43− with inorganic and organic substances’ oxidation. Arr, a heterodimer membrane-bound enzyme, catalyzes this process in Desulfurispirillum indicum strain S5. Arr enzymes constitute two catalytic subunits including ArrA (Molybdopterin center + [4Fe-4S] cluster) and ArrB ([Fe-4S] center + 3 or 4 [4Fe-4S] clusters). In the presence of oxygen and NO3 as electron acceptors in the respiration process, the expression of Arr genes is repressed [215,216]. ArrA and ArrB also are found in both the anaerobic Gram-positive bacterium, Chrysiogenes arsenatis, and the Gram-negative bacterium Shewanella sp. Strain ANA-3 [217].
The expression of three genes, including ACR1, ACR2, and ACR3, provides the ability to tolerate arsenic oxyanions in Saccharomyces cerevisiae. ACR1 encoded a transcriptional regulator while ACR2 and ACR3 are arsenate reductase and plasma membrane AsO33− efflux pump, respectively. Overexpression of cytoplasmic thiols and pumping As (III) out of the cell followed by reducing As (V) is another resistance mechanism observed in S. cerevisiae. In addition to this mechanism, fungi can sequestrate arsenic oxyanions in vacuole intracellularly. Bacteria and fungi methylate arsenic oxyanions and transform them into volatile species to decrease their adverse effects. In a high concentration of AsO43−, the level of GiPT transcript elevated, and a few hours later, GiArsA (putative AsO33− efflux) level increased. GiPT acts as a high-affinity phosphate transporter that uptake AsO43− into the cell. Mechanisms of AsO43− reduction in AMF is still unclear. It was suggested that an arsenate reductase might exist and transform AsO43− into AsO33−. Then, AsO33− expels from the cell by GiArsA and GvArsA as AsO33− efflux pumps [218]. As a result of toxic metal accumulation, oxidative stress increases. Microorganisms have evolved some strategies to conquer the adverse consequences of oxidative stresses. In Aspergillus niger, the concentration of cytoplasmic thiols, malondialdehyde, proline, and anti-oxidative stress enzymes such as catalase, superoxide dismutase, and succinate dehydrogenase elevated as a result of an increase in AsO33− up to a certain level. In this study performed by Mukherjee et al. (2010), A. niger was exposed to different AsO33− doses ranging from 25 to 100 mg/L. The results showed that growth rate and thiol content were maximum at 25 mg/L and 100 mg/L AsO33−, respectively. The most level of AsO33− absorption and intracellular AsO33− was observed at 75 mg/L.
Different efflux proteins such as ArsB, Acr3, ArsP, ArsJ, and MSF1 were evolved in arsenic-resistant bacteria for arsenic resistance. Some metalloids such as AsO33−-, AsO43−-, Sb(OH)3-, and methyl AsO33−-induced the expression of arsK. arsK is a novel efflux system for metalloid transportation and is regulated by ArsR2. arsK and arsR2 are integrated into one operon and transcribed under one promotor’s regulation [219]. Different mechanisms in the entrance and transportation of arsenic in E. coli and S. cerevisiae is summarized in Table 2. It is suggested that this mechanism also is used for antimony uptake [142].
Major intrinsic proteins (MIPs) are one of the transporter systems that participated in metalloid homeostasis in microorganisms. MIPs can expel them from the cell at a high level of metalloids and assimilate them when microorganisms need a beneficial metalloid. In S. cerevisiae, As (III) is conjugated to GSH, and Ycf1p has transported them by vacuolization. In S. cerevisiae, As (III) conjugated to GSH, and Ycf1p transported them by vacuolization. The presence of As (III) in the medium initiates a MAP kinase signaling cascade that can be started with As (III) and leads to inhibition of As (III) influx by Fps1p closure. Some genes are involved in metalloid uptake, which was concluded in Table 3.

3.2. Selenium

Studies showed that microorganisms often used sulfate-thiosulfate permease and the SulP- type permease to transport selenium compounds. In addition to these proteins, other research revealed that membrane protein YedE might also be applied for selenium oxyanion transport [231].
Transport of SeO32− in E. coli occurred by at least three systems, including Sulphate transport complex ABC, sulfate permease, and one or more unknown transporters that even worked when the other two transporters were inactivated. cysAWTP operon encodes sulfate transport complex ABC and consists of CysT and CysP. A polyol transporter found in Rhodobacter sphaeroides, is responsible for transferring SeO32− into cells [232,233].
Many genes were found that are involved indirectly in SeO42− reduction, such as fnr, tatC, and menD. These genes have central roles in expressing anaerobic genes, the exportation of folded proteins across the inner membrane, and the menaquinone biosynthesis pathway, respectively. Theisen and Yee (2014) found a functional selenate reductase gene, namely ynfE, in Citrobacter fereundii that encodes a molybdenum-binding Tat-secreted protein. An FNR binding site was found upstream of the ynfE gene, suggesting that SeO42− reduction in this bacterium occurred in anaerobic conditions. ynfE gene is located in the ynfEGHdmsD operon [234], and a high structural identity is observed between the selenate reductase of this bacterium and selenate reductases found in Gammaproteobacteria [235]. FNR is a general regulator system for fumarate and nitrate reductases and well regulates all CISMS (complex iron sulfur-molybdo enzymes). The menaquinone is the electron carrier to CISMS and TatC is a protein translocase that accepts CISM enzymes to help in assembly at the cytoplasmic membrane. Thus, both FNR and TatC are required for the YnfE protein to mature but have nothing specifically to do with SeO32− metabolism, and SeO32− can be processed by CISM-type enzymes including even NarGHI in E. coli [236].
In anaerobic conditions, the fnr gene is responsible for regulating selenate reductase activity in Enterobacter cloacae SLD1a-1. Recombinant Escherichia coli S17-1 containing the fnr gene from E. cloacae can reduce SeO42− to Se (0) with similar rates to those for E. cloacae. Even SeNPs biosynthesized by this bacterium have the same size and shape as SeNPs produced by E. cloacae. Lack of SeO42− reduction activity and SeNPs biosynthesis are observed in E. cloacae mutants in the fnr gene [237].
Another operon involved in SeO42− reduction is the srdBCA operon, which encodes a respiratory selenate reductase in Gram-positive bacteria such as Bacillus selenatarsenatis. Nitrate and nitrite reductase, protein, or peptide-containing thiol groups, chaperones, and some elongation factors also may be involved in selenium oxyanion transformation to Se (0) [238].
The SerABC complex is a soluble periplasmic enzyme in Thauera selenatis, consisting of three subunits, including SerA, SerB, and SerC, and a b-type cytochrome. In contrast, this complex in E. cloaca is a membrane-bound enzyme, and molybdenum, iron, and sulfur are the prosthetic groups of enzymes. SerABC complex is a selenate reductase that can only reduce SeO42− and showed substrate specificity [239,240].
The ygfKLMN operon encodes another selenate reductase in E. coli K12. The deletion of the ygfk gene in this operon results in the loss of SeO42− reduction in this strain. Besides enzymes involved in SeO42− and SeO32− reduction, the mutation in genes involved in electron carrier biosynthesis, such as menaquinone, can also affect selenium oxyanion’s bioreduction activity. This is because most all of the selenate and selenite reductases are complex iron-sulfur molybdenum enzymes that interact with the quinone pool anaerobically where menaquinone dominates over quinone [241]. Menaquinone biosynthesis is performed by menFDHBCE gene cluster, menA, and ubiE, menD, menC, or menE gene mutants were unable in SeO42− reduction [242].

3.3. Boron

The transportation of some metalloids such as AsO33−, Sb(OH)3, boron, and silicon across the membrane facilitates by MIPs in yeast. Furthermore, it was suggested that nodulin 26-like intrinsic protein (NIP) subfamily proteins facilitate these metalloids’ transportation. Borate is a carbonate analog and can be transported across the membrane by bicarbonate transporter-like BOR1 in plants. YNL275w is a homolog of BOR1 in Saccharomyces cerevisiae and also can be applied as an efflux boron transporter in the yeast cell membrane [118]. In contrast, AsO43− and Sb(OH)6 are phosphate analogs, and transcription of them is facilitated by phosphate transporters [243].

3.4. Tellurium

Different enzymes have been found in various organisms to participate in TeO32− reduction [244] and consist of: nitrate reductases (Nar and Nap) [245], catalase, NADH dehydrogenase (type II), and dihydrolipoamide dehydrogenase (lpdA), alkyl hydroperoxide reductase (AhpF), thioredoxin reductase (TrxB), dihydrolipoamide dehydrogenase, glutathione reductase (GorA), NADH: flavorubredoxin reductase (NorW), mercuric reductase (MerA), and the putative oxidoreductase YkgC [246]. The tehAB genes involved in TeO32− resistance is found on the chromosome of E. coli where tehB is S-adenosylmethionine cofactor dependent and likely a telluromethylase [247,248]. In addition to the phosphate transporter, TeO32− can enter the cell via a monocarboxylate transporter in R. capsulatus, and acetate, lactate, and pyruvate can compete with TeO32− in this process [249]. The ter operon consists of seven genes, ZABCDEF, but little is understood about the mechanism, yet it seems to be associated with the E. coli pathogenicity island [250]. The terD gene seems to be able to give resistance on its own and is found in other species [250,251,252]. In another study, Maltman et al. (2017) found a new tellurite reductase with a 117 kDa molecular weight. In addition to TeO32−, this membrane-associated enzyme can reduce TeO42− [253]. Other gene involved in TeO32− resistance includes kilAtelAB [254]. Many of these are found on large conjugative plasmids [255].
Some transporters and pathways in metalloid biotransformation in bacteria and fungi were summarized in Figure 2 and Figure 3, respectively.

4. Remediation under More Than One Metal (loid) Challenge

For bioremediation of contaminated areas with two or more metalloids, it is essential to know the microbial diversity and type of bacteria present in these regions. Indigenous microorganisms have different mechanisms for tolerating high levels of these contaminants, so their potential can be used for bioremediation. For example, the diversity of diazotrophic bacteria was investigated in two mines in China with high levels of arsenic and antimony. Due to low amounts of nitrogen, diazotrophic bacteria are important, as they are capable of both nitrogen fixation and metalloid tolerance in these regions. The most abundant diazotroph phyla include Pseudomonadota and Cyanobacteria, and Sinorhizobium, Dechloromonas, Trichormus, Herbaspirillum, Desmonostoc, and Klebsiella are prevalent genera in these areas [256].
Liu et al. (2017) investigated the simultaneous bioremediation of wastewater containing arsenic and antimony oxyanions by SRB. They used lactate and ethanol as carbon sources. Results showed that ethanol was a better electron donor and carbon source than lactate and it might be related to the lower amount of sulfide formation by ethanol in comparison with lactate consumption. In presence of both HAsO42− (As (V)) and Sb(OH)6 (Sb (V)), the rate of As (V) removal was low and it improved remarkably by adding ferrous ions (Fe (II)). They suggested that sorption/co-precipitation by FeS is involved in increasing As (V) remediation by SRB [257].
An interesting response of microorganisms is observed in response to the simultaneous presence of contaminants such as TeO32− and SeO32− or TeO32− and mercury. In both cases, the presence of mercury or SeO32− with TeO32− increases the bacterial resistance to TeO32− in some microorganisms. The mechanism and cause of these reactions are still unknown [258,259]. In fact, it appears that SeO32− will actually protect against TeO32− challenge [260].

5. Metalloid Reduction through Biodegradation of Associated Organic Compounds

Microbial tolerance to metalloids has been investigated for over 40 years. Most of these investigations have concentrated on planktonic forms of single-species microorganisms [261,262,263,264,265,266], while microbial consortia and biofilms are the common forms of microorganisms present in the environment where “natural” exposure to metalloids occur [256,267,268,269,270,271,272,273]. A microbial biofilm comprises one or more species generally attached to a surface and in addition to the microbial cells, contains a considerable amount of extracellular matrix of exopolymeric substance (EPS; also referred to as exopolysaccharides). The matrix is a mixture of various biochemical polymers (nucleic acids, proteins, carbohydrates—EPS often also refers to exopolysaccharides) [274]. This matrix is of a defense nature but can also be used as a carbon source as well as a source of functional groups to bind ions and organic compounds. It acts as a barrier to the entry of some inhibitory compounds such as toxins and enables the exchange of electron and genetic information thorough its matrix [275]. Studies have shown that biofilms have a remarkable ability to remove various contaminants than planktonic species [276,277]. In addition to biofilm, microorganisms can usually form microbial consortia which referred to communities composed of more than two different species of microorganisms that can be used for different biotechnological applications such as bioremediation. These consortia can tolerate and remove significant amounts of pollutants rather than pure culture because of synergistic interactions between microbial species composing these consortia [278,279,280,281,282,283,284,285].
The idea of a biofilm community is illustrated by the ability of a consortium consisting of methanotrophic bacteria and archaea to reduce SeO42−. The use of consortium resulted in a 1.8-fold increase in SeO42− reduction compared to the use of each of the bacterial and archaeal species separately [286]. This subject gained limited attention in the literature at the time. Precise underlying mechanisms are still unknown but now has attracted much more interest in recent years.
For designing optimized co-culture, a great extent of information is required about microbial community structure, metabolic requirements, interactions, and mechanisms. Consortia are composed of greater than two different microbial species that co-exist and interact together. These interactions between microorganisms are always negative, positive, or neutral and are categorized into six different forms: cooperation (commensalism and mutualistic relationships), prey and predator or parasite and its host, neutral interaction, competition, and amensalism. Cooperation is the best mode for consortia formation because of cross-feeding (syntropy) and metabolite exchange between these consortia [287,288].
In an exciting state of cooperation, one microorganism produces compounds and provides them to another microorganism, while another microorganism fulfills another microbe’s requirements by performing a similar action [289,290]. A microorganism may produce growth factors or enzyme-inducing compounds that cause more growth or more production rates in another microorganism. Sometimes microorganisms produce more products or increase the biomass of other microbes present in the consortium by removing or reducing stress factors such as toxins or inhibitors [291,292,293]. In certain cases, a competitive or antagonistic relationship between microorganisms leads to the production of beneficial products. Finally, the degradation of complex biodegradable compounds such as hydrocarbons by a microorganism in the consortium leads to smaller and easier intermediates for the consumption of other microbes and the formation of a metabolic network between the members of the consortium [288]. The summary of these mechanisms is illustrated in Figure 4.
Some microorganisms can couple reductions of metalloid oxyanions with hydrocarbon oxidation. In this regard, Luo et al. (2018) studied the bioreduction of SeO42− using denitrifying anaerobic methane-oxidizing biofilm. SeO42− was added to culture media with a concentration of 20–60 µmol/L in a membrane bioreactor (MBR) bioreactor with a reduced rate of 2.8–12.4 µmol/L/d. In the lack of methane, no SeO42− was reduced. Fluorescence in-situ hybridization (FISH) and 16S rRNA gene sequencing analysis were applied to detect biofilm microbial community structure. Results showed that Candidatus Methanoperedens nitroreducens and Candidatus Methylomirabilis oxyfera were dominant SeO42− reducers in the presence of methane as a carbon source [294].
In addition to SeO42−, the reduction of SeO32− also can be coupled by the oxidation of methane. In this regard, Bai et al. (2019) used denitrifying methanotrophic microorganisms for bioreduction. The addition of NO3 can inhibit the process. The authors proposed two hypothetical mechanisms for this process: methane oxidation by methanotrophs and the transfer of electrons to SeO32− by the same microorganism. The other involves a synergistic relationship between methanotrophs (electron transfer) and another microbe that can reduce SeO32− [295].
Liu et al. (2013) investigated the simultaneous removal of pyrene and arsenic by bacterial and fungal co-culture. This co-culture consists of two bacterial strains, namely, Bacillus and Sphingomonas, and Fusarium as a fungus. They made five different culture conditions; culturing isolates independently, bacterial culture, and a combination of three strains with arsenic (20 µM) as an energy source and pyrene (100 mg/L) as a carbon source. Results showed that the highest amount of pyrene degradation and arsenic methylation occurred in the presence of a mix-culture of all three strains because of the synergistic effect. The bioremediation of pollutants by co-culture was also decreased to nine days compared to 63 days As volatilization time by bacterial strains (only 13.9% arsenic removal) [296].
Due to diverse reductase enzymes, strains of the genus Pseudomonas strains are competent bacteria in the bioremediation of contaminants. In a study conducted by Feng et al. (2014), these strains’ ability to remove AsO43− and poly aromatic hydrocarbons (PAH) was investigated individually and concurrently. These strains are capable of complete phenanthrene (of 60 mg/L) and half pyrene (20 mg/L) elimination in 60 h and reduction of 1.5 mM AsO43− in 48 h. The arsenic reduction rate will be significantly decreased, while PAH is exploited as the sole source of carbon and energy. The addition of lactate as an electron source increases the ability to degrade simultaneous pyrene and AsO43− reduction in Pseudomonas strains [297].
Despite the biodegradation of a broad spectrum of organic contaminants, Methanotrophs can transform inorganic pollutants such as SeO32− into safe elemental Se. In this regard, Eswayah et al. (2017) studied the ability of pure culture of Methylococcus capsulatus and Methylosinus trichosporium in simultaneous methane oxidation and SeO32− remediation. GC-MS analysis showed that these bacteria could volatize SeO32− in addition to producing elemental Se. Localization of elemental Se showed that this process occurred on the bacterial cell wall. Volatile Se forms produced by Mc. capsulatus include DMSe (di methyl selenide), DMDSe (di methyl di selenide), DMSeS (dimethyl selenenyl sulfide), Methyl selenol, and Methylseleno acetate while Ms. trichosporium can only volatize SeO32− into DMDSe, and DMSeS [298]. Although the pure culture of these two bacteria could not transform SeO42− into its non-toxic form, another study by Lai et al. (2016) showed the ability of cultured mixtures of methanotrophs to remove SeO42−. In this study, the complete bioremediation of SeO42− was achieved using methane as a carbon source and electron donor in a membrane bioreactor (MBfR). Additionally, the effect of NO3 was investigated on the SeO42− bioreduction process. Results showed that NO3 could interfere with SeO42− bioreduction and inhibit this process slightly. The simultaneous reduction of NO3 and SeO42− occurred even when the amount of methane was low. Authors suggested two different mechanisms for this process: i. some methanotroph microorganisms such as Methylomonas can cause couple reduction in SeO42− with the oxidation of methane. ii. The synergistic relationship between methanotrophs and SeO42− reducing microorganisms in this way that some methanotroph microorganisms can oxidize methane and provide electrons for other microorganisms that can reduce senate into elemental selenium (Se0) [299].
Since AsO43− is less toxic than AsO33−, AsO33−-oxidizing microorganisms can also be used for bioremediation purposes. In a study conducted by Tang et al. (2013), the ability of a heterotrophic bacterial consortium in simultaneous phenanthrene biodegradation and AsO33− oxidation was investigated. This consortium can remediate 71.4% of phenanthrene (200 mg/L) and oxidize 96.2% of AsO33− (60 mg/L) after 48 h incubation simultaneously. Dominant genera of this consortium include Pseudomonas, Pusillimonas, Alcaligenes, and Achromobacter. The shift in this consortium’s community structure in the presence of contaminates is listed in Table 4 [300].
Mix-culture of sulfate-reducing bacteria (SRB) was isolated from an Sb mine in Guangxi, China, and applied to eradicate antimony and arsenic co-contaminants from wastewater. In this study, Liu et al. (2018) used ethanol and lactate as carbon sources and electron donors. The amount of arsenic and antimony removal increased in the presence of ethanol rather than lactate, indicating that ethanol is a better carbon source than lactate. After 12 days, this mixed culture can reduce 97.8 and 26.4% of 5 mg/L of both antimony and arsenic in the presence of ethanol, respectively. The addition of Fe (II) can elevate these amounts to 99.4% (Antimony) and 98.2% (arsenic) through co-precipitation [257].
Zhang et al. (2016) applied a mixed SRB culture for antimony bioremediation of a stream contaminated by coalmine effluents. This mixed culture could remediate 93% of Sb (V) (5 mg/L) over 11 days of incubation. Due to bioreduction, insoluble antimony sulfide (Sb2S3) was produced and detected microscopically. Further studies showed that microbial biosorption had an insignificant role in the antimony bioreduction of this SRB mix culture [301].
By producing EPS, many microorganisms can withstand environmental stress and the invasion of bacteriophages, and the effects of certain toxins, creating anaerobic chambers and separating aerobic and anaerobic microenvironments. This is especially important in bioremediation, where some microorganisms grow under anaerobic conditions and use oxyanions as the last electron acceptor in respiratory reactions. Besides, EPS can be used as a carbon source by microorganisms. It can also act as a barrier against the loss of valuable compounds around microorganisms. Since metalloid bioremediation is based on redox reactions, the presence of structures such as nanowires that form between bacteria and archaea and are involved in electron transfer should also be investigated in studies [302,303,304].

6. Conclusions

Despite the considerable research to date that has been completed on the bioremediation of metalloids so far, we still have poor knowledge about the underlying mechanisms and requirements of the microorganisms involved in this process. Proteomics and metabolomics studies can be perfect methods for illustrating the involved mechanisms of bioremediation. Furthermore, some software can be designed for predicting the suitable amount of nutrients, enzyme cofactors, electron shuttles, type of microorganisms, and physiochemical factors that are required for efficient ex-situ bioremediation of polluted sites. The relationship between sulfur and nitrogen cycles with the detoxification of metalloids and the role of microorganisms involved in these two cycles in the bioremediation of contaminated areas are fascinating and more investigations are needed in this field of study. Moreover, proteomic and transcriptomic studies can help reveal the synergistic relationships of microorganisms that reduce metalloids and decrease their toxic effects.

Author Contributions

Conceptualization, E.L., H.M., R.J.T. and M.A.A.; investigation, E.L. and H.M.; writing—original draft preparation, E.L.; writing—review and editing, E.L., H.M., R.J.T. and M.A.A.; visualization, E.L.; supervision, H.M., R.J.T. and M.A.A.; project administration, H.M. and M.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

Funding was provided to RJT from the Natural Sciences and Engineering Research Council of Canada.

Data Availability Statement

Data is contained within the article.

Acknowledgments

Assistance provided by Kimia Kermanshahian was greatly appreciated.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miazek, K.; Iwanek, W.; Remacle, C.; Richel, A.; Goffin, D. Effect of metals, metalloids and metallic nanoparticles on microalgae growth and industrial product biosynthesis: A review. Int. J. Mol. Sci. 2015, 16, 23929–23969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sekhon, B. Metalloid compounds as drugs. Res. Pharm. Sci. 2013, 8, 145. [Google Scholar] [PubMed]
  3. Zambonino, M.C.; Quizhpe, E.M.; Jaramillo, F.E.; Rahman, A.; Santiago Vispo, N.; Jeffryes, C.; Dahoumane, S.A. Green synthesis of selenium and tellurium nanoparticles: Current trends, biological properties and biomedical applications. Int. J. Mol. Sci. 2021, 22, 989. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, S.; Xing, C.; Huang, D.; Zhou, C.; Ding, B.; Guo, Z.; Peng, Z.; Wang, D.; Zhu, X.; Liu, S. Eradication of tumor growth by delivering novel photothermal selenium-coated tellurium nanoheterojunctions. Sci. Adv. 2020, 6, eaay6825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Zheng, D.; Guo, Y.; Zhang, M.; Feng, X.; Zhu, L.; Qiu, L.; Jin, X.; Zhao, G. Anisotropic charge carrier transport of optoelectronic functional selenium-containing organic semiconductor materials. J. Comput. Chem. 2020, 41, 976–985. [Google Scholar] [CrossRef]
  6. Wu, Y.; Tong, J.; Ruan, L.; Luo, F.; Liu, G.; Zhang, R.; Han, X.; Zhang, Y.; Tian, F.; Zhang, X. N-type diamond semiconductor induced by co-doping selenium and boron. Comput. Mater. Sci. 2021, 196, 110515. [Google Scholar] [CrossRef]
  7. Huang, W.; Wang, M.; Hu, L.; Wang, C.; Xie, Z.; Zhang, H. Recent advances in semiconducting monoelemental selenium nanostructures for device applications. Adv. Funct. Mater. 2020, 30, 2003301. [Google Scholar] [CrossRef]
  8. Mehdi, Y.; Hornick, J.-L.; Istasse, L.; Dufrasne, I. Selenium in the environment, metabolism and involvement in body functions. Molecules 2013, 18, 3292–3311. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, X.-H.; Adam, J.-L.; Bureau, B. Chalcogenide glasses. In Springer Handbook of Glass; Springer: Cham, Switzerland, 2019; pp. 525–552. [Google Scholar]
  10. Månberger, A.; Johansson, B. The geopolitics of metals and metalloids used for the renewable energy transition. Energy Strategy Rev. 2019, 26, 100394. [Google Scholar] [CrossRef]
  11. Babula, P.; Adam, V.; Opatrilova, R.; Zehnalek, J.; Havel, L.; Kizek, R. Uncommon heavy metals, metalloids and their plant toxicity: A review. In Organic Farming, Pest Control and Remediation of Soil Pollutants; Springer: Berlin/Heidelberg, Germany, 2009; pp. 275–317. [Google Scholar]
  12. Veerapandian, M.; Kim, H.Y.; Seo, Y.-T.; Lee, K.-N.; Yun, K.; Lee, M.-H. Metalloid polymer nanoparticle functionalized graphene oxide working electrode for durable glucose sensing. Mater. Res. Bull. 2014, 49, 593–600. [Google Scholar] [CrossRef]
  13. Feinberg, A.; Stenke, A.; Peter, T.; Hinckley, E.-L.S.; Driscoll, C.T.; Winkel, L.H. Reductions in the deposition of sulfur and selenium to agricultural soils pose risk of future nutrient deficiencies. Commun. Earth Environ. 2021, 2, 101. [Google Scholar] [CrossRef]
  14. Kopp, B.; Zalko, D.; Audebert, M. Genotoxicity of 11 heavy metals detected as food contaminants in two human cell lines. Environ. Mol. Mutagen. 2018, 59, 202–210. [Google Scholar] [CrossRef]
  15. Aguilar, N.C.; Faria, M.C.; Pedron, T.; Batista, B.L.; Mesquita, J.P.; Bomfeti, C.A.; Rodrigues, J.L. Isolation and characterization of bacteria from a brazilian gold mining area with a capacity of arsenic bioaccumulation. Chemosphere 2020, 240, 124871. [Google Scholar] [CrossRef]
  16. Khamkhash, A.; Srivastava, V.; Ghosh, T.; Akdogan, G.; Ganguli, R.; Aggarwal, S. Mining-related selenium contamination in Alaska, and the state of current knowledge. Minerals 2017, 7, 46. [Google Scholar] [CrossRef]
  17. Kavlak, G.; Graedel, T. Global anthropogenic selenium cycles for 1940–2010. Resour. Conserv. Recycl. 2013, 73, 17–22. [Google Scholar] [CrossRef]
  18. Staicu, L.C.; Oremland, R.S.; Tobe, R.; Mihara, H. Bacteria versus selenium: A view from the inside out. In Selenium in Plants; Springer: Cham, Switzerland, 2017; pp. 79–108. [Google Scholar]
  19. Lemly, A.D. Aquatic selenium pollution is a global environmental safety issue. Ecotoxicol. Environ. Saf. 2004, 59, 44–56. [Google Scholar] [CrossRef] [Green Version]
  20. Gerson, A.R.; Fan, R.; Qian, G.; Schumann, R.C.; Olin, P.; Howard, D.L.; Smart, R.S.C. Examination of multiple sources of selenium release from coal wastes and strategies for remediation. J. Hazard. Mater. 2022, 422, 126924. [Google Scholar] [CrossRef]
  21. Hao, L.; Zhang, J.; Zhang, S.; Ma, S.; Li, B.; Long, J.; Fan, J.; Luo, K. Distribution characteristics and main influencing factors of selenium in surface soil of natural selenium-rich area: A case study in Langao County, China. Environ. Geochem. Health 2021, 43, 333–346. [Google Scholar] [CrossRef]
  22. Ibrahim, A.S.; Al-Farawati, R. Selenium Concentration Levels in Two Polluted Lagoons, Eastern Red Sea Coastal Waters. Water 2023, 15, 687. [Google Scholar] [CrossRef]
  23. Rayman, M.P. The importance of selenium to human health. Lancet 2000, 356, 233–241. [Google Scholar] [CrossRef] [Green Version]
  24. Piacenza, E.; Presentato, A.; Ambrosi, E.; Speghini, A.; Turner, R.J.; Vallini, G.; Lampis, S. Physical–chemical properties of biogenic selenium nanostructures produced by Stenotrophomonas maltophilia SeITE02 and Ochrobactrum sp. MPV1. Front. Microbiol. 2018, 9, 3178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Presentato, A.; Piacenza, E.; Anikovskiy, M.; Cappelletti, M.; Zannoni, D.; Turner, R.J. Biosynthesis of selenium-nanoparticles and-nanorods as a product of selenite bioconversion by the aerobic bacterium Rhodococcus aetherivorans BCP1. New Biotechnol. 2018, 41, 1–8. [Google Scholar] [CrossRef] [PubMed]
  26. Stoytcheva, Z.R.; Berry, M.J. Transcriptional regulation of mammalian selenoprotein expression. Biochim. Et Biophys. Acta (BBA)-Gen. Subj. 2009, 1790, 1429–1440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Gladyshev, V.N.; Arnér, E.S.; Berry, M.J.; Brigelius-Flohé, R.; Bruford, E.A.; Burk, R.F.; Carlson, B.A.; Castellano, S.; Chavatte, L.; Conrad, M. Selenoprotein gene nomenclature. J. Biol. Chem. 2016, 291, 24036–24040. [Google Scholar] [CrossRef] [Green Version]
  28. Liu, X.; Wang, Y.; Han, S.; Zhang, Y.; Zou, Y.; Su, S.; Zhou, H.; Zhang, X.; Liang, H.; Hou, J. A spatial ecological study on serum selenium and Keshan disease in Heilongjiang province, China. Biol. Trace Elem. Res. 2020, 199, 3253–3261. [Google Scholar] [CrossRef]
  29. Hou, J.; Zhu, L.; Chen, C.; Feng, H.; Li, D.; Sun, S.; Xing, Z.; Wan, X.; Wang, X.; Li, F. Association of selenium levels with the prevention and control of Keshan disease: A cross-sectional study. J. Trace Elem. Med. Biol. 2021, 68, 126832. [Google Scholar] [CrossRef]
  30. Zhu, Y.-h.; Wang, X.-f.; Yang, G.; Wei, J.; Tan, W.-h.; Wang, L.-x.; Guo, X.; Lammi, M.J.; Xu, J.-h. Efficacy of long-term selenium supplementation in the treatment of chronic Keshan disease with congestive heart failure. Curr. Med. Sci. 2019, 39, 237–242. [Google Scholar] [CrossRef]
  31. Jia, Y.; Wang, R.; Su, S.; Qi, L.; Wang, Y.; Wang, Y.; Zou, Y.; Liu, X.; Zhang, Y.; Hou, J. A county-level spatial study of serum selenoprotein P and Keshan disease. Front. Nutr. 2022, 9, 47. [Google Scholar] [CrossRef]
  32. Rannem, T.; Ladefoged, K.; Hylander, E.; Christiansen, J.; Laursen, H.; Kristensen, J.H.; Linstow, M.; Beyer, N.; Liguori, R.; Dige-Petersen, H. The effect of selenium supplementation on skeletal and cardiac muscle in selenium-depleted patients. J. Parenter. Enter. Nutr. 1995, 19, 351–355. [Google Scholar] [CrossRef]
  33. Peng, A.; Yang, C.; Rui, H.; Li, H. Study on the pathogenic factors of Kashin-Beck disease. J. Toxicol. Environ. Health Part A Curr. Issues 1992, 35, 79–90. [Google Scholar] [CrossRef]
  34. Zhang, D.; Zhang, D.; Yang, X.; Li, Q.; Zhang, R.; Xiong, Y. The Role of Selenium-Mediated Notch/Hes1 Signaling Pathway in Kashin–Beck Disease Patients and Cartilage Injury Models. Biol. Trace Elem. Res. 2022. [Google Scholar] [CrossRef]
  35. Wang, J.; Zhao, S.; Yang, L.; Gong, H.; Li, H.; Nima, C. Assessing the health loss from Kashin-Beck disease and its relationship with environmental selenium in Qamdo district of Tibet, China. Int. J. Environ. Res. Public Health 2021, 18, 11. [Google Scholar] [CrossRef]
  36. Wang, L.; Yin, J.; Yang, B.; Qu, C.; Lei, J.; Han, J.; Guo, X. Serious selenium deficiency in the serum of patients with Kashin–Beck disease and the effect of nano-selenium on their chondrocytes. Biol. Trace Elem. Res. 2020, 194, 96–104. [Google Scholar] [CrossRef]
  37. Ning, Y.; Hu, M.; Chen, S.; Zhang, F.; Yang, X.; Zhang, Q.; Gong, Y.; Huang, R.; Liu, Y.; Chen, F. Investigation of selenium nutritional status and dietary pattern among children in Kashin-Beck disease endemic areas in Shaanxi Province, China using duplicate portion sampling method. Environ. Int. 2022, 164, 107255. [Google Scholar] [CrossRef]
  38. Saito, Y.; Hashimoto, T.; Sasaki, M.; Hanaoka, S.; Sugai, K. Effect of selenium deficiency on cardiac function of individuals with severe disabilities under long-term tube feeding. Dev. Med. Child Neurol. 1998, 40, 743–748. [Google Scholar] [CrossRef] [Green Version]
  39. Bomer, N.; Grote Beverborg, N.; Hoes, M.F.; Streng, K.W.; Vermeer, M.; Dokter, M.M.; IJmker, J.; Anker, S.D.; Cleland, J.G.; Hillege, H.L. Selenium and outcome in heart failure. Eur. J. Heart Fail. 2020, 22, 1415–1423. [Google Scholar] [CrossRef] [Green Version]
  40. Shimada, B.K.; Alfulaij, N.; Seale, L.A. The impact of selenium deficiency on cardiovascular function. Int. J. Mol. Sci. 2021, 22, 10713. [Google Scholar] [CrossRef]
  41. Mirdamadi, A.; Rafiei, R.; Kahazaipour, G.; Fouladi, L. Selenium level in patients with heart failure versus normal individuals. Int. J. Prev. Med. 2019, 10, 210. [Google Scholar]
  42. Tang, C.; Li, S.; Zhang, K.; Li, J.; Han, Y.; Zhan, T.; Zhao, Q.; Guo, X.; Zhang, J. Selenium deficiency-induced redox imbalance leads to metabolic reprogramming and inflammation in the liver. Redox Biol. 2020, 36, 101519. [Google Scholar] [CrossRef]
  43. Lenz, M.; Lens, P.N. The essential toxin: The changing perception of selenium in environmental sciences. Sci. Total Environ. 2009, 407, 3620–3633. [Google Scholar] [CrossRef]
  44. Long, Z.; Xiang, J.; Song, J.; Lu, Y.; Yin, H.; Zhu, Y.; Liu, X.; Qin, L.; Bañuelos, G.S.; Wang, Z. Soil selenium concentration and residents daily dietary intake in a selenosis area: A preliminary study in Yutangba Village, Enshi City, China. Bull. Environ. Contam. Toxicol. 2020, 105, 798–805. [Google Scholar] [CrossRef] [PubMed]
  45. Yuan, L.; Yin, X.; Zhu, Y.; Li, F.; Huang, Y.; Liu, Y.; Lin, Z. Selenium in plants and soils, and selenosis in Enshi, China: Implications for selenium biofortification. In Phytoremediation and Biofortification; Springer: Dordrecht, The Netherlands, 2012; pp. 7–31. [Google Scholar]
  46. Gheorghiu, M.L.; Badiu, C. Selenium involvement in mitochondrial function in thyroid disorders. Hormones 2020, 19, 25–30. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, Z.; Li, S.; Jiang, H.; Liu, B.; Lv, Z.; Guo, C.; Zhang, H. Effects of selenium on apoptosis and abnormal amino acid metabolism induced by excess fatty acid in isolated rat hepatocytes. Mol. Nutr. Food Res. 2017, 61, 1700016. [Google Scholar] [CrossRef]
  48. Chawla, R.; Filippini, T.; Loomba, R.; Cilloni, S.; Dhillon, K.S.; Vinceti, M. Exposure to a high selenium environment in Punjab, India: Biomarkers and health conditions. Sci. Total Environ. 2020, 719, 134541. [Google Scholar] [CrossRef]
  49. Hasan, M.A.; Ahmed, K.M.; Sracek, O.; Bhattacharya, P.; Von Broemssen, M.; Broms, S.; Fogelström, J.; Mazumder, M.L.; Jacks, G. Arsenic in shallow groundwater of Bangladesh: Investigations from three different physiographic settings. Hydrogeol. J. 2007, 15, 1507–1522. [Google Scholar] [CrossRef]
  50. Saha, G.C.; Ali, M.A. Dynamics of arsenic in agricultural soils irrigated with arsenic contaminated groundwater in Bangladesh. Sci. Total Environ. 2007, 379, 180–189. [Google Scholar] [CrossRef]
  51. Rahman, M.M.; Dong, Z.; Naidu, R. Concentrations of arsenic and other elements in groundwater of Bangladesh and West Bengal, India: Potential cancer risk. Chemosphere 2015, 139, 54–64. [Google Scholar] [CrossRef]
  52. Hasan, M.K.; Shahriar, A.; Jim, K.U. Water pollution in Bangladesh and its impact on public health. Heliyon 2019, 5, e02145. [Google Scholar] [CrossRef] [Green Version]
  53. Chakraborti, D.; Rahman, M.M.; Ahamed, S.; Dutta, R.N.; Pati, S.; Mukherjee, S.C. Arsenic groundwater contamination and its health effects in Patna district (capital of Bihar) in the middle Ganga plain, India. Chemosphere 2016, 152, 520–529. [Google Scholar] [CrossRef]
  54. Li, Y.; Bi, Y.; Mi, W.; Xie, S.; Ji, L. Land-use change caused by anthropogenic activities increase fluoride and arsenic pollution in groundwater and human health risk. J. Hazard. Mater. 2021, 406, 124337. [Google Scholar] [CrossRef]
  55. Nath, A.; Samanta, S.; Banerjee, S.; Danda, A.A.; Hazra, S. Threat of arsenic contamination, salinity and water pollution in agricultural practices of Sundarban Delta, India, and mitigation strategies. SN Appl. Sci. 2021, 3, 1–15. [Google Scholar] [CrossRef]
  56. Rahman, A.; Mondal, N.; Fauzia, F. Arsenic enrichment and its natural background in groundwater at the proximity of active floodplains of Ganga River, northern India. Chemosphere 2021, 265, 129096. [Google Scholar] [CrossRef]
  57. Alsubih, M.; El Morabet, R.; Khan, R.A.; Khan, N.A.; ul Haq Khan, M.; Ahmed, S.; Qadir, A.; Changani, F. Occurrence and health risk assessment of arsenic and heavy metals in groundwater of three industrial areas in Delhi, India. Environ. Sci. Pollut. Res. 2021, 28, 63017–63031. [Google Scholar] [CrossRef]
  58. Wu, R.; Podgorski, J.; Berg, M.; Polya, D.A. Geostatistical model of the spatial distribution of arsenic in groundwaters in Gujarat State, India. Environ. Geochem. Health 2021, 43, 2649–2664. [Google Scholar] [CrossRef]
  59. Tong, J.; Guo, H.; Wei, C. Arsenic contamination of the soil–wheat system irrigated with high arsenic groundwater in the Hetao Basin, Inner Mongolia, China. Sci. Total Environ. 2014, 496, 479–487. [Google Scholar] [CrossRef]
  60. Zhang, Y.; Xu, B.; Guo, Z.; Han, J.; Li, H.; Jin, L.; Chen, F.; Xiong, Y. Human health risk assessment of groundwater arsenic contamination in Jinghui irrigation district, China. J. Environ. Manag. 2019, 237, 163–169. [Google Scholar] [CrossRef]
  61. Han, L.; Gao, B.; Hao, H.; Lu, J.; Xu, D. Arsenic pollution of sediments in China: An assessment by geochemical baseline. Sci. Total Environ. 2019, 651, 1983–1991. [Google Scholar] [CrossRef]
  62. Gong, Y.; Qu, Y.; Yang, S.; Tao, S.; Shi, T.; Liu, Q.; Chen, Y.; Wu, Y.; Ma, J. Status of arsenic accumulation in agricultural soils across China (1985–2016). Environ. Res. 2020, 186, 109525. [Google Scholar] [CrossRef]
  63. Ran, H.; Guo, Z.; Yi, L.; Xiao, X.; Zhang, L.; Hu, Z.; Li, C.; Zhang, Y. Pollution characteristics and source identification of soil metal (loid) s at an abandoned arsenic-containing mine, China. J. Hazard. Mater. 2021, 413, 125382. [Google Scholar] [CrossRef]
  64. Li, C. Arsenic pollution in shallow drinking wells in Yuncheng Basin, China: Occurrence and mechanisms. In Environmental Arsenic in a Changing World; CRC Press: Boca Raton, FL, USA, 2019; pp. 613–614. [Google Scholar]
  65. Kanel, S.R.; Malla, G.B.; Choi, H. Modeling and study of the mechanism of mobilization of arsenic contamination in the groundwater of Nepal in South Asia. Clean Technol. Environ. Policy 2013, 15, 1077–1082. [Google Scholar] [CrossRef]
  66. Mueller, B.; Chan, M.C.; Hug, S.J. Unique Geochemistry of Arsenic-Contaminated Groundwater and Corresponding Mitigation Efforts in Southern Nepal. In ACS EST Water; ACS Publications: Washington, DC, USA, 2023. [Google Scholar]
  67. Phan, K.; Sthiannopkao, S.; Kim, K.-W.; Wong, M.H.; Sao, V.; Hashim, J.H.; Yasin, M.S.M.; Aljunid, S.M. Health risk assessment of inorganic arsenic intake of Cambodia residents through groundwater drinking pathway. Water Res. 2010, 44, 5777–5788. [Google Scholar] [CrossRef] [PubMed]
  68. Postma, D.; Larsen, F.; Thai, N.T.; Trang, P.T.K.; Jakobsen, R.; Nhan, P.Q.; Long, T.V.; Viet, P.H.; Murray, A.S. Groundwater arsenic concentrations in Vietnam controlled by sediment age. Nat. Geosci. 2012, 5, 656–661. [Google Scholar] [CrossRef]
  69. Van Geen, A.; Win, K.H.; Zaw, T.; Naing, W.; Mey, J.L.; Mailloux, B. Confirmation of elevated arsenic levels in groundwater of Myanmar. Sci. Total Environ. 2014, 478, 21–24. [Google Scholar] [CrossRef] [Green Version]
  70. Shahid, M.; Niazi, N.K.; Dumat, C.; Naidu, R.; Khalid, S.; Rahman, M.M.; Bibi, I. A meta-analysis of the distribution, sources and health risks of arsenic-contaminated groundwater in Pakistan. Environ. Pollut. 2018, 242, 307–319. [Google Scholar] [CrossRef] [PubMed]
  71. Tabassum, R.A.; Shahid, M.; Dumat, C.; Niazi, N.K.; Khalid, S.; Shah, N.S.; Imran, M.; Khalid, S. Health risk assessment of drinking arsenic-containing groundwater in Hasilpur, Pakistan: Effect of sampling area, depth, and source. Environ. Sci. Pollut. Res. 2019, 26, 20018–20029. [Google Scholar] [CrossRef]
  72. Ur Rehman, H.; Ahmed, S.; Ur Rahman, M.; Mehmood, M.S. Arsenic contamination, induced symptoms, and health risk assessment in groundwater of Lahore, Pakistan. Environ. Sci. Pollut. Res. 2022, 29, 49796–49807. [Google Scholar] [CrossRef]
  73. Jat Baloch, M.Y.; Zhang, W.; Zhang, D.; Al Shoumik, B.A.; Iqbal, J.; Li, S.; Chai, J.; Farooq, M.A.; Parkash, A. Evolution Mechanism of Arsenic Enrichment in Groundwater and Associated Health Risks in Southern Punjab, Pakistan. Int. J. Environ. Res. Public Health 2022, 19, 13325. [Google Scholar] [CrossRef]
  74. Masood, N.; Farooqi, A.; Zafar, M.I. Health risk assessment of arsenic and other potentially toxic elements in drinking water from an industrial zone of Gujrat, Pakistan: A case study. Environ. Monit. Assess. 2019, 191, 1–15. [Google Scholar] [CrossRef]
  75. Ashraf, A.; Chen, X.; Ramamurthy, R. Modelling heavy metals contamination in groundwater of Southern Punjab, Pakistan. Int. J. Environ. Sci. Technol. 2021, 18, 2221–2236. [Google Scholar] [CrossRef]
  76. Winkel, L.; Berg, M.; Stengel, C.; Rosenberg, T. Hydrogeological survey assessing arsenic and other groundwater contaminants in the lowlands of Sumatra, Indonesia. Appl. Geochem. 2008, 23, 3019–3028. [Google Scholar] [CrossRef]
  77. Barringer, J.L.; Mumford, A.; Young, L.Y.; Reilly, P.A.; Bonin, J.L.; Rosman, R. Pathways for arsenic from sediments to groundwater to streams: Biogeochemical processes in the Inner Coastal Plain, New Jersey, USA. Water Res. 2010, 44, 5532–5544. [Google Scholar] [CrossRef]
  78. Sorg, T.J.; Chen, A.S.; Wang, L. Arsenic species in drinking water wells in the USA with high arsenic concentrations. Water Res. 2014, 48, 156–169. [Google Scholar] [CrossRef]
  79. Moncur, M.C.; Paktunc, D.; Birks, S.J.; Ptacek, C.J.; Welsh, B.; Thibault, Y. Source and distribution of naturally occurring arsenic in groundwater from Alberta’s Southern Oil Sands Regions. Appl. Geochem. 2015, 62, 171–185. [Google Scholar] [CrossRef] [Green Version]
  80. Bailey, A.S.; Jamieson, H.E.; Radková, A.B. Geochemical characterization of dust from arsenic-bearing tailings, Giant Mine, Canada. Appl. Geochem. 2021, 135, 105119. [Google Scholar] [CrossRef]
  81. Francisca, F.M.; Perez, M.E.C. Assessment of natural arsenic in groundwater in Cordoba Province, Argentina. Environ. Geochem. Health 2009, 31, 673. [Google Scholar] [CrossRef]
  82. Leybourne, M.I.; Cameron, E.M. Source, transport, and fate of rhenium, selenium, molybdenum, arsenic, and copper in groundwater associated with porphyry–Cu deposits, Atacama Desert, Chile. Chem. Geol. 2008, 247, 208–228. [Google Scholar] [CrossRef]
  83. Parga, J.R.; Cocke, D.L.; Valenzuela, J.L.; Gomes, J.A.; Kesmez, M.; Irwin, G.; Moreno, H.; Weir, M. Arsenic removal via electrocoagulation from heavy metal contaminated groundwater in La Comarca Lagunera Mexico. J. Hazard. Mater. 2005, 124, 247–254. [Google Scholar] [CrossRef]
  84. Rowland, H.A.; Omoregie, E.O.; Millot, R.; Jimenez, C.; Mertens, J.; Baciu, C.; Hug, S.J.; Berg, M. Geochemistry and arsenic behaviour in groundwater resources of the Pannonian Basin (Hungary and Romania). Appl. Geochem. 2011, 26, 1–17. [Google Scholar] [CrossRef] [Green Version]
  85. Mudzielwana, R.; Gitari, M.W.; Akinyemi, S.A.; Talabi, A.O.; Ndungu, P. Hydrogeochemical characteristics of arsenic rich groundwater in Greater Giyani Municipality, Limpopo Province, South Africa. Groundw. Sustain. Dev. 2020, 10, 100336. [Google Scholar] [CrossRef]
  86. Shaji, E.; Santosh, M.; Sarath, K.; Prakash, P.; Deepchand, V.; Divya, B. Arsenic contamination of groundwater: A global synopsis with focus on the Indian Peninsula. Geosci. Front. 2020, 12, 101079. [Google Scholar] [CrossRef]
  87. Lampis, S.; Zonaro, E.; Bertolini, C.; Cecconi, D.; Monti, F.; Micaroni, M.; Turner, R.J.; Butler, C.S.; Vallini, G. Selenite biotransformation and detoxification by Stenotrophomonas maltophilia SeITE02: Novel clues on the route to bacterial biogenesis of selenium nanoparticles. J. Hazard. Mater. 2017, 324, 3–14. [Google Scholar] [CrossRef] [PubMed]
  88. Tsai, S.-L.; Singh, S.; Chen, W. Arsenic metabolism by microbes in nature and the impact on arsenic remediation. Curr. Opin. Biotechnol. 2009, 20, 659–667. [Google Scholar] [CrossRef] [PubMed]
  89. Churbanov, M.; Snopatin, G.; Shiryaev, V.; Plotnichenko, V.; Dianov, E. Recent advances in preparation of high-purity glasses based on arsenic chalcogenides for fiber optics. J. Non-Cryst. Solids 2011, 357, 2352–2357. [Google Scholar] [CrossRef]
  90. Das, A.; Shamirian, A.; Snee, P.T. Arsenic silylamide: An effective precursor for arsenide semiconductor nanocrystal synthesis. Chem. Mater. 2016, 28, 4058–4064. [Google Scholar] [CrossRef]
  91. Lechtman, H.; Klein, S. The production of copper–arsenic alloys (arsenic bronze) by cosmelting: Modern experiment, ancient practice. J. Archaeol. Sci. 1999, 26, 497–526. [Google Scholar] [CrossRef] [Green Version]
  92. Fitzmaurice, A.G.; Bilgin, A.A.; O’Day, P.A.; Illera, V.; Burris, D.R.; Reisinger, H.J.; Hering, J.G. Geochemical and hydrologic controls on the mobilization of arsenic derived from herbicide application. Appl. Geochem. 2009, 24, 2152–2162. [Google Scholar] [CrossRef] [Green Version]
  93. Qi, Y.; Donahoe, R.J. The environmental fate of arsenic in surface soil contaminated by historical herbicide application. Sci. Total Environ. 2008, 405, 246–254. [Google Scholar] [CrossRef]
  94. Kazi, F.K.M.; Cooper, P.A. Method to recover and reuse chromated copper arsenate wood preservative from spent treated wood. Waste Manag. 2006, 26, 182–188. [Google Scholar] [CrossRef]
  95. Peryea, F.; Creger, T. Vertical distribution of lead and arsenic in soils contaminated with lead arsenate pesticide residues. Water Air Soil Pollut. 1994, 78, 297–306. [Google Scholar] [CrossRef]
  96. Hu, Y.; Cheng, H.; Tao, S.; Schnoor, J.L. China’s ban on phenylarsonic feed additives, a major step toward reducing the human and ecosystem health risk from arsenic. Environ. Sci. Technol. 2019, 53, 12177–12187. [Google Scholar] [CrossRef]
  97. Douer, D.; Tallman, M.S. Arsenic trioxide: New clinical experience with an old medication in hematologic malignancies. J. Clin. Oncol. 2005, 23, 2396–2410. [Google Scholar] [CrossRef]
  98. Liu, J.T.; Bain, L.J. Arsenic inhibits hedgehog signaling during P19 cell differentiation. Toxicol. Appl. Pharmacol. 2014, 281, 243–253. [Google Scholar] [CrossRef] [Green Version]
  99. Prakash, C.; Soni, M.; Kumar, V. Biochemical and molecular alterations following arsenic-induced oxidative stress and mitochondrial dysfunction in rat brain. Biol. Trace Elem. Res. 2015, 167, 121–129. [Google Scholar] [CrossRef]
  100. Kitchin, K.T.; Wallace, K. Arsenite binding to synthetic peptides: The effect of increasing length between two cysteines. J. Biochem. Mol. Toxicol. 2006, 20, 35–38. [Google Scholar] [CrossRef]
  101. Rosenblatt, A.E.; Burnstein, K.L. Inhibition of androgen receptor transcriptional activity as a novel mechanism of action of arsenic. Mol. Endocrinol. 2009, 23, 412–421. [Google Scholar] [CrossRef] [Green Version]
  102. Ter Welle, H.; Slater, E. Uncoupling of respiratory-chain phosphorylation by arsenate and evidence for the existence of a stable X? Pintermediate of oxidative phosphorylation. Biochim. Et Biophys. Acta. Enzymol. 1964, 89, 385–388. [Google Scholar]
  103. Yu, R.C.; Hsu, K.-H.; Chen, C.-J.; Froines, J.R. Arsenic methylation capacity and skin cancer. Cancer Epidemiol. Prev. Biomark. 2000, 9, 1259–1262. [Google Scholar]
  104. Ferreccio, C.; González, C.; Milosavjlevic, V.; Marshall, G.; Sancha, A.M.; Smith, A.H. Lung cancer and arsenic concentrations in drinking water in Chile. Epidemiology 2000, 673–679. [Google Scholar] [CrossRef]
  105. Lee, A.M.; Fraumeni Jr, J.F. Arsenic and respiratory cancer in man: An occupational study. J. Natl. Cancer Inst. 1969, 42, 1045–1052. [Google Scholar]
  106. Bates, M.N.; Rey, O.A.; Biggs, M.L.; Hopenhayn, C.; Moore, L.E.; Kalman, D.; Steinmaus, C.; Smith, A.H. Case-control study of bladder cancer and exposure to arsenic in Argentina. Am. J. Epidemiol. 2004, 159, 381–389. [Google Scholar] [CrossRef] [Green Version]
  107. Wang, W.; Cheng, S.; Zhang, D. Association of inorganic arsenic exposure with liver cancer mortality: A meta-analysis. Environ. Res. 2014, 135, 120–125. [Google Scholar] [CrossRef] [PubMed]
  108. Ahsan, H.; Chen, Y.; Parvez, F.; Zablotska, L.; Argos, M.; Hussain, I.; Momotaj, H.; Levy, D.; Cheng, Z.; Slavkovich, V. Arsenic exposure from drinking water and risk of premalignant skin lesions in Bangladesh: Baseline results from the Health Effects of Arsenic Longitudinal Study. Am. J. Epidemiol. 2006, 163, 1138–1148. [Google Scholar] [CrossRef] [PubMed]
  109. Halatek, T.; Sinczuk-Walczak, H.; Rabieh, S.; Wasowicz, W. Association between occupational exposure to arsenic and neurological, respiratory and renal effects. Toxicol. Appl. Pharmacol. 2009, 239, 193–199. [Google Scholar] [CrossRef] [PubMed]
  110. Mazumder, D.G. Effect of chronic intake of arsenic-contaminated water on liver. Toxicol. Appl. Pharmacol. 2005, 206, 169–175. [Google Scholar] [CrossRef]
  111. Islam, L.N.; Nabi, A.; Rahman, M.M.; Khan, M.A.; Kazi, A.I. Association of clinical complications with nutritional status and the prevalence of leukopenia among arsenic patients in Bangladesh. Int. J. Environ. Res. Public Health 2004, 1, 74–82. [Google Scholar] [CrossRef] [Green Version]
  112. Chen, Y.; Graziano, J.H.; Parvez, F.; Liu, M.; Slavkovich, V.; Kalra, T.; Argos, M.; Islam, T.; Ahmed, A.; Rakibuz-Zaman, M. Arsenic exposure from drinking water and mortality from cardiovascular disease in Bangladesh: Prospective cohort study. Bmj 2011, 342, d2431. [Google Scholar] [CrossRef] [Green Version]
  113. Chen, Y.; Wu, F.; Liu, M.; Parvez, F.; Slavkovich, V.; Eunus, M.; Ahmed, A.; Argos, M.; Islam, T.; Rakibuz-Zaman, M. A prospective study of arsenic exposure, arsenic methylation capacity, and risk of cardiovascular disease in Bangladesh. Environ. Health Perspect. 2013, 121, 832–838. [Google Scholar] [CrossRef] [Green Version]
  114. Heck, J.E.; Chen, Y.; Grann, V.R.; Slavkovich, V.; Parvez, F.; Ahsan, H. Arsenic exposure and anemia in Bangladesh: A population-based study. J. Occup. Environ. Med. 2008, 50, 80–87. [Google Scholar] [CrossRef]
  115. Hopenhayn, C.; Bush, H.M.; Bingcang, A.; Hertz-Picciotto, I. Association between arsenic exposure from drinking water and anemia during pregnancy. J. Occup. Environ. Med. 2006, 48, 635–643. [Google Scholar] [CrossRef]
  116. Tanmoy, P.; Saha, N.C. Environmental arsenic and selenium contamination and approaches towards its bioremediation through the exploration of microbial adaptations: A review. Pedosphere 2019, 29, 554–568. [Google Scholar]
  117. Moriya, F.; Furumiya, J.; Hashimoto, Y. A case of fatal arsenic poisoning. Forensic Toxicol. 2006, 24, 88–91. [Google Scholar] [CrossRef]
  118. Tanaka, M.; Fujiwara, T. Physiological roles and transport mechanisms of boron: Perspectives from plants. Pflügers Arch. -Eur. J. Physiol. 2008, 456, 671–677. [Google Scholar] [CrossRef]
  119. Uluisik, I.; Karakaya, H.C.; Koc, A. The importance of boron in biological systems. J. Trace Elem. Med. Biol. 2018, 45, 156–162. [Google Scholar] [CrossRef]
  120. Concha, G.; Broberg, K.; Grandér, M.; Cardozo, A.; Palm, B.; Vahter, M. High-level exposure to lithium, boron, cesium, and arsenic via drinking water in the Andes of northern Argentina. Environ. Sci. Technol. 2010, 44, 6875–6880. [Google Scholar] [CrossRef]
  121. Martos-Villa, R.; Mata, M.P.; Williams, L.B.; Nieto, F.; Arroyo Rey, X.; Sainz-Díaz, C.I. Evidence of Hydrocarbon-Rich Fluid Interaction with Clays: Clay Mineralogy and Boron Isotope Data from Gulf of Cádiz Mud Volcano Sediments. Minerals 2020, 10, 651. [Google Scholar] [CrossRef]
  122. Hitit, A.; Yazici, Z.O.; Şahin, H.; Öztürk, P.; Aşgın, A.M.; Hitit, B. A novel Ni-based bulk metallic glass containing high amount of tungsten and boron. J. Alloy. Compd. 2019, 807, 151661. [Google Scholar] [CrossRef]
  123. Vargas-Gonzalez, L.; Speyer, R.F.; Campbell, J. Flexural strength, fracture toughness, and hardness of silicon carbide and boron carbide armor ceramics. Int. J. Appl. Ceram. Technol. 2010, 7, 643–651. [Google Scholar] [CrossRef]
  124. Sarac, N.; Ugur, A.; Boran, R.; Elgin, E.S. The use of boron compounds for stabilization of lipase from Pseudomonas aeruginosa ES3 for the detergent industry. J. Surfactants Deterg. 2015, 18, 275–285. [Google Scholar] [CrossRef]
  125. Da Silva, R.C.; Baird, R.; Degryse, F.; McLaughlin, M.J. Slow and Fast-Release Boron Sources in Potash Fertilizers: Spatial Variability, Nutrient Dissolution and Plant Uptake. Soil Sci. Soc. Am. J. 2018, 82, 1437–1448. [Google Scholar] [CrossRef] [Green Version]
  126. Jin, E.; Li, S.; Ren, M.; Hu, Q.; Gu, Y.; Li, K. Boron affects immune function through modulation of splenic T lymphocyte subsets, cytokine secretion, and lymphocyte proliferation and apoptosis in rats. Biol. Trace Elem. Res. 2017, 178, 261–275. [Google Scholar] [CrossRef]
  127. Routray, I.; Ali, S. Boron induces lymphocyte proliferation and modulates the priming effects of lipopolysaccharide on macrophages. PLoS ONE 2016, 11, e0150607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Hunt, C.; Herbel, J. Effects of dietary boron on calcium and mineral metabolism in the streptozotocin-injected, vitamin D3-deprived rat. Magnes. Trace Elem. 1991, 10, 387–408. [Google Scholar]
  129. Sharma, A.; Mani, V.; Pal, R.P.; Sarkar, S.; Datt, C. Boron supplementation in peripartum Murrah buffaloes: The effect on calcium homeostasis, bone metabolism, endocrine and antioxidant status. J. Trace Elem. Med. Biol. 2020, 62, 126623. [Google Scholar] [CrossRef] [PubMed]
  130. Kabu, M.; Akosman, M.S. Biological effects of boron. In Reviews of Environmental Contamination and Toxicology; Springer: Berlin/Heidelberg, Germany, 2013; pp. 57–75. [Google Scholar]
  131. Donoiu, I.; Militaru, C.; Obleagă, O.; Hunter, J.M.; Neamţu, J.; Biţă, A.; Scorei, I.R.; Rogoveanu, O.C. Effects of boron-containing compounds on cardiovascular disease risk factors–a review. J. Trace Elem. Med. Biol. 2018, 50, 47–56. [Google Scholar] [CrossRef] [PubMed]
  132. Farfán-García, E.; Castillo-Mendieta, N.; Ciprés-Flores, F.; Padilla-Martínez, I.; Trujillo-Ferrara, J.; Soriano-Ursúa, M. Current data regarding the structure-toxicity relationship of boron-containing compounds. Toxicol. Lett. 2016, 258, 115–125. [Google Scholar] [CrossRef]
  133. Barth, R.F.; Vicente, M.H.; Harling, O.K.; Kiger, W.; Riley, K.J.; Binns, P.J.; Wagner, F.M.; Suzuki, M.; Aihara, T.; Kato, I. Current status of boron neutron capture therapy of high grade gliomas and recurrent head and neck cancer. Radiat. Oncol. 2012, 7, 1–21. [Google Scholar] [CrossRef] [Green Version]
  134. Kohno, J.; Kawahata, T.; Otake, T.; Morimoto, M.; Mori, H.; Ueba, N.; Nishio, M.; Kinumaki, A.; Komatsubara, S.; Kawashima, K. Boromycin, an anti-HIV antibiotic. Biosci. Biotechnol. Biochem. 1996, 60, 1036–1037. [Google Scholar] [CrossRef]
  135. Irschik, H.; Schummer, D.; Gerth, K.; Höfle, G.; Reichenbach, H. The tartrolons, new boron-containing antibiotics from a myxobacterium, Sorangium cellulosum. J. Antibiot. 1995, 48, 26–30. [Google Scholar] [CrossRef] [Green Version]
  136. Lee, J.; Dewick, P.; Gorst-Allman, C.; Spreafico, F.; Kowal, C.; Chang, C.; McInnes, A.; Walter, J.; Keller, P.; Floss, H. Further studies on the biosynthesis of the boron-containing antibiotic aplasmomycin. J. Am. Chem. Soc. 1987, 109, 5426–5432. [Google Scholar] [CrossRef]
  137. Harris, W.R.; Amin, S.A.; Küpper, F.C.; Green, D.H.; Carrano, C.J. Borate binding to siderophores: Structure and stability. J. Am. Chem. Soc. 2007, 129, 12263–12271. [Google Scholar] [CrossRef]
  138. Wolkenstein, K.; Gross, J.H.; Falk, H. Boron-containing organic pigments from a Jurassic red alga. Proc. Natl. Acad. Sci. USA 2010, 107, 19374–19378. [Google Scholar] [CrossRef] [Green Version]
  139. USGS. Mineral Commodity Summaries; Nova Science Publishers: Hauppauge, NY, USA, 2011. [Google Scholar]
  140. Zanetti, M.; Camino, G.; Canavese, D.; Morgan, A.B.; Lamelas, F.J.; Wilkie, C.A. Fire retardant halogen− antimony− clay synergism in polypropylene layered silicate nanocomposites. Chem. Mater. 2002, 14, 189–193. [Google Scholar] [CrossRef]
  141. Hu, X.; Guo, X.; He, M.; Li, S. pH-dependent release characteristics of antimony and arsenic from typical antimony-bearing ores. J. Environ. Sci. 2016, 44, 171–179. [Google Scholar] [CrossRef]
  142. Filella, M.; Belzile, N.; Lett, M.-C. Antimony in the environment: A review focused on natural waters. III. Microbiota relevant interactions. Earth-Sci. Rev. 2007, 80, 195–217. [Google Scholar] [CrossRef]
  143. Chen, C.; Fu, K.; Lu, Y.; Zhu, J.; Xue, L.; Hu, Y.; Zhang, X. Use of a tin antimony alloy-filled porous carbon nanofiber composite as an anode in sodium-ion batteries. RSC Adv. 2015, 5, 30793–30800. [Google Scholar] [CrossRef]
  144. Haldar, A.K.; Sen, P.; Roy, S. Use of antimony in the treatment of leishmaniasis: Current status and future directions. Mol. Biol. Int. 2011, 2011, 571242. [Google Scholar] [CrossRef] [Green Version]
  145. Tanoue, A.; Yoo, W.J.; Kobayashi, S. Antimony/N-hydroxyphthalimide as a catalyst system for cross-dehydrogenative coupling reactions under aerobic conditions. Adv. Synth. Catal. 2013, 355, 269–273. [Google Scholar]
  146. Kelepertsis, A.; Alexakis, D.; Skordas, K. Arsenic, antimony and other toxic elements in the drinking water of Eastern Thessaly in Greece and its possible effects on human health. Environ. Geol. 2006, 50, 76–84. [Google Scholar] [CrossRef]
  147. Cavallo, D.; Iavicoli, I.; Setini, A.; Marinaccio, A.; Perniconi, B.; Carelli, G.; Iavicoli, S. Genotoxic risk and oxidative DNA damage in workers exposed to antimony trioxide. Environ. Mol. Mutagen. 2002, 40, 184–189. [Google Scholar] [CrossRef]
  148. An, Y.-J.; Kim, M. Effect of antimony on the microbial growth and the activities of soil enzymes. Chemosphere 2009, 74, 654–659. [Google Scholar] [CrossRef]
  149. Presentato, A.; Piacenza, E.; Anikovskiy, M.; Cappelletti, M.; Zannoni, D.; Turner, R.J. Rhodococcus aetherivorans BCP1 as cell factory for the production of intracellular tellurium nanorods under aerobic conditions. Microb. Cell Factories 2016, 15, 204. [Google Scholar] [CrossRef] [Green Version]
  150. Presentato, A.; Piacenza, E.; Darbandi, A.; Anikovskiy, M.; Cappelletti, M.; Zannoni, D.; Turner, R.J. Assembly, growth and conductive properties of tellurium nanorods produced by Rhodococcus aetherivorans BCP1. Sci. Rep. 2018, 8, 3923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Wedepohl, K.H. The composition of the continental crust. Geochim. Et Cosmochim. Acta 1995, 59, 1217–1232. [Google Scholar] [CrossRef]
  152. Kim, D.-H.; Kim, M.-G.; Jiang, S.; Lee, J.-H.; Hur, H.-G. Promoted reduction of tellurite and formation of extracellular tellurium nanorods by concerted reaction between iron and Shewanella oneidensis MR-1. Environ. Sci. Technol. 2013, 47, 8709–8715. [Google Scholar] [PubMed]
  153. Zare, B.; Faramarzi, M.A.; Sepehrizadeh, Z.; Shakibaie, M.; Rezaie, S.; Shahverdi, A.R. Biosynthesis and recovery of rod-shaped tellurium nanoparticles and their bactericidal activities. Mater. Res. Bull. 2012, 47, 3719–3725. [Google Scholar] [CrossRef]
  154. Zare, B.; Nami, M.; Shahverdi, A.-R. Tracing tellurium and its nanostructures in biology. Biol. Trace Elem. Res. 2017, 180, 171–181. [Google Scholar] [CrossRef] [PubMed]
  155. Ba, L.A.; Döring, M.; Jamier, V.; Jacob, C. Tellurium: An element with great biological potency and potential. Org. Biomol. Chem. 2010, 8, 4203–4216. [Google Scholar] [CrossRef]
  156. Brodsky, M.; Halpert, G.; Albeck, M.; Sredni, B. The anti-inflammatory effects of the tellurium redox modulating compound, AS101, are associated with regulation of NFκB signaling pathway and nitric oxide induction in macrophages. J. Inflamm. 2010, 7, 3. [Google Scholar] [CrossRef] [Green Version]
  157. Halpert, G.; Sredni, B. The effect of the novel tellurium compound AS101 on autoimmune diseases. Autoimmun. Rev. 2014, 13, 1230–1235. [Google Scholar] [CrossRef]
  158. Tavakoli, P.; Ghaffarifar, F.; Delavari, H.; KarimiPourSaryazdi, A.; Dayer, M.S.; Nasiri, V.; Ahmadi, S. Synthesis of Tellurium Oxide (TeO2) Nanorods and Nanoflakes and Evaluation of Its Efficacy Against Leishmania major In Vitro and In Vivo. Acta Parasitol. 2021, 67, 143–152. [Google Scholar] [CrossRef]
  159. Zare, B.; Sepehrizadeh, Z.; Faramarzi, M.A.; Soltany-Rezaee-Rad, M.; Rezaie, S.; Shahverdi, A.R. Antifungal activity of biogenic tellurium nanoparticles against C andida albicans and its effects on squalene monooxygenase gene expression. Biotechnol. Appl. Biochem. 2014, 61, 395–400. [Google Scholar] [CrossRef]
  160. Gerhardsson, L. Tellurium. In Handbook on the Toxicology of Metals; Elsevier: Amsterdam, The Netherlands, 2015; pp. 1217–1228. [Google Scholar]
  161. Marwede, M.; Reller, A. Future recycling flows of tellurium from cadmium telluride photovoltaic waste. Resour. Conserv. Recycl. 2012, 69, 35–49. [Google Scholar] [CrossRef] [Green Version]
  162. Zheng, J.; Yang, L.; Deng, Y.; Zhang, C.; Zhang, Y.; Xiong, S.; Ding, C.; Zhao, J.; Liao, C.; Gong, D. A review of public and environmental consequences of organic germanium. Crit. Rev. Environ. Sci. Technol. 2020, 50, 1384–1409. [Google Scholar] [CrossRef]
  163. Wiche, O.; Székely, B.; Moschner, C.; Heilmeier, H. Germanium in the soil-plant system—A review. Environ. Sci. Pollut. Res. 2018, 25, 31938–31956. [Google Scholar] [CrossRef]
  164. Ji, X.; Page, R.L.; Chaudhuri, S.; Liu, W.; Yu, S.Y.; Mohney, S.E.; Badding, J.V.; Gopalan, V. Single-Crystal Germanium Core Optoelectronic Fibers. Adv. Opt. Mater. 2017, 5, 1600592. [Google Scholar] [CrossRef]
  165. Shiryaev, V.; Karaksina, E.; Churbanov, M.; Kotereva, T.; Stepanov, B.; Ketkova, L.; Evdokimov, I.; Koltashev, V.; Plotnichenko, V.; Filatov, A. Special pure germanium-rich Ga-Ge-As-Se glasses for active mid-IR fiber optics. Mater. Res. Bull. 2018, 107, 430–437. [Google Scholar] [CrossRef]
  166. Yu, B.; Sun, X.; Calebotta, G.; Dholakia, G.; Meyyappan, M. One-dimensional germanium nanowires for future electronics. J. Clust. Sci. 2006, 17, 579–597. [Google Scholar] [CrossRef]
  167. Aliane, A.; Ouvrier-Buffet, J.; Ludurczak, W.; André, L.; Kaya, H.; Vialle, C.; Benwadih, M.; Goudon, V.; Becker, S.; Hartmann, J. Fabrication and characterization of sensitive vertical PiN germanium photodiodes as infrared detectors. Semicond. Sci. Technol. 2020, 35, 035013. [Google Scholar] [CrossRef]
  168. Yamazaki, N.; Nakahama, S.; Hirao, A.; Shiraishi, Y.; Phung, H.M. Anionic Polymerization of p-Styrenyl Substituted Derivatives of Silicon, Germanium and Tin. In Contemporary Topics in Polymer Science; Springer: Berlin/Heidelberg, Germany, 1984; pp. 379–385. [Google Scholar]
  169. Zhang, J.; Zhang, B.; Li, W.; Ouyang, B.; Fan, M.; Wang, H. Behaviour of silicon ointment for power-cable insulation under external heating. J. Therm. Anal. Calorim. 2020, 140, 2749–2756. [Google Scholar] [CrossRef]
  170. Zhang, H.; Yuan, X.; Xiong, T.; Wang, H.; Jiang, L. Bioremediation of co-contaminated soil with heavy metals and pesticides: Influence factors, mechanisms and evaluation methods. Chem. Eng. J. 2020, 398, 125657. [Google Scholar] [CrossRef]
  171. Srivastava, J.; Naraian, R.; Kalra, S.; Chandra, H. Advances in microbial bioremediation and the factors influencing the process. Int. J. Environ. Sci. Technol. 2014, 11, 1787–1800. [Google Scholar] [CrossRef] [Green Version]
  172. Kumar, V.; Shahi, S.; Singh, S. Bioremediation: An eco-sustainable approach for restoration of contaminated sites. In Microbial Bioprospecting for Sustainable Development; Springer: Singapore, 2018; pp. 115–136. [Google Scholar]
  173. Boopathy, R. Factors limiting bioremediation technologies. Bioresour. Technol. 2000, 74, 63–67. [Google Scholar] [CrossRef]
  174. Shukla, A.; Srivastava, S.; D’Souza, S. An integrative approach toward biosensing and bioremediation of metals and metalloids. Int. J. Environ. Sci. Technol. 2018, 15, 2701–2712. [Google Scholar] [CrossRef]
  175. Kaur, D.; Singh, A.; Kumar, A.; Gupta, S. Genetic engineering approaches and applicability for the bioremediation of metalloids. In Plant Life under Changing Environment; Elsevier: Amsterdam, The Netherlands, 2020; pp. 207–235. [Google Scholar]
  176. Gadd, G.M. Metals, minerals and microbes: Geomicrobiology and bioremediation. Microbiology 2010, 156, 609–643. [Google Scholar] [CrossRef] [PubMed]
  177. Sher, S.; Rehman, A. Use of heavy metals resistant bacteria—A strategy for arsenic bioremediation. Appl. Microbiol. Biotechnol. 2019, 103, 6007–6021. [Google Scholar] [CrossRef]
  178. Tamayo-Figueroa, D.P.; Castillo, E.; Brandão, P.F. Metal and metalloid immobilization by microbiologically induced carbonates precipitation. World J. Microbiol. Biotechnol. 2019, 35, 58. [Google Scholar] [CrossRef]
  179. Talukdar, D.; Jasrotia, T.; Sharma, R.; Jaglan, S.; Kumar, R.; Vats, R.; Kumar, R.; Mahnashi, M.H.; Umar, A. Evaluation of novel indigenous fungal consortium for enhanced bioremediation of heavy metals from contaminated sites. Environ. Technol. Innov. 2020, 20, 101050. [Google Scholar] [CrossRef]
  180. Iyer, A.; Mody, K.; Jha, B. Biosorption of heavy metals by a marine bacterium. Mar. Pollut. Bull. 2005, 50, 340–343. [Google Scholar] [CrossRef]
  181. Deng, X.; Wang, P. Isolation of marine bacteria highly resistant to mercury and their bioaccumulation process. Bioresour. Technol. 2012, 121, 342–347. [Google Scholar] [CrossRef]
  182. Paul, D.; Kazy, S.K.; Banerjee, T.D.; Gupta, A.K.; Pal, T.; Sar, P. Arsenic biotransformation and release by bacteria indigenous to arsenic contaminated groundwater. Bioresour. Technol. 2015, 188, 14–23. [Google Scholar] [CrossRef]
  183. Barman, D.; Jha, D.K.; Bhattacharjee, K. Metallotolerant Bacteria: Insights into Bacteria Thriving in Metal-Contaminated Areas. In Microbial Versatility in Varied Environments; Springer: Berlin/Heidelberg, Germany, 2020; pp. 135–164. [Google Scholar]
  184. Chen, S.-Y.; Lin, J.-G. Bioleaching of heavy metals from contaminated sediment by indigenous sulfur-oxidizing bacteria in an air-lift bioreactor: Effects of sulfur concentration. Water Res. 2004, 38, 3205–3214. [Google Scholar] [CrossRef]
  185. Roh, Y.; Chon, C.-M.; Moon, J.-W. Metal reduction and biomineralization by an alkaliphilic metal-reducing bacterium, Alkaliphilus metalliredigens (QYMF). Geosci. J. 2007, 11, 415–423. [Google Scholar] [CrossRef]
  186. Torsvik, V.; Øvreås, L. Microbial diversity and function in soil: From genes to ecosystems. Curr. Opin. Microbiol. 2002, 5, 240–245. [Google Scholar] [CrossRef]
  187. Kirk, J.L.; Beaudette, L.A.; Hart, M.; Moutoglis, P.; Klironomos, J.N.; Lee, H.; Trevors, J.T. Methods of studying soil microbial diversity. J. Microbiol. Methods 2004, 58, 169–188. [Google Scholar] [CrossRef]
  188. Subedi, G.; Taylor, J.; Hatam, I.; Baldwin, S.A. Simultaneous selenate reduction and denitrification by a consortium of enriched mine site bacteria. Chemosphere 2017, 183, 536–545. [Google Scholar] [CrossRef]
  189. Gonzalez-Gil, G.; Lens, P.N.; Saikaly, P.E. Selenite reduction by anaerobic microbial aggregates: Microbial community structure, and proteins associated to the produced selenium spheres. Front. Microbiol. 2016, 7, 571. [Google Scholar] [CrossRef] [Green Version]
  190. Nguyen, T.H.; Ha, M.-G.; Kang, H.Y. Kinetics of microbial selenite reduction by novel bacteria isolated from activated sludge. J. Environ. Manag. 2019, 236, 746–754. [Google Scholar] [CrossRef]
  191. Ghosh, A.; Mohod, A.M.; Paknikar, K.M.; Jain, R.K. Isolation and characterization of selenite-and selenate-tolerant microorganisms from selenium-contaminated sites. World J. Microbiol. Biotechnol. 2008, 24, 1607–1611. [Google Scholar] [CrossRef]
  192. Hunter, W.J.; Manter, D.K. Reduction of selenite to elemental red selenium by Pseudomonas sp. strain CA5. Curr. Microbiol. 2009, 58, 493–498. [Google Scholar] [CrossRef]
  193. Wasi, S.; Tabrez, S.; Ahmad, M. Use of Pseudomonas spp. for the bioremediation of environmental pollutants: A review. Environ. Monit. Assess. 2013, 185, 8147–8155. [Google Scholar] [CrossRef]
  194. Presentato, A.; Piacenza, E.; Cappelletti, M.; Turner, R.J. Interaction of Rhodococcus with metals and biotechnological applications. In Biology of Rhodococcus; Springer: Cham, Switzerland, 2019; pp. 333–357. [Google Scholar]
  195. Presentato, A.; Piacenza, E.; Turner, R.J.; Zannoni, D.; Cappelletti, M. Processing of Metals and Metalloids by Actinobacteria: Cell Resistance Mechanisms and Synthesis of Metal (loid)-Based Nanostructures. Microorganisms 2020, 8, 2027. [Google Scholar] [CrossRef]
  196. Firrincieli, A.; Presentato, A.; Favoino, G.; Marabottini, R.; Allevato, E.; Stazi, S.R.; Scarascia Mugnozza, G.; Harfouche, A.; Petruccioli, M.; Turner, R.J. Identification of resistance genes and response to arsenic in Rhodococcus aetherivorans BCP1. Front. Microbiol. 2019, 10, 888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Cappelletti, M.; Presentato, A.; Piacenza, E.; Firrincieli, A.; Turner, R.J.; Zannoni, D. Biotechnology of Rhodococcus for the production of valuable compounds. Appl. Microbiol. Biotechnol. 2020, 104, 8567–8594. [Google Scholar] [CrossRef] [PubMed]
  198. Retamal-Morales, G.; Mehnert, M.; Schwabe, R.; Tischler, D.; Zapata, C.; Chávez, R.; Schlömann, M.; Levicán, G. Detection of arsenic-binding siderophores in arsenic-tolerating Actinobacteria by a modified CAS assay. Ecotoxicol. Environ. Saf. 2018, 157, 176–181. [Google Scholar] [CrossRef] [PubMed]
  199. Narasingarao, P.; Häggblom, M.M. Identification of anaerobic selenate-respiring bacteria from aquatic sediments. Appl. Environ. Microbiol. 2007, 73, 3519–3527. [Google Scholar] [CrossRef] [Green Version]
  200. Rosenfeld, C.E.; James, B.R.; Santelli, C.M. Persistent bacterial and fungal community shifts exhibited in selenium-contaminated reclaimed mine soils. Appl. Environ. Microbiol. 2018, 84, e01394-18. [Google Scholar] [CrossRef] [Green Version]
  201. Valverde, A.; González-Tirante, M.; Medina-Sierra, M.; Santa-Regina, I.; García-Sánchez, A.; Igual, J.M. Diversity and community structure of culturable arsenic-resistant bacteria across a soil arsenic gradient at an abandoned tungsten–tin mining area. Chemosphere 2011, 85, 129–134. [Google Scholar] [CrossRef]
  202. Ghosh, S.; Sar, P. Identification and characterization of metabolic properties of bacterial populations recovered from arsenic contaminated ground water of North East India (Assam). Water Res. 2013, 47, 6992–7005. [Google Scholar] [CrossRef]
  203. Jackson, C.; Harrison, K.; Dugas, S. Enumeration and characterization of culturable arsenate resistant bacteria in a large estuary. Syst. Appl. Microbiol. 2005, 28, 727–734. [Google Scholar] [CrossRef]
  204. Das, S.; Liu, C.-C.; Jean, J.-S.; Liu, T. Dissimilatory arsenate reduction and in situ microbial activities and diversity in arsenic-rich groundwater of Chianan Plain, Southwestern Taiwan. Microb. Ecol. 2016, 71, 365–374. [Google Scholar] [CrossRef]
  205. Mishra, R.R.; Prajapati, S.; Das, J.; Dangar, T.K.; Das, N.; Thatoi, H. Reduction of selenite to red elemental selenium by moderately halotolerant Bacillus megaterium strains isolated from Bhitarkanika mangrove soil and characterization of reduced product. Chemosphere 2011, 84, 1231–1237. [Google Scholar] [CrossRef]
  206. Davolos, D.; Pietrangeli, B. Phylogenetic analysis on the arsenic-resistant bacteria isolated from three different freshwater environments. Chem. Ecol. 2011, 27, 79–87. [Google Scholar] [CrossRef]
  207. Chen, S.; Shao, Z. Isolation and diversity analysis of arsenite-resistant bacteria in communities enriched from deep-sea sediments of the Southwest Indian Ocean Ridge. Extremophiles 2009, 13, 39–48. [Google Scholar] [CrossRef]
  208. Xiao, E.; Krumins, V.; Dong, Y.; Xiao, T.; Ning, Z.; Xiao, Q.; Sun, W. Microbial diversity and community structure in an antimony-rich tailings dump. Appl. Microbiol. Biotechnol. 2016, 100, 7751–7763. [Google Scholar] [CrossRef]
  209. Wang, N.; Zhang, S.; He, M. Bacterial community profile of contaminated soils in a typical antimony mining site. Environ. Sci. Pollut. Res. 2018, 25, 141–152. [Google Scholar] [CrossRef]
  210. Maltman, C.; Walter, G.; Yurkov, V. A diverse community of metal (loid) oxide respiring bacteria is associated with tube worms in the vicinity of the Juan de Fuca Ridge black smoker field. PLoS ONE 2016, 11, e0149812. [Google Scholar] [CrossRef] [Green Version]
  211. Sato, T.; Kobayashi, Y. The ars operon in the skin element of Bacillus subtilis confers resistance to arsenate and arsenite. J. Bacteriol. 1998, 180, 1655–1661. [Google Scholar] [CrossRef] [Green Version]
  212. Rosenstein, R.; Peschel, A.; Wieland, B.; Götz, F. Expression and regulation of the antimonite, arsenite, and arsenate resistance operon of Staphylococcus xylosus plasmid pSX267. J. Bacteriol. 1992, 174, 3676–3683. [Google Scholar] [CrossRef] [Green Version]
  213. Carlin, A.; Shi, W.; Dey, S.; Rosen, B.P. The ars operon of Escherichia coli confers arsenical and antimonial resistance. J. Bacteriol. 1995, 177, 981–986. [Google Scholar] [CrossRef] [Green Version]
  214. Yang, H.-C.; Fu, H.-L.; Lin, Y.-F.; Rosen, B.P. Pathways of arsenic uptake and efflux. Curr. Top. Membr. 2012, 69, 325–358. [Google Scholar]
  215. Kumari, N.; Jagadevan, S. Genetic identification of arsenate reductase and arsenite oxidase in redox transformations carried out by arsenic metabolising prokaryotes–A comprehensive review. Chemosphere 2016, 163, 400–412. [Google Scholar] [CrossRef] [PubMed]
  216. Rauschenbach, I.; Yee, N.; Häggblom, M.M.; Bini, E. Energy metabolism and multiple respiratory pathways revealed by genome sequencing of Desulfurispirillum indicum strain S5. Environ. Microbiol. 2011, 13, 1611–1621. [Google Scholar] [CrossRef] [PubMed]
  217. Malasarn, D.; Keeffe, J.R.; Newman, D.K. Characterization of the arsenate respiratory reductase from Shewanella sp. strain ANA-3. J. Bacteriol. 2008, 190, 135–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. González-Chávez, M.d.C.A.; del Pilar Ortega-Larrocea, M.; Carrillo-González, R.; López-Meyer, M.; Xoconostle-Cázares, B.; Gomez, S.K.; Harrison, M.J.; Figueroa-López, A.M.; Maldonado-Mendoza, I.E.J.F.B. Arsenate induces the expression of fungal genes involved in As transport in arbuscular mycorrhiza. Fungal Biol. 2011, 115, 1197–1209. [Google Scholar] [CrossRef] [PubMed]
  219. Li, J.; Wang, Q.; Oremland, R.S.; Kulp, T.R.; Rensing, C.; Wang, G. Microbial antimony biogeochemistry: Enzymes, regulation, and related metabolic pathways. Appl. Environ. Microbiol. 2016, 82, 5482–5495. [Google Scholar] [CrossRef] [Green Version]
  220. Rosen, B.P.; Tamás, M.J. Arsenic transport in prokaryotes and eukaryotic microbes. In MIPs and Their Role in the Exchange of Metalloids; Springer: New York, NY, USA, 2010; pp. 47–55. [Google Scholar]
  221. Shen, M.W.; Shah, D.; Chen, W.; Da Silva, N. Enhanced arsenate uptake in Saccharomyces cerevisiae overexpressing the Pho84 phosphate transporter. Biotechnol. Prog. 2012, 28, 654–661. [Google Scholar] [CrossRef]
  222. Bobrowicz, P.; Wysocki, R.; Owsianik, G.; Goffeau, A.; Ułaszewski, S. Isolation of three contiguous genes, ACR1, ACR2 and ACR3, involved in resistance to arsenic compounds in the yeast Saccharomyces cerevisiae. Yeast 1997, 13, 819–828. [Google Scholar] [CrossRef]
  223. Wysocki, R.; Bobrowicz, P.; Ułaszewski, S. The Saccharomyces cerevisiae ACR3 gene encodes a putative membrane protein involved in arsenite transport. J. Biol. Chem. 1997, 272, 30061–30066. [Google Scholar] [CrossRef] [Green Version]
  224. Meng, Y.-L.; Liu, Z.; Rosen, B.P. As (III) and Sb (III) uptake by GlpF and efflux by ArsB in Escherichia coli. J. Biol. Chem. 2004, 279, 18334–18341. [Google Scholar] [CrossRef] [Green Version]
  225. Wysocki, R.; Chéry, C.C.; Wawrzycka, D.; Van Hulle, M.; Cornelis, R.; Thevelein, J.M.; Tamás, M.J. The glycerol channel Fps1p mediates the uptake of arsenite and antimonite in Saccharomyces cerevisiae. Mol. Microbiol. 2001, 40, 1391–1401. [Google Scholar] [CrossRef]
  226. Sanders, O.I.; Rensing, C.; Kuroda, M.; Mitra, B.; Rosen, B.P. Antimonite is accumulated by the glycerol facilitator GlpF in Escherichia coli. J. Bacteriol. 1997, 179, 3365–3367. [Google Scholar] [CrossRef] [Green Version]
  227. Stevens, S.Y.; Hu, W.; Gladysheva, T.; Rosen, B.P.; Zuiderweg, E.R.; Lee, L. Secondary structure and fold homology of the ArsC protein from the Escherichia coli arsenic resistance plasmid R773. Biochemistry 1999, 38, 10178–10186. [Google Scholar] [CrossRef]
  228. Mukhopadhyay, R.; Shi, J.; Rosen, B.P. Purification and characterization of Acr2p, the Saccharomyces cerevisiae arsenate reductase. J. Biol. Chem. 2000, 275, 21149–21157. [Google Scholar] [CrossRef] [Green Version]
  229. Fu, H.L.; Jiang, X.; Rosen, B.P. Metalloid transport systems. In Biological Chemistry of Arsenic, Antimony and Bismuth; Wiley: Hoboken, NJ, USA, 2011; pp. 181–207. [Google Scholar]
  230. Maciaszczyk-Dziubinska, E.; Wawrzycka, D.; Wysocki, R. Arsenic and antimony transporters in eukaryotes. Int. J. Mol. Sci. 2012, 13, 3527–3548. [Google Scholar] [CrossRef] [Green Version]
  231. Lin, J.; Peng, T.; Jiang, L.; Ni, J.-Z.; Liu, Q.; Chen, L.; Zhang, Y. Comparative genomics reveals new candidate genes involved in selenium metabolism in prokaryotes. Genome Biol. Evol. 2015, 7, 664–676. [Google Scholar] [CrossRef] [Green Version]
  232. Wang, D.; Rensing, C.; Zheng, S. Microbial reduction and resistance to selenium: Mechanisms, applications and prospects. J. Hazard. Mater. 2021, 421, 126684. [Google Scholar] [CrossRef]
  233. Tugarova, A.V.; Kamnev, A.A. Proteins in microbial synthesis of selenium nanoparticles. Talanta 2017, 174, 539–547. [Google Scholar] [CrossRef]
  234. Connelly, K.R.; Stevenson, C.; Kneuper, H.; Sargent, F. Biosynthesis of selenate reductase in Salmonella enterica: Critical roles for the signal peptide and DmsD. Microbiology 2016, 162, 2136. [Google Scholar] [CrossRef] [Green Version]
  235. Theisen, J.; Yee, N. The molecular basis for selenate reduction in Citrobacter freundii. Geomicrobiol. J. 2014, 31, 875–883. [Google Scholar] [CrossRef]
  236. Avazéri, C.; Turner, R.J.; Pommier, J.; Weiner, J.H.; Giordano, G.; Verméglio, A. Tellurite reductase activity of nitrate reductase is responsible for the basal resistance of Escherichia coli to tellurite. Microbiology 1997, 143, 1181–1189. [Google Scholar] [CrossRef] [Green Version]
  237. Yee, N.; Ma, J.; Dalia, A.; Boonfueng, T.; Kobayashi, D. Se (VI) reduction and the precipitation of Se (0) by the facultative bacterium Enterobacter cloacae SLD1a-1 are regulated by FNR. Appl. Environ. Microbiol. 2007, 73, 1914–1920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Kuroda, M.; Yamashita, M.; Miwa, E.; Imao, K.; Fujimoto, N.; Ono, H.; Nagano, K.; Sei, K.; Ike, M. Molecular cloning and characterization of the srdBCA operon, encoding the respiratory selenate reductase complex, from the selenate-reducing bacterium Bacillus selenatarsenatis SF-1. J. Bacteriol. 2011, 193, 2141–2148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Dridge, E.J.; Watts, C.A.; Jepson, B.J.; Line, K.; Santini, J.M.; Richardson, D.J.; Butler, C.S. Investigation of the redox centres of periplasmic selenate reductase from Thauera selenatis by EPR spectroscopy. Biochem. J. 2007, 408, 19–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Lowe, E.C.; Bydder, S.; Hartshorne, R.S.; Tape, H.L.; Dridge, E.J.; Debieux, C.M.; Paszkiewicz, K.; Singleton, I.; Lewis, R.J.; Santini, J.M. Quinol-cytochrome c oxidoreductase and cytochrome c4 mediate electron transfer during selenate respiration in Thauera selenatis. J. Biol. Chem. 2010, 285, 18433–18442. [Google Scholar] [CrossRef] [Green Version]
  241. Rothery, R.A.; Workun, G.J.; Weiner, J.H. The prokaryotic complex iron–sulfur molybdoenzyme family. Biochim. Et Biophys. Acta (BBA)-Biomembr. 2008, 1778, 1897–1929. [Google Scholar] [CrossRef] [Green Version]
  242. Ma, J.; Kobayashi, D.Y.; Yee, N. Role of menaquinone biosynthesis genes in selenate reduction by Enterobacter cloacae SLD1a-1 and Escherichia coli K12. Environ. Microbiol. 2009, 11, 149–158. [Google Scholar] [CrossRef]
  243. Kesawat, M.S.; Shivaraj, S.; Kim, D.K.; Kumar, M.; Hahn, B.S.; Deshmukh, R. Metalloids and Their Role in the Biological System. In Metalloids in Plants; Wiley: Hoboken, NJ, USA, 2020; pp. 1–17. [Google Scholar]
  244. Presentato, A.; Turner, R.J.; Vásquez, C.C.; Yurkov, V.; Zannoni, D. Tellurite-dependent blackening of bacteria emerges from the dark ages. Environ. Chem. 2019, 16, 266–288. [Google Scholar] [CrossRef]
  245. Alavi, S.; Amoozegar, M.A.; Khajeh, K. Enzyme (s) responsible for tellurite reducing activity in a moderately halophilic bacterium, Salinicoccus iranensis strain QW6. Extremophiles 2014, 18, 953–961. [Google Scholar] [CrossRef]
  246. Castro, M.E.; Molina, R.; Díaz, W.; Pichuantes, S.E.; Vásquez, C.C. The dihydrolipoamide dehydrogenase of Aeromonas caviae ST exhibits NADH-dependent tellurite reductase activity. Biochem. Biophys. Res. Commun. 2008, 375, 91–94. [Google Scholar] [CrossRef]
  247. Liu, M.; Turner, R.J.; Winstone, T.L.; Saetre, A.; Dyllick-Brenzinger, M.; Jickling, G.; Tari, L.W.; Weiner, J.H.; Taylor, D.E. Escherichia coli TehB RequiresS-Adenosylmethionine as a Cofactor To Mediate Tellurite Resistance. J. Bacteriol. 2000, 182, 6509–6513. [Google Scholar] [CrossRef] [Green Version]
  248. Choudhury, H.G.; Beis, K. Crystallization and initial X-ray diffraction analysis of the tellurite-resistance S-adenosyl-l-methionine transferase protein TehB from Escherichia coli. Acta Crystallogr. Sect. F: Struct. Biol. Cryst. Commun. 2010, 66, 1496–1499. [Google Scholar] [CrossRef] [Green Version]
  249. Borghese, R.; Marchetti, D.; Zannoni, D. The highly toxic oxyanion tellurite (TeO 3 2−) enters the phototrophic bacterium Rhodobacter capsulatus via an as yet uncharacterized monocarboxylate transport system. Arch. Microbiol. 2008, 189, 93–100. [Google Scholar] [CrossRef]
  250. Nguyen, T.T.H.; Kikuchi, T.; Tokunaga, T.; Iyoda, S.; Iguchi, A. Diversity of the tellurite resistance gene operon in Escherichia coli. Front. Microbiol. 2021, 12, 1314. [Google Scholar] [CrossRef]
  251. Toptchieva, A.; Sisson, G.; Bryden, L.J.; Taylor, D.E.; Hoffman, P.S. An inducible tellurite-resistance operon in Proteus mirabilis. Microbiology 2003, 149, 1285–1295. [Google Scholar] [CrossRef] [Green Version]
  252. Pan, Y.-R.; Lou, Y.-C.; Seven, A.B.; Rizo, J.; Chen, C. NMR structure and calcium-binding properties of the tellurite resistance protein TerD from Klebsiella pneumoniae. J. Mol. Biol. 2011, 405, 1188–1201. [Google Scholar] [CrossRef]
  253. Maltman, C.; Donald, L.J.; Yurkov, V. Tellurite and tellurate reduction by the aerobic anoxygenic phototroph Erythromonas ursincola, strain KR99 is carried out by a novel membrane associated enzyme. Microorganisms 2017, 5, 20. [Google Scholar] [CrossRef] [Green Version]
  254. Turner, R.J.; Weiner, J.H.; Taylor, D.E. In vivo complementation and site-specific mutagenesis of the tellurite resistance determinant kilAteIAB from IncPα plasmid RK2Ter. Microbiology 1994, 140, 1319–1326. [Google Scholar] [CrossRef] [Green Version]
  255. Taylor, D.E. Bacterial tellurite resistance. Trends Microbiol. 1999, 7, 111–115. [Google Scholar] [CrossRef]
  256. Li, Y.; Lin, H.; Gao, P.; Yang, N.; Xu, R.; Sun, X.; Li, B.; Xu, F.; Wang, X.; Song, B. Synergistic Impacts of Arsenic and Antimony Co-contamination on Diazotrophic Communities. Microb. Ecol. 2021, 84, 44–58. [Google Scholar] [CrossRef]
  257. Liu, F.; Zhang, G.; Liu, S.; Fu, Z.; Chen, J.; Ma, C. Bioremoval of arsenic and antimony from wastewater by a mixed culture of sulfate-reducing bacteria using lactate and ethanol as carbon sources. Int. Biodeterior. Biodegrad. 2018, 126, 152–159. [Google Scholar] [CrossRef]
  258. Rodríguez-Rojas, F.; Díaz-Vásquez, W.; Undabarrena, A.; Muñoz-Díaz, P.; Arenas, F.; Vásquez, C. Mercury-mediated cross-resistance to tellurite in Pseudomonas spp. isolated from the Chilean Antarctic territory. Metallomics 2016, 8, 108–117. [Google Scholar] [CrossRef] [PubMed]
  259. Bajaj, M.; Winter, J. Se (IV) triggers faster Te (IV) reduction by soil isolates of heterotrophic aerobic bacteria: Formation of extracellular SeTe nanospheres. Microb. Cell Factories 2014, 13, 168. [Google Scholar] [CrossRef] [PubMed]
  260. Vrionis, H.A.; Wang, S.; Haslam, B.; Turner, R.J. Selenite protection of tellurite toxicity toward Escherichia coli. Front. Mol. Biosci. 2015, 2, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Hosseini, F.; Lashani, E.; Moghimi, H. Simultaneous bioremediation of phenol and tellurite by Lysinibacillus sp. EBL303 and characterization of biosynthesized Te nanoparticles. Sci. Rep. 2023, 13, 1243. [Google Scholar] [CrossRef] [PubMed]
  262. Reddy, G.K.K.; Pathak, S.; Nancharaiah, Y. Aerobic reduction of selenite and tellurite to elemental selenium and tellurium nanostructures by Alteromonas sp. under saline conditions. Int. Biodeterior. Biodegrad. 2023, 179, 105571. [Google Scholar] [CrossRef]
  263. Liang, X.; Perez, M.A.M.J.; Zhang, S.; Song, W.; Armstrong, J.G.; Bullock, L.A.; Feldmann, J.; Parnell, J.; Csetenyi, L.; Gadd, G.M. Fungal transformation of selenium and tellurium located in a volcanogenic sulfide deposit. Environ. Microbiol. 2020, 22, 2346–2364. [Google Scholar] [CrossRef] [Green Version]
  264. Ao, B.; He, F.; Lv, J.; Tu, J.; Tan, Z.; Jiang, H.; Shi, X.; Li, J.; Hou, J.; Hu, Y. Green synthesis of biogenetic Te (0) nanoparticles by high tellurite tolerance fungus Mortierella sp. AB1 with antibacterial activity. Front. Microbiol. 2022, 13, 1020179. [Google Scholar] [CrossRef]
  265. Cai, X.; Xue, C.; Owens, G.; Chen, Z. Removal of As (III) using a microorganism sustained secrete laccase-straw oxidation system. J. Hazard. Mater. 2023, 448, 130967. [Google Scholar] [CrossRef]
  266. Kujala, K.; Besold, J.; Mikkonen, A.; Tiirola, M.; Planer-Friedrich, B. Abundant and diverse arsenic-metabolizing microorganisms in peatlands treating arsenic-contaminated mining wastewaters. Environ. Microbiol. 2020, 22, 1572–1587. [Google Scholar] [CrossRef]
  267. Turner, R.J.; Borghese, R.; Zannoni, D. Microbial processing of tellurium as a tool in biotechnology. Biotechnol. Adv. 2012, 30, 954–963. [Google Scholar] [CrossRef]
  268. Harrison, J.J.; Ceri, H.; Stremick, C.A.; Turner, R.J. Biofilm susceptibility to metal toxicity. Environ. Microbiol. 2004, 6, 1220–1227. [Google Scholar] [CrossRef]
  269. Sun, W.; Sun, X.; Li, B.; Häggblom, M.M.; Han, F.; Xiao, E.; Zhang, M.; Wang, Q.; Li, F. Bacterial response to antimony and arsenic contamination in rice paddies during different flooding conditions. Sci. Total Environ. 2019, 675, 273–285. [Google Scholar] [CrossRef]
  270. Nguyen, D.D.; Ha, M.-G.; Kang, H.Y. Potential of versatile bacteria isolated from activated sludge for the bioremediation of arsenic and antimony. J. Water Process Eng. 2021, 39, 101890. [Google Scholar] [CrossRef]
  271. Park, S.-C.; Boyanov, M.I.; Kemner, K.M.; O’Loughlin, E.J.; Kwon, M.J. Distribution and speciation of Sb and toxic metal (loid) s near an antimony refinery and their effects on indigenous microorganisms. J. Hazard. Mater. 2021, 403, 123625. [Google Scholar] [CrossRef]
  272. Li, Y.; Xu, Z.; Wu, J.; Mo, P. Efficiency and mechanisms of antimony removal from wastewater using mixed cultures of iron-oxidizing bacteria and sulfate-reducing bacteria based on scrap iron. Sep. Purif. Technol. 2020, 246, 116756. [Google Scholar] [CrossRef]
  273. He, Q.; Liu, Y.; Wan, D.; Liu, Y.; Xiao, S.; Wang, Y.; Shi, Y. Enhanced biological antimony removal from water by combining elemental sulfur autotrophic reduction and disproportionation. J. Hazard. Mater. 2022, 434, 128926. [Google Scholar] [CrossRef]
  274. Wingender, J.; Neu, T.R.; Flemming, H.-C. What are bacterial extracellular polymeric substances. In Microbial Extracellular Polymeric Substances; Springer: Berlin/Heidelberg, Germany, 1999; pp. 1–19. [Google Scholar]
  275. Flemming, H.-C.; Wingender, J. The biofilm matrix. Nat. Rev. Microbiol. 2010, 8, 623–633. [Google Scholar] [CrossRef]
  276. Harrison, J.J.; Ceri, H.; Stremick, C.; Turner, R.J. Differences in biofilm and planktonic cell mediated reduction of metalloid oxyanions. FEMS Microbiol. Lett. 2004, 235, 357–362. [Google Scholar] [CrossRef]
  277. Grujić, S.; Vasić, S.; Radojević, I.; Čomić, L.; Ostojić, A. Comparison of the Rhodotorula mucilaginosa biofilm and planktonic culture on heavy metal susceptibility and removal potential. Water Air Soil Pollut. 2017, 228, 73. [Google Scholar] [CrossRef]
  278. Piacenza, E.; Presentato, A.; Zonaro, E.; Lampis, S.; Vallini, G.; Turner, R.J. Microbial-based bioremediation of selenium and tellurium compounds. In Biosorption; Derco, J., Ed.; Intech Open: London, UK, 2018; pp. 117–147. [Google Scholar]
  279. Kamyabi, A.; Nouri, H.; Moghimi, H. Synergistic effect of Sarocladium sp. and Cryptococcus sp. co-culture on crude oil biodegradation and biosurfactant production. Appl. Biochem. Biotechnol. 2017, 182, 324–334. [Google Scholar] [CrossRef]
  280. Ghorbannezhad, H.; Moghimi, H.; Dastgheib, S.M.M. Evaluation of heavy petroleum degradation using bacterial-fungal mixed cultures. Ecotoxicol. Environ. Saf. 2018, 164, 434–439. [Google Scholar] [CrossRef] [PubMed]
  281. Demeter, M.A.; Lemire, J.; George, I.; Yue, G.; Ceri, H.; Turner, R.J. Harnessing oil sands microbial communities for use in ex situ naphthenic acid bioremediation. Chemosphere 2014, 97, 78–85. [Google Scholar] [CrossRef] [PubMed]
  282. Demeter, M.A.; Lemire, J.; Yue, G.; Ceri, H.; Turner, R.J. Culturing oil sands microbes as mixed species communities enhances ex situ model naphthenic acid degradation. Front. Microbiol. 2015, 6, 936. [Google Scholar] [CrossRef] [PubMed]
  283. Frankel, M.L.; Demeter, M.A.; Lemire, J.A.; Turner, R.J. Evaluating the metal tolerance capacity of microbial communities isolated from Alberta oil sands process water. PLoS ONE 2016, 11, e0148682. [Google Scholar] [CrossRef] [Green Version]
  284. Frankel, M.L.; Bhuiyan, T.I.; Veksha, A.; Demeter, M.A.; Layzell, D.B.; Helleur, R.J.; Hill, J.M.; Turner, R.J. Removal and biodegradation of naphthenic acids by biochar and attached environmental biofilms in the presence of co-contaminating metals. Bioresour. Technol. 2016, 216, 352–361. [Google Scholar] [CrossRef] [Green Version]
  285. Golby, S.; Ceri, H.; Marques, L.L.; Turner, R.J. Mixed-species biofilms cultured from an oil sand tailings pond can biomineralize metals. Microb. Ecol. 2014, 68, 70–80. [Google Scholar] [CrossRef]
  286. Shi, L.-D.; Lv, P.-L.; Wang, M.; Lai, C.-Y.; Zhao, H.-P. A mixed consortium of methanotrophic archaea and bacteria boosts methane-dependent selenate reduction. Sci. Total Environ. 2020, 732, 139310. [Google Scholar] [CrossRef]
  287. Che, S.; Men, Y. Synthetic microbial consortia for biosynthesis and biodegradation: Promises and challenges. J. Ind. Microbiol. Biotechnol. 2019, 46, 1343–1358. [Google Scholar] [CrossRef]
  288. Zhang, S.; Merino, N.; Okamoto, A.; Gedalanga, P. Interkingdom microbial consortia mechanisms to guide biotechnological applications. Microb. Biotechnol. 2018, 11, 833–847. [Google Scholar] [CrossRef] [Green Version]
  289. Bernstein, H.C.; Carlson, R.P. Microbial consortia engineering for cellular factories: In vitro to in silico systems. Comput. Struct. Biotechnol. J. 2012, 3, e201210017. [Google Scholar] [CrossRef] [Green Version]
  290. Payne, S.M.; Mey, A.R.; Wyckoff, E.E. Vibrio iron transport: Evolutionary adaptation to life in multiple environments. Microbiol. Mol. Biol. Rev. 2016, 80, 69–90. [Google Scholar] [CrossRef] [Green Version]
  291. Nishio, K.; Hashimoto, K.; Watanabe, K. Light/electricity conversion by a self-organized photosynthetic biofilm in a single-chamber reactor. Appl. Microbiol. Biotechnol. 2010, 86, 957–964. [Google Scholar] [CrossRef]
  292. Wang, H.; Laughinghouse, H.D.; Anderson, M.A.; Chen, F.; Willliams, E.; Place, A.R.; Zmora, O.; Zohar, Y.; Zheng, T.; Hill, R.T. Novel bacterial isolate from Permian groundwater, capable of aggregating potential biofuel-producing microalga Nannochloropsis oceanica IMET1. Appl. Environ. Microbiol. 2012, 78, 1445–1453. [Google Scholar] [CrossRef] [Green Version]
  293. Agbakpe, M.; Ge, S.; Zhang, W.; Zhang, X.; Kobylarz, P. Algae harvesting for biofuel production: Influences of UV irradiation and polyethylenimine (PEI) coating on bacterial biocoagulation. Bioresour. Technol. 2014, 166, 266–272. [Google Scholar] [CrossRef]
  294. Luo, J.-H.; Chen, H.; Hu, S.; Cai, C.; Yuan, Z.; Guo, J. Microbial selenate reduction driven by a denitrifying anaerobic methane oxidation biofilm. Environ. Sci. Technol. 2018, 52, 4006–4012. [Google Scholar] [CrossRef]
  295. Bai, Y.-N.; Wang, X.-N.; Lu, Y.-Z.; Fu, L.; Zhang, F.; Lau, T.-C.; Zeng, R.J. Microbial selenite reduction coupled to anaerobic oxidation of methane. Sci. Total Environ. 2019, 669, 168–174. [Google Scholar] [CrossRef]
  296. Liu, S.; Hou, Y.; Sun, G. Synergistic degradation of pyrene and volatilization of arsenic by cocultures of bacteria and a fungus. Front. Environ. Sci. Eng. 2013, 7, 191–199. [Google Scholar] [CrossRef]
  297. Feng, T.; Lin, H.; Tang, J.; Feng, Y. Characterization of polycyclic aromatic hydrocarbons degradation and arsenate reduction by a versatile Pseudomonas isolate. Int. Biodeterior. Biodegrad. 2014, 90, 79–87. [Google Scholar] [CrossRef]
  298. Eswayah, A.S.; Smith, T.J.; Scheinost, A.C.; Hondow, N.; Gardiner, P.H. Microbial transformations of selenite by methane-oxidizing bacteria. Appl. Microbiol. Biotechnol. 2017, 101, 6713–6724. [Google Scholar] [CrossRef] [Green Version]
  299. Lai, C.-Y.; Wen, L.-L.; Shi, L.-D.; Zhao, K.-K.; Wang, Y.-Q.; Yang, X.; Rittmann, B.E.; Zhou, C.; Tang, Y.; Zheng, P. Selenate and nitrate bioreductions using methane as the electron donor in a membrane biofilm reactor. Environ. Sci. Technol. 2016, 50, 10179–10186. [Google Scholar] [CrossRef]
  300. Tang, J.; Feng, T.; Cui, C.; Feng, Y. Simultaneous biodegradation of phenanthrene and oxidation of arsenite by a dual-functional bacterial consortium. Int. Biodeterior. Biodegrad. 2013, 82, 173–179. [Google Scholar] [CrossRef]
  301. Zhang, G.; Ouyang, X.; Li, H.; Fu, Z.; Chen, J. Bioremoval of antimony from contaminated waters by a mixed batch culture of sulfate-reducing bacteria. Int. Biodeterior. Biodegrad. 2016, 115, 148–155. [Google Scholar] [CrossRef]
  302. Sutherland, I.W. Biofilm exopolysaccharides: A strong and sticky framework. Microbiology 2001, 147, 3–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Zhang, X.; Bishop, P.L. Biodegradability of biofilm extracellular polymeric substances. Chemosphere 2003, 50, 63–69. [Google Scholar] [CrossRef]
  304. Elliott, A.; Plach, J.; Droppo, I.; Warren, L. Collaborative microbial Fe-redox cycling by pelagic floc bacteria across wide ranging oxygenated aquatic systems. Chem. Geol. 2014, 366, 90–102. [Google Scholar] [CrossRef]
Figure 1. General mechanisms of metalloid tolerance in microorganisms.
Figure 1. General mechanisms of metalloid tolerance in microorganisms.
Microorganisms 11 00891 g001
Figure 2. Arsenic and antimony transporters and remediation mechanisms in bacteria (Although this microorganism is shown as a Gram-negative bacterium, transporters and mechanisms belong to different species of Gram-positive and Gram-negative bacteria).
Figure 2. Arsenic and antimony transporters and remediation mechanisms in bacteria (Although this microorganism is shown as a Gram-negative bacterium, transporters and mechanisms belong to different species of Gram-positive and Gram-negative bacteria).
Microorganisms 11 00891 g002
Figure 3. Metalloid transporter and protein involved in metalloid uptake and detoxification in fungi.
Figure 3. Metalloid transporter and protein involved in metalloid uptake and detoxification in fungi.
Microorganisms 11 00891 g003
Figure 4. Some mechanisms proposed to attenuate the toxicity of metalloids in the environment.
Figure 4. Some mechanisms proposed to attenuate the toxicity of metalloids in the environment.
Microorganisms 11 00891 g004
Table 1. The ecology of arsenic and selenium-resistant bacteria in different contaminated sites.
Table 1. The ecology of arsenic and selenium-resistant bacteria in different contaminated sites.
LocationSample SitePhylum/ClassGenus/SpeciesRef.
Louisiana, USALake Pontchartrain estuaryGammaproteobacteriaAeromonas, Rheinheimera, Pseudomonas[203]
BetaproteobacteriaRhodoferax, Variovorax, Rubrivorax, Janthinobacterium, Zoogloea
AlphaproteobacteriaBrevundimonas, Agrobacterium, Aquaspirillum, Sphingomonas, Porphyrobacter
BacillotaBacillus
ActinomycetotaCurtobacterium
BacteroidotaFlavobacterium
TaiwanChianan plain, soil, and groundwatersAlphaproteobacteriaSphingomonas, Microvirga, Phenylobacterium, Ochrobactrum, Sphingosinicella[204]
GammaproteobacteriaPseudomonas, Rheinheimera, Steroidobacter, Marinobacter, Azotobacter, Lysobacter, Acinetobacter
BetaproteobacteriaJanthinobacterium, Azonexus, Azohydromonas, Herbaspirillum, Massilia
DeltaproteobacteriaSyntrophobacter
BacillotaBacillus, Clostridium, Tumebacillus, Vulcanibacillus, Ammoniphilus, Shimazuella, Paenibacillus, Falsibacillus, Anaerobacter,
ActinomycetotaMarmoricola, Nocardioides, Pimelobacter, Arthrobacter,
BacteroidotaFlavisolibacter, Pontibacter, Parasegetibacter, Ohtaekwangia, Pedobacter,
ChloroflexotaSphaerobacter
NitrospiriaNitrospira,
GemmatimonadotaGemmatimonas,
India Groundwater of Asam AlphaproteobacteriaBrevundimonas, Bosea, Caulobacter, Rhizobium[205]
BetaproteobacteriaAcidovorax, Undibacterium, Herbaspirillum, Ralstonia
GammaproteobacteriaAcinetobacter, Pseudomonas
ActinomycetotaRhodococcus
BacillotaStaphylococcus, Bacillus
IndiaNawanshahr in Punjab, Maharashtra, and Andhra PradeshBetaproteobacteriaAchromobacter xylosoxidans, Delftia tsuruhatensis, Delftia tsuruhatensis[191]
GammaproteobacteriaPseudomonas sp., Enterobacter cowanii
BacilliBacillus fusiformis, Bacillus sp., Bacillus sphaericus
Table 2. Comparison of arsenic transportation in bacteria and yeast.
Table 2. Comparison of arsenic transportation in bacteria and yeast.
FunctionE. coliS. cerevisiaeRef.
As (V) entrance to cellsPhosphate transporters
(PiT and Pst)
Phosphate transporters (Pho84p)[220,221]
As resistance operonArsGene cluster similar to ars[213,222]
As (III) extrusion transporter
(Active transportation)
ATP-fueled
(ArsA and ArsB)
H+-coupled (Acr3p)
(Proton gradient generated by plasma membrane proton ATPase Pma1p)
[223,224]
As (III) efflux
(Passive transportation)
MIP
(GlpF)
MIP (Fps1p)[224,225,226]
As (V) to As (III) reductionArsCAcr2p[227,228]
Table 3. Genes involved in metalloid uptake in S. cerevisiae [229,230].
Table 3. Genes involved in metalloid uptake in S. cerevisiae [229,230].
Gene/ProteinDeletionOverexpressionRef.
BOR1More susceptible rather than wild type to the toxic level of boronIncreases resistance to boric acid and decreases efflux rate[229,230]
FPS1Reduction of boron accumulationNo effect on the accumulation
DUR3Reduction of boron accumulationNo effect on the accumulation
fps1-Δ1 allele lacking
N-terminal extension
-hypertensive to As (III) and Sb (III) and accumulated more As (III)
FPS1Increased resistance to As (III) and Sb (III)-
HOG1Increased sensitivity to As (III) and Sb (III) due to upregulation of metalloid intake-
Table 4. Dominant taxa in the presence of contaminants.
Table 4. Dominant taxa in the presence of contaminants.
ContaminantsDominant Groups in the Presence of ContaminantsRef.
Phenanthrene Pseudomonas and Achromobacter spp.(Tang et al., 2013) [300]
AsO33− Pseudomonas and Alcaligenes spp.
Phenanthrene + AsO33− Gammaproteobacteria (Pseudomonas) and Bettaproteobacteria (Achromobacter and Alcaligenes)
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

Lashani, E.; Amoozegar, M.A.; Turner, R.J.; Moghimi, H. Use of Microbial Consortia in Bioremediation of Metalloid Polluted Environments. Microorganisms 2023, 11, 891. https://doi.org/10.3390/microorganisms11040891

AMA Style

Lashani E, Amoozegar MA, Turner RJ, Moghimi H. Use of Microbial Consortia in Bioremediation of Metalloid Polluted Environments. Microorganisms. 2023; 11(4):891. https://doi.org/10.3390/microorganisms11040891

Chicago/Turabian Style

Lashani, Elham, Mohammad Ali Amoozegar, Raymond J. Turner, and Hamid Moghimi. 2023. "Use of Microbial Consortia in Bioremediation of Metalloid Polluted Environments" Microorganisms 11, no. 4: 891. https://doi.org/10.3390/microorganisms11040891

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