Next Article in Journal
Humic Substances Isolated from Recycled Biomass Trigger Jasmonic Acid Biosynthesis and Signalling
Next Article in Special Issue
Using Mediterranean Native Plants for the Phytoremediation of Mining Sites: An Overview of the Past and Present, and Perspectives for the Future
Previous Article in Journal
Environmental Factors Affecting Monoterpene Emissions from Terrestrial Vegetation
Previous Article in Special Issue
A Review on Remediation of Iron Ore Mine Tailings via Organic Amendments Coupled with Phytoremediation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Microbial-Assisted Remediation of Cadmium-Contaminated Soil

1
Department of Agronomy, Faculty of Agriculture and Environment, The Islamia University of Bahawalpur, Bahawalpur 63100, Pakistan
2
Key Laboratory of Vegetation Restoration and Management of Degraded Ecosystems, South China Botanical Garden, Chinese Academy of Sciences, Guangzhou 510650, China
3
University of Chinese Academy of Sciences, Beijing 100039, China
4
Department of Botany, The Islamia University of Bahawalpur, Bahawalpur 63100, Pakistan
5
Institute of Soil and Environmental Sciences, University of Agriculture, Faisalabad 38040, Pakistan
6
Department of Soil and Environmental Sciences, Ghazi University, D. G. Khan 32200, Pakistan
7
Institute of Marine and Environmental Technology, University of Maryland Center for Environmental Science, Baltimore, MD 21202, USA
8
Land Resources Research Institute (LRRI), National Agricultural Research Centre (NARC), Islamabad, Pakistan
9
Department of Agronomy, University of Agriculture, Faisalabad 38040, Pakistan
10
Department of Botany, University of Agriculture, Faisalabad 38040, Pakistan
11
Department of Agriculture, Extension, Azad Jammu & Kashmir, Pakistan
12
Soil and Environmental Biotechnology Division, National Institute for Biotechnology and Genetic Engineering, Constituent College of Pakistan Institute of Engineering and Applied Sciences, Faisalabad, Pakistan
13
Tasmanian Institute of Agriculture, University of Tasmania, Hobart, TAS 7001, Australia
*
Author to whom correspondence should be addressed.
Plants 2023, 12(17), 3147; https://doi.org/10.3390/plants12173147
Submission received: 16 March 2023 / Revised: 29 August 2023 / Accepted: 29 August 2023 / Published: 31 August 2023

Abstract

:
Soil contamination with cadmium (Cd) is a severe concern for the developing world due to its non-biodegradability and significant potential to damage the ecosystem and associated services. Industries such as mining, manufacturing, building, etc., rapidly produce a substantial amount of Cd, posing environmental risks. Cd toxicity in crop plants decreases nutrient and water uptake and translocation, increases oxidative damage, interferes with plant metabolism and inhibits plant morphology and physiology. However, various conventional physicochemical approaches are available to remove Cd from the soil, including chemical reduction, immobilization, stabilization and electro-remediation. Nevertheless, these processes are costly and unfriendly to the environment because they require much energy, skilled labor and hazardous chemicals. In contrasting, contaminated soils can be restored by using bioremediation techniques, which use plants alone and in association with different beneficial microbes as cutting-edge approaches. This review covers the bioremediation of soils contaminated with Cd in various new ways. The bioremediation capability of bacteria and fungi alone and in combination with plants are studied and analyzed. Microbes, including bacteria, fungi and algae, are reported to have a high tolerance for metals, having a 98% bioremediation capability. The internal structure of microorganisms, their cell surface characteristics and the surrounding environmental circumstances are all discussed concerning how microbes detoxify metals. Moreover, issues affecting the effectiveness of bioremediation are explored, along with potential difficulties, solutions and prospects.

1. Introduction

In the last two decades, the quality of human life has improved significantly. However, developmental activities have occurred at the expense of the environment’s quality [1]. Soil and the environment are contaminated due to higher concentrations of metalloids and heavy metals (HMs) resulting from rapidly expanding industrial wastes, excessive use of automobiles, resource extraction, petrochemical spillage, metallurgy and anthropogenic activities [2]. A heavy metal is any metallic substance with a relatively higher density and is toxic even at low concentrations [3]. Heavy metals include elements such as aluminum (Al), arsenic (As), antimony (Sb), beryllium (Be), cadmium (Cd), chromium (Cr), copper (Cu), lead (Pb), mercury (Hg) and nickel (Ni) [4]. The persistent nature of these toxic HMs causes harm to humans, plants and animals at higher levels [5].
Cadmium is one of the most dangerous HMs to living organisms [6], mainly due to its higher toxicity and severe extent of bioaccumulation [7]. It adversely impacts human health by accumulating in the kidney and causing renal tubular damage and emphysema [8]. Cd has persisted in soil for decades, depending on multiple factors, including soil type, redox potential, pH, clay contents, organic matter, plant uptake and leaching [9]. Cd presents a unique concern due to its notable mobility in soil environments. Unlike some heavy metals, Cd exhibits a relatively high degree of mobility within soils, facilitated by factors such as soil pH, organic matter content and redox potential. This mobility renders Cd more hazardous even at relatively low soil concentrations, as it can readily leach into groundwater and accumulate in crops, posing environmental and human health risks. Cd toxicity negatively affects plant functioning by inhibiting carbon fixation, reducing chlorophyll synthesis and minimizing photosynthetic activity [10]. Cd-induced phytotoxicity leads to plant morphological alterations, such as chlorosis and the suppression of lateral root formation [11]. Additionally, Cd exposure induces osmotic stress in plants by reducing relative leaf water content, stomatal conductance and transpiration, ultimately leading to tissue damage [12]. Furthermore, the toxicity of Cd results in the overproduction of reactive oxygen species (ROS), damaging plant membranes and destroying cell organelles [13]. Cd toxicity also reduces the uptake and transportation of mineral elements, leading to stunted growth with ultimate yield penalties on field crops [14]. The increased mobility of Cd underscores the urgency of effective remediation strategies to mitigate its potential widespread contamination and its subsequent adverse effects on ecosystems and agriculture.
Cd remediation mitigates or eliminates Cd contamination from environmental systems, particularly soil. Cadmium, a highly toxic heavy metal, poses significant health and ecological risks even at relatively low concentrations due to its mobility within soils [12,15]. This process involves various strategies to reduce Cd’s presence, minimizing its potential harm to human health, ecosystems and agricultural productivity. Remediation methods can be broadly categorized into physicochemical approaches, which involve chemical treatments and physical processes, and bioremediation, which employs living organisms such as microorganisms and plants to transform or remove Cd from the soil matrix. The most commonly employed approaches include chemical oxidation and reduction, precipitation, electrochemical treatment, solvent extraction, ion exchange, filtration, reverse osmosis, recovery by evaporation and soil washing with chelating chemicals [16,17]. However, one major drawback of these traditional processes is the creation of toxic heaps, sludge and secondary pollutants [18]. Therefore, it is necessary to continuously monitor the stability of immobilized HMs such as Cd [19]. Moreover, conventional remediation techniques can only remove Cd to a certain degree. In addition, traditional remediation requires expensive chemicals, significant energy and investment [20].
In contrast to physicochemical procedures, bioremediation is an environment-friendly technique that utilizes plants and microorganisms (such as fungi, bacteria and algae) to aid in the restoration of contaminated soil to its original state [21]. Bioremediation harnesses the natural metabolic capabilities of these organisms to convert Cd into less harmful forms, offering a sustainable and eco-friendly solution to Cd contamination. Biological techniques such as biosorption and bioaccumulation offer an advantage in removing HMs from polluted resources [18,22]. In a natural ecosystem, microbes are widely distributed and thrive in HM-polluted environments [12]. However, the ability of microbes to remediate contaminants can halt when they run out of food [23]. An enrichment method for the isolation of microbes that combines the properties of (1) the degradation of a chosen pollutant and (2) excellent root colonization has been developed [24,25] to ensure that these microbes can access the best available food source in soil, namely root exudates [26]. Plant root exudates, including organic acids, alcohols and sugars, serve as energy sources for soil microflora and promote microbial activity and growth [27]. According to Sabae et al. [28], some root exudates may also function as chemotactic signals for microbes. Furthermore, plant roots enhance water movement and loosen the rhizosphere, which improves microbial colonization [29,30]. As a result, these microbes transform hazardous HMs into non-toxic forms. Throughout bioremediation, these microbes transform organic pollutants into end products, including H2O, CO2 and metabolites, which are the primary substrates for cell growth [31]. Microbes maintain a defense system against HM contamination in the rhizosphere via two mechanisms: (i) the biosynthesis of enzymes that break down specific contaminants and (ii) persistence that can withstand associated HMs [18].
Despite numerous individual efforts to evaluate the potential of various microbes for remediating Cd, such as bacteria, fungi and algae, no comprehensive review covers the multivariate features of plant growth-promoting microbes and their strategies and mechanisms for decontaminating Cd-contaminated soil. This review covers some new aspects and dimensions of the bioremediation of Cd-contaminated soils. Here, we mainly review the recently published literature from 2010 to 2022. The main objective of this review is to highlight the bioremediation potential of various microorganisms, especially bacteria, fungi and algae, individually and in combination with plants. Different mechanisms, i.e., indirect and direct mechanisms, adapted by microorganisms to detoxify Cd, are also discussed. Furthermore, factors, i.e., water content, temperature, pH, nutrient availability, moisture content and pollutant bioavailability, which can influence the bioremediation of Cd in contaminated soil, are also explored. Finally, the present review explores field application knowledge through case studies, challenges and prospects.

2. Review Methodology

The review methodology employed in this study involved a comprehensive search for peer-reviewed research articles using a range of relevant keywords. The keywords used in the search included terms such as “contaminant”, “Cd uptake”, “toxicity”, “accumulation”, “dynamics”, “seed germination”, “oxidative stress”, “antioxidant enzymes”, “photosynthetic rate”, “growth patterns”, “plant physiology”, “lipid peroxidation”, “nutrients uptake”, “mitigation measures”, “microbes”, “immobilization”, “bioremediation”, “mechanism”, “PGPRs”, “membrane and enzyme technology”, “genetic and metabolic engineering”, “metagenomics” and “nanoparticle”.
Six prominent databases were utilized to ensure a wide range of literature sources: Sciencedirect, Google Scholar, Web of Science, Researchgate, Scopus and freefullpdf. These databases are renowned for their vast collections of peer-reviewed articles and are widely recognized in the academic community. The search strategy resulted in identifying approximately 336 relevant articles published between 2010 and 2023. These articles were selected based on their relevance to the research topic and the inclusion of pertinent information regarding the effects of contaminants, specifically Cd, on various aspects of plant biology and physiology and bioremediation techniques to decontaminate Cd. To further enhance the comprehensiveness of the review, the reference lists of the identified articles were examined, and any additional relevant papers cited within these articles were also reviewed. By employing this rigorous methodology, the researchers aimed to ensure that the review encompassed a wide range of up-to-date and credible information on the subject matter, thereby strengthening the overall validity and reliability of the findings presented in the study.

3. Sources of Cadmium

According to an annual United Nations Environment Program (UNEP) report, the amount of Cd released into the environment via different sources varies between 150 and 2600 tons [32]. Natural sources of Cd include rock weathering and air soil particles, e.g., from deserts, sea spray, forest fires, biogenic materials, volcanoes and hydrothermal vents [33]. Various rock types contain significant amounts of Cd, ranging from 0.006 to 8.4 ppm. It is estimated that wind-blown ash deposits approximately 0.25 × 106 kg of Cd annually, volcanic eruptions contribute 0.5 × 106 kg, wildfires contribute 0.01 × 106 kg, and salt seal aerosols contribute 0.002 × 106 kg [34]. Cadmium can enter the soil via the long-term application of fertilizers, pesticides and animal manure [35].

4. Cadmium Toxicity and Plants

4.1. Seed Germination and Seedling Growth

Seed germination is considered the most essential activity following the onset of emergence and is accompanied by a release from dormancy [36]. Growth regulators, such as gibberellic acid (GA), auxin and abscisic acid (ABA), regulate seed germination and early seedling growth [37]. Phytohormones GA and ABA work antagonistically, by which elevated levels of ABA inhibit seed germination and regulate seed dormancy, and GA induces germination [36]. The germination mechanism depends on the seeds’ GA/ABA ratio, which acts as a central hub during integration with environmental stresses [38]. Some other plant hormones, including strigolactones, cytokinin and brassinosteroids, either induce or retard seed germination [39]. Cd inhibits germination and reduces the growth of germinating seedlings [40,41]. The suppression of seed germination mainly occurs due to inhibiting metabolic and physiological processes [42,43]. Cd toxicity in seeds reduces water absorption capacity and starch digestion and impairs the growth of growing embryos [44]. Besides hormonal disturbance, ROS that are too low cause the failure of normal seed germination, and excessive Cd toxicity leads to higher ROS (H2O2, free radicals, singlet O2) and damages the growing embryo [45]. Excess ROS accumulation occurs due to low levels of cellular antioxidants (SOD, POD and CAT) [46]. At low levels, H2O2 favors seed germination and acts as an oxidative spell of germination. A low concentration of H2O2 actively oxidizes several proteins, enzymes and mRNAs [47]. Furthermore, Cd contamination impairs germination by reducing seed water uptake, blocking the transport of soluble sugar to the embryonic axis and decreasing the starch release capacity of the embryo due to the inactivation of α-amylase [48]. The inactivation of α-amylase mainly occurs when chemically similar Cd ions starve beneficial Ca ions in seeds [36]. During the early phase of seed germination, Cd and Ca-calmodulin compete to replace Cd with Ca ions for regulating normal germination and maintaining membrane integrity [43].

4.2. Cadmium-Induced Changes in Growth and Development

Cd is considered a non-essential element for normal plant growth and development. It can damage growth-related traits in various plants [49], as shown in Figure 1. Cd ions bind to functional proteins and make them dysfunctional. This leads to the degradation of the photosynthetic apparatus and a reduction in photosynthetic pigments, ultimately reducing biomass production and plant growth [50]. The absorption and translocation of Cd2+ reduce the leaf surface area and, subsequently, photosynthetic products [49]. Long-term exposure to Cd at the root level leads to mucilaginous, decomposing and necrotic roots. These changes ultimately lead to leaf chlorosis, rolling and premature leaf falls [15,51]. Additionally, excessive Cd accumulation in the rhizosphere disturbs the root system, inhibiting primary and lateral root growth, causing root stiffness and turning the roots brownish and twisted [52]. Cd toxicity also reduces mitotic divisions and proliferates the cortical cells of roots, thereby reducing root length and minimizing dry biomass [53]. To mitigate the adverse effect of Cd on roots, plants increase root parenchyma and cortical cell area to make an efficient flow of nutrients and water [54].

4.3. Impact on Amino Acids, Proteins and Organic Osmolytes

During HM stress, plants synthesize various low-molecular-weight organic osmolytes, such as amino acids, total soluble sugars and proteins [55]. These organic osmolytes act as signaling molecules and free radical scavengers and can modulate the stomatal aperture while reducing oxidative stress [56]. Amino acids regulate pH, enzyme synthesis and redox homeostasis [57]. During HM stress, amino acids upregulate osmotic adjustments, maintain integral proteins and ion homeostasis, neutralize redox potential and scavenge ROS by maintaining plant antioxidant levels [58]. However, Cd toxicity induces damage to proteins in the cell cytoplasm [49]. Its uptake by roots can reduce proteins via increased H2O2, LPX and free radicals [59].

4.4. Plant Water Relations

Exposure to Cd stress leads to adverse changes in the water status of plants [60]. Cd reduces the extent of water availability and nutrient translocation at the root level, disturbing short-distance symplast and apoplast pathways [53]. Excess Cd ions in root cells also lower the water status above ground [61]. Water balance disturbances can lead to low membrane integrity during lipid peroxidation [62]. Cd negatively influences the physiological mechanism of the cell water status and gas exchange, impairing plant metabolic processes [63]. Exposure to Cd in plants results in imbalanced nutrient and water uptake, reducing photosynthetic performance and biomass production [64]. Moreover, low water availability increases Cd ion sequestration in root cells and causes oxidative damage to root cells [65].

4.5. Impact on Photosynthesis

Cadmium poses a severe threat to the photosynthetic system of plants [66], and its accumulation in leaves can cause oxidative stress and a decline in transpiration rate, leading to stomatal closure [67]. Exposure to Cd damages various crucial components of plants’ photosynthetic systems, including photosystems, reaction centers and antenna complexes [68]. Photosynthetic efficiency in leaves depends on the availability of Fe2+ and the synthesis of other accessory pigments. Cd can inhibit the activity of Fe3+ reductase, causing a reduction in Fe2+ and leading to a decline in the photosynthetic performance of plants [69]. Cd-induced low pigment synthesis can result in a minimum density of chloroplast and chlorosis [70]. Furthermore, Cd can disrupt the shape of chloroplast and inflate thylakoids [71]. The toxicity of Cd also adversely impacts mesophyll structure and poorly developed anatomical structures that alter the biochemistry of photosynthesis. Negative interactions with SH groups can inhibit photosynthetic enzymes [72]. The impact of Cd toxicity in different plant species is presented in Table 1.

5. The Role of Microbes in the Bioremediation of Cd-Contaminated Soils

Microorganisms have been utilized for the remediation of HM pollution through various techniques, including the immobilization, adhesion, processing, oxidation and volatilization of HMs [12,15]. To maximize bioremediation efficiency, it is imperative to identify the underlying mechanisms that govern the behavior and proliferation of microorganisms in contaminated sites and their reactions to environmental fluctuations [86]. Bioremediation techniques include interactions between microbes and metals, biosorption, biotransformation, biomineralization, bioaccumulation and bioleaching. Microorganisms that depend on chemicals for growth and development can remove them from soil [87]. In addition to dissolving metals, microbes can oxidize and reduce transition metals. Cell membranes can be harmed by different organic solvents [31]. However, microbial cells can evolve defense mechanisms, such as creating solvent efflux or hydrophobic pumps, to prevent damage to the outer membrane [88]. Plasmid-encoded or energy-dependent metal efflux systems have been found in numerous bacteria that resist metals such as Cr, Cd and As [89]. The microbe-assisted phytoremediation of Cd from soil under different experimental settings is detailed in Table 2.

5.1. Remediation of Cd by Bacteria

Bacterial biomasses, both dead and alive, can be used in bioremediation. Cd can be removed through the processes of biosorption and bioaccumulation [111]. By modifying the cell wall–plasma membrane complex and depositing Cd into the cell wall, bacteria can resist the harmful effects of Cd [15]. Cd enters bacterial cells through the absorption mechanisms of divalent cations such as manganese (Mn2+) or zinc (Zn2+) [112]. The surface of bacterial cells contains several functional groups, including carboxyl, phosphonate, sulfonate, hydroxyl and amide groups, which can absorb Cd from soil solutions [113]. Some of the most effective Cd-bioremediating bacteria include Streptomyces R25, Fomitopsis pinicola CCBAS 535 [114] and Pseudomonas aeruginosa [111]. Compared to the control, Bacillus subtilis L. and Saccharomyces cerevisiae L. absorbed 75.76 and 69.56% of Cd from contaminated soil after five days of inoculation [26]. Bacillus subtilis L. can improve water absorption and minimize electrolyte leakage (EL) to promote plant growth and reduce Cd toxicity [115]. Bacillus licheniformis L. increases the dispersion of Cd and its accumulation in plants under contaminated soil conditions, which lowers the amount of hazardous Cd in soil [25,116]. In wheat (Triticum aestivum), the inoculation of Bacillus siamensis L. reduced Cd toxicity by reducing the malondialdehyde (MDA) content and increasing the catalase (CAT) and superoxidase (SOD) contents [117]. Moreover, the inoculation of Bacillus siamensis L. increased wheat crop yield under Cd stress by increasing membrane stability, total soluble sugars, amino acid synthesis and photosynthetic activity [117].
Moreover, the application of plant growth-promoting rhizobacteria (PGPR) can be a significant factor in bacterial-assisted Cd bioremediation [12,15]. PGPR inoculants, such as Rhodococcus sp. 4N-4, Flavobacterium sp. 5P-4, Variovorax paradoxus 2C-1, [118], Flavobacterium sp., Kluyvera ascorbata SUD165 and SUD165/26, Pseudomonas tolaasii RP23, P. Fluorescens RS9, Rhodococcus sp., Variovorax paradoxus [118,119], Pseudomonas aeruginosa [120,121], Pseudomonas sp., Bradyrhizobium sp., Ochrobactrum cytisi [122], Bacillus megaterium [123] and Rhodobacter sphaeroides [18], have significantly been used to mitigate Cd toxicity in various agricultural and horticultural crops grown in Cd-contaminated soils. In addition, PGPR release antifungal chemicals, such as hydrogen cyanide, and mobilize nutrients, particularly phosphates, from soil to defend plants from fungal disease [124]. Certain PGPR release 1-aminocyclopropane-1-carboxylate (ACC) deaminase, which helps plants recover from biotic and abiotic stress [124,125]. Therefore, PGPR can increase species’ capacity to remove Cd through bioremediation and may be used in phytoremediation techniques.

5.2. Remediation of Cd by Fungi

Mycoremediation is the process of using fungi to bioremediate Cd [18]. It involves utilizing fungi’s extracellular enzymes or potential metabolic capacity to reduce organic and inorganic contaminants in natural resources [126]. Fungi have been widely accepted for their involvement in the remediation of Cd due to their physical contact, low cost, wide availability, increased cell-to-surface ratio and fungal enzymatic activities with the surrounding environment, as well as their ability to be farmed on a large scale [127]. Fungi have demonstrated enormous physiological and metabolic capacity to digest harmful substances and lower the associated environmental concerns with these molecules via chemical changes or affecting chemical bioavailability [128]. Mycoremediation involves intra- and extracellular precipitation, valence transformation and an active uptake mechanism [129,130]. Mycelium’s role in fungus degradation makes it an active participant in bioremediation. Fungal mycelia penetrate the air spaces of polluted soils [131]. They secrete extracellular enzymes, organic acids and complex organic compounds, which aid in the solubilization and chelation of metal ions [132].
Mycoremediation involves several mechanical routes, including extrusion, sequestration, biotransformation, avoidance/exclusion and biodegradation [130]. It is a useful technology regarding the hyperaccumulation of contaminants. Hyper-accumulators tend to accumulate pollutants with low concentrations due to their low biomass, whereas fungi accumulate more contaminants [133]. Interactions between fungal species and hyperaccumulator plants, leguminous plants and other herbs can lead to an efficient phytoremediation strategy [134]. Arbuscular mycorrhizal fungus (AMF) creates a physical connection directly between soil and plants, increasing the rhizosphere surface area and improving nutrient absorption [135]. However, the increased surface area also increases the vulnerability of plants to contaminants. Several regulating factors affect exposure and, subsequently, metal toxicity, such as plant and fungal species, their ecotypes, the bioavailability of pollutants, soil properties, soil fertility, root growth and light intensity [136]. In Cd bioremediation, Microsphaeropsis sp. LSE10 was found to have the highest removal efficiency of 247.5 mg g−1 compared to other fungal species [137]. Mycorrhizae can act as a barrier that prevents contaminants from passing through to plants. Metals can also attach to fungal hyphae, indicating that they may react to contamination [138,139]. Fungal vesicles, spores, extraradical mycelia and intraradical mycelia are essential for the accumulation of metals and the chelation of contaminants [140].
Moreover, plants inoculated with AMF may produce molecules that chelate Cd complexes, such as phytochelatins, metallothioneins and glutathione [141]. According to Zhang et al. [139], glomalin produced by AMF mycelia can also bind more metals. Hence, it can significantly immobilize HMs and promote host plant tolerance to harsh situations. Certain fungal species, such as Trichoderma spp. and Piriformospora indica, are adaptable due to their ability to grow in soils with high contaminants [142] Data suggest that Trichoderma metal tolerance strains can significantly influence the bioaccumulation of Cd and other contaminants [143]. Trichoderma simmonsii L. (UTFC 10063) has the potential to bioaccumulate Cd and reduce its Cd toxicity by 46.1% [144]. Aspergillus niger L. significantly eliminated Cd ions in soils by 84% [145]. Reports indicate that Trichoderma atroviride L. affects rapeseed’s uptake and the translocation of Cd, Ni and Zn [146]. Other fungal varieties, such as Trichoderma mutant L. [147], Talamyces emersonii L., Basidiomycetes [148,149], Trichoderma harzianum L., Trichoderma tomentosum L. and Trichoderma asperellum L. [150], aid in the remediation of Cd from contaminated agricultural soils [144]. It is important to exploit fungi that can remediate contaminated soil through bioaccumulation, bio-volatilization and biosorption to reduce Cd contamination from agricultural soils.

5.3. Remediation of Cd by Algae

Phyco-remediation involves using algae and cyanobacteria for Cd removal, assimilation, degradation, etc. [151]. Due to its greater algal availability, low operational costs, low sludge generation, facile application and low nutritional demand, bioremediation with algae is advantageous [152]. In controlled circumstances, Kumar et al. [153] investigated the brown color variant of Kappaphycus alvarezii that absorbed 3.365 mg of Cd 100 g−1 fresh weights. Previous studies have shown that the Chlorella, Ulva, Sargassum, Fucus and Ascophyllum species can absorb HMs. Cd accumulates on several cell wall layers of Spirulina maxima [154,155,156]. Of Microcystis aeruginosa, 90% have a strong affinity for Cd2+. The Cd2+ concentration of 1 mg L−1 showed strong population growth for Porphyridium cruentum and reduced its population growth at 5 mg L−1 [157]. According to Saunders et al. [158], Cd can be removed from coal-fired power plant wastewater using the algae species Hydrodictyon, Oedogonium and Rhizoclonium. Spirogyra hyaline’s dry algal biomass has been shown by Kumar and Oommen [159] as a useful biosorbent for Cd remediation. The effectiveness of Chlorella vulgaris and Chlamydomonas reinhardtii in removing Cd was established by Kotrba et al. [160]. According to Tuzen and Sari [161], Chlamydomonas reinhardtii biomass can be biosorbents in the removal of Cd2+. The algal species or strains, sorption process, immobilization techniques, manipulation of Cd binding sites, economic viability of remediation technologies, etc., all play a role in the phyco-remediation of Cd commercially. Table 3 thoroughly summarizes the efficiency of various bacterial, fungal and algal cultures for Cd bio-removal.

6. Mechanisms Involved in Bioremediation by Microbes

6.1. Direct Mechanisms

6.1.1. Nitrogen Fixation

Nitrogen (N) is a vital nutrient for plant growth, as it is required for chlorophyll production, photosynthesis and cell division. N is often the key limiting element for plant growth and development [196]. It is the most abundant element in Earth’s atmosphere but only exists in its inert state in certain modified prokaryotes, such as some cyanobacteria, actinomycetes and eubacteria [197]. Legumes can form symbiotic relationships with rhizobia that fix N in soil, such as Azorhizobium, Allorhizobium, Bradyrhizobium, Mesorhizobium, Rhizobium or Sinorhizobium [198]. Several PGPR, such as Rhizobium, which forms a symbiotic relationship with root nodules of leguminous plants, can continually transform atmospheric N into nitrate and ammonium for plant use by producing nitrogenase enzymes [199].
Certain PGPR may improve plant absorption of N in rhizosphere soil [200]. For instance, inoculation with PGPR can boost N absorption and encourage the development of tomato (Solanum lycopersicum) plants under a rising N supply [201], similar to Singh et al. [202], who selected 22 isolates from the rhizosphere of sugarcane (Saccharum officinarum). Similarly, several Bacillus isolates have been shown to exhibit N-fixing and biocontrol abilities. Rhizobia produce Nod Factors (NFs) in response to plant root exudates that include (iso) flavonoids, which start the symbiotic process leading to the initiation of bacterial infection. Nodules, which develop in roots and, in rare instances, stems, result from this molecular conversation. Plants stipulate a carbon supply to microbes to fuel the symbiotic biological N fixation (BNF) process and a microaerophilic environment within nodules, consistent with nitrogenase (Nase) complex functioning. The Nase enzyme converts dinitrogen from the atmosphere to ammonia, which is subsequently converted into organic forms and is expelled from nodules to support plant development. Because N-limited conditions are a typical characteristic of metal-contaminated soils, symbiotic BNF also makes legumes the ideal pioneers to invade and repair the quality and health of these habitats [203]. This ability, combined with the deep-dwelling root systems and large biomass of legumes, makes them suitable candidates for the effective phytoremediation of Cd [204].

6.1.2. Phosphate Solubilization

Phosphate-solubilizing bacteria (PSB) have been found to enhance plant growth in Cd-contaminated soils by providing phosphorus (P), making them useful for the remediation of Cd. Despite being a crucial component for plant development, soil seldom contains enough P. Both organic (Po, average 50%) and inorganic (Pi, average 50%) forms of P are found in nature [205,206]. Roots can absorb them because none are soluble (typically no more than 5%) [207]. Plants may absorb monobasic (H2PO4) and dibasic (HPO42−) ions. Pi is mostly soluble when the soil’s pH decreases due to the synthesis of low-molecular-weight organic acids. On the other hand, phosphoric esters are hydrolyzed by phosphatase during the mineralization of organic P [208]. PSB activity produces enough P, significantly reducing the need for chemical fertilizers [209]. PSB dissipates inorganic phosphates through organic acid secretion, which increases phosphate solubility by ionizing protons, lowering the pH and combining PO43− to create HPO42− or H2PO4−. The organic acid may also form Al3+, Ca2+ and Fe3+ complexes, making PO43− available for plant uptake. Similarly, PSB-supported phytoextraction may enhance the mobility of Cd in soil, as revealed by various research. Endophytic Rahnella sp. JN6 efficiently solubilized 8.8 mg L−1 Cd and 133.54 mg L−1 phosphate in a liquid culture and increased Cd accumulation in mustard (Brassica napus) [210]. Burkholderia sp. J62 boosted maize and tomato biomass by solubilizing 25.8 mg L−1 Cd and 234 mg L−1 phosphate in a culture solution. The favorable impact on plants encourages increased phytoextraction or phytostabilization effectiveness in Cd-contaminated soil [211].

6.1.3. Phytohormone Production

Metals may also affect plant growth and development. Numerous studies have shown that plants grown under metal stress conditions experience damage to their membrane system, which affects the structure and function of organelles and various physiological and biochemical processes in their tissues. Lignans, or phytohormones, are active chemical compounds produced by plants that may have specific physiological effects on plants even in very low quantities [212]. They are strongly connected to root growth and may control various plant life cycle functions. Environmental challenges, such as HM toxicity, harsh temperatures, nutrient deficiencies and drought, which may cause a range of unfavorable physiological and chemical responses in plants, continually affect plant development during phytoremediation [213,214]. PGPR may increase plant tolerance to such stresses and promote plant development in Cd-contaminated soil by preserving the nutritional status and modifying phytohormonal balance by synthesizing plant growth regulators. Several studies have shown that PGPR are important in phytohormone synthesis, which may govern plant growth, development and physiological processes and influence biological and non-biological stress responses. Various PGPR have been identified that produce phytohormones, such as IAA, auxins, cytokinin and gibberellins, during harsh conditions, ultimately promoting plant development and enhancing the plant’s ability to withstand under environmental stress due to Cd contamination [215,216]. The Bacillus sp. MT7 strain was isolated from maize rhizosphere soil and tested in a tomato rhizosphere [217]. The results showed that MT7 produced 14.44 g mL−1 IAA 4 days after inoculation, promoting plant growth and showing tolerance against Cd stress. Pseudomonas fluorescens may boost the wedge’s development and physiological processes (Sedum alfredii) by generating IAA and improving plant Cd absorption via modulating Cd expression and transport genes. Pan et al. [218] showed the ability of ABA-producing B. subtilis to reduce Cd accumulation in Arabidopsis thaliana. Numerous bacterial species can produce gibberellic acid to reduce metal toxicity by lowering Cd absorption and lipid peroxidation, affecting hormonal balance and controlling the activities of proteases, catalase and peroxidase.

6.1.4. Antagonistic Role of PGPR

Antagonizing bacteria are crucial biocontrol agents in the rhizosphere zone, as they protect the plant from disease caused by pathogenic bacteria [219], as shown in Figure 2. Possible mechanisms for antagonistic activity may consist of antibiotics that inhibit pathogenic activity, the place for colonization, competition for nutrients, parasitism and mycophagy [220,221]. Biological control agents are more sustainable options for farming and are well accepted in many nations. Numerous biocontrol agents, including bacteria, fungi and actinomycetes, have been investigated for their potential effectiveness against various phytopathogens. Many have also been found to act on PGPR [222]. Different bacterial species, such as Pseudomonas, Bacillus, Klebsiella, Azospirillum spp., etc., have adapted to rhizosphere soil and have effectively defended plants against diseases [216,223]. Aspergillus flavus, often found in soil, has the potential to infect crops, and its aflatoxin negatively impacts the majority of crops [224]. A filamentous fungus called Alternaria alternata is responsible for causing diseases such as leafspot in Aloe vera and stem canker in tomato plants [225]. The fungus Fusarium oxysporum is a typical member of rhizosphere microbial communities, and in most cases, they are non-pathogenic. Pathogenic Fusarium, on the other hand, may attack plant roots and result in Fusarium wilt [226]. Various research has stated that certain soil bacteria, particularly PGPR, may be able to synthesize the cytokinins and gibberellins that control plant growth and development [227]. Using more efficient and genetically engineered bacteria (GEB) has shown a more significant requirement for removing Cd from polluted sites. Metal-binding peptides such as phytochelatins and metallothioneins have been found to improve HM binding. Phytochelatins are known to bind HMs, particularly Cd, with great affinity, by creating thiolate complexes. Phytochelatin coding genes have recently been cloned from plants and fungi and functionally expressed in Escherichia coli.
According to Paul and Bhakta [228], GEB, with both a high bioaccumulation capability and a strong affinity for the target metal, have demonstrated that they preferentially collect metal ions from multi-component pollution. The unique Cd transport mechanism and the metallothionein protein significantly enhance GEB’s ability to accumulate Cd2+ from multi-component metal-contaminated sites [229]. Dixit et al. [88] discovered that Mesorhizobium huakuii modified with an Arabidopsis thaliana gene coding for phytochelatins accumulated more Cd2+. In addition, according to Hou et al. [230], the amount of Cd accumulation in the recombinant Escherichia coli strain was about 25-fold greater than that in the control strain. Oliva-Arancibia et al. [231] found that Pseudomonas putida 06909 reduced Cd cellular toxicity. Recently, recombinant Caulobacter crescentus strain JS4022/p723-6H, expressing the RsaA-6His fusion protein, could remove 94.3–99.9% of Cd (II), whereas the control strain could only remove 11.4–37.0% [232]. Without calling the findings into doubt, there is significant promise for using recombinant technology to remove target HMs efficiently.

6.1.5. Siderophore Secretion

PGPR secrete a variety of Fe carriers known as siderophores to sustain metabolic activity. These siderophores may make complexes with HMs in the soil to increase their bioavailability via biochemical reactions to facilitate absorption through plants’ roots [216,233]. Low-molecular-weight compounds known as siderophores may bind ferric ions and make them available for microbial cells. PGPR bacteria can synthesize siderophores, which thrive and produce at their best in harsh environmental circumstances with low nutrition availability or under HM toxicity. For instance, Gazitúa et al. [234] identified six bacteria in the rhizomes of plants for culturing in the substrate under harsh conditions, which were isolated from the plant rhizospheres growing in floating tailings. The results demonstrated that bacterial strains improved the aboveground and belowground biomass of plants both by PGPR and metal-resistant bacterial strains. This is primarily because they may produce iron carriers and improve phosphate breakdown, which can lessen the detrimental effects of excessive HM concentrations and a deficiency of heavy elements [235]. Cadmium can also be transported through Fe transporters. However, two different systems are reported to uptake Fe from the soil. The first involves reduced Fe (II) uptake, whereas the second involves chelated Fe (III) uptake. Fe (III)-phytosiderophore complexes (Fe(III)-PS) assist transportation in barley and maize through yellow stripe 1/yellow stripe-like 1 (YS1/YSL1). It has been reported that the SnYSL3 and OsYSL2 proteins transport Cd in black nightshade (Solanum nigrum) and Oryza sativa, respectively [236].
Many plant-related bacteria make siderophores an essential PGPR feature in delivering iron to plants while protecting them against fungal infections. Cd2+ induces the synthesis of green-pigmented ‘pyoverdine’ siderophores in Pseudomonas. Mustard and pumpkin plants inoculated with siderophore synthesizing P. aeruginosa showed greater iron content in leaves in Cd-contaminated soil. Desferrioxamine E, Desferrioxamine B and coelichelin were identified as Cd-induced siderophores [237]. Streptomyces tendae synthesize these siderophores, aiding plant development, enabling soil metal solubilization and increasing Cd and Fe absorption in sunflowers (Helianthus annus L.). Siderophore production was demonstrated to boost iron absorption while decreasing Cd uptake in Streptomyces [238]. Pseudomonas putida enhanced plant growth by producing pyoverdines and reducing Pb and Cd absorption in mung beans. Plants inoculated with siderophore-producing bacteria were shown to either boost or limit Cd absorption, depending on the plant, bacterium and metal combination.

6.1.6. Volatile Organic Compounds

Low-molecular-weight molecules known as volatile organic compounds (VOCs) typically have less than 12 carbon atoms and a vapor pressure of at least 0.01 kPa at an ambient temperature. They may be linked to additional elements, including oxygen (O), nitrogen (N), bromine (Br), sulfur (S), fluorine (Fl) and chlorine (Cl) [239]. These substances are known as biogenic VOCs when living things synthesize them. Research has revealed that these chemicals are significant in various processes supporting plant growth, the induction of systemic resistance (ISR) and plant chemical signaling [240]. Bacterial VOCs such as acetoin and 2,3-butanediol may interact with plants, triggering plant defense and growth promotion mechanisms and assisting host plants in absorbing nutrients such as iron (Fe) and S. VOCs generated by Bacillus sp. B55 significantly enhanced the S feeding of Nicotiana attenuata. VOCs are important in most PGPR as bioprotectants through induced biopesticides, systemic resistance and phytostimulators. Such actions may help plants develop quicker, which is essential for the successful phytoextraction of Cd-contaminated soil. The volatile dimethylhexadecylamine (DMHDA), which is involved in promoting the growth and development of barrel clover (Medicago truncatula) seedlings, is one of these VOCs with advantageous activities in plants and is produced by the rhizobacterium Artrobacter agilis strain UMCV2, particularly under Fe deficit conditions [241]. DMHDA boosted the biomass, ferric reductase activity and chlorophyll content.
Additionally, DMHDA induces M. truncatula plant roots to release protons that aid in acidifying the rhizosphere, enabling Fe absorption under limited circumstances and increasing the amount of this element in plants inoculated with bacterium [242]. Later research revealed that the UMCV2 strain might live as an endophyte and colonize them in plant tissues. Additionally, there is evidence that VOCs may benefit the rhizosphere and boost plant immune systems. For instance, Bacillus subtilis synthesized volatile 2,3-butanediol, stimulating plant growth and physiological reactions. Additionally, 2,3-butanediol-exposed plant roots reacted by producing more root exudates. These findings imply that 2,3-butanediol induces the release of root exudates that regulate the activity of fungi and bacteria in the rhizosphere [243]. Rojas-Solis et al. [244] stated that two Bacillus strains synthesized volatile chemicals with synergistic actions to improve plant growth and reduce possible diseases such as Botrytis cinerea. Dimethyl disulfide (DMDS), a flammable substance, inhibited the growth of B. cinerea’s mycelium. Compounds such as ‘DMDS’ have yet to be assessed for their role in rhizosphere colonization. According to Sharma [245], Juncus maritimus synthesizes malonate and oxalate, which function as Cd-complexing agents, boosting Cd transport and solubility in soils. Pseudomonas sp. produces organic acids and enhances soil mineral nutrition and Cd availability by increasing host-mediated VOC secretion, dramatically increasing the shoot plant biomass of Solanum nigrum and Cd acquisition in aerial portions [245].

6.2. Indirect Mechanisms

6.2.1. Production of Antibiotics

To combat damage caused by phytopathogens, the primary approach of plant growth-promoting bacteria is the production of antibiotics. The biocontrol capabilities of bacterial strains such as Pseudomonas are mainly dependent on root colonization, the stimulation of plant systemic resistance and the production of antifungal antibiotics [246]. The synthesis of one or more antibiotics is often linked to the potential of rhizobacteria as biocontrol alternatives against plant pathogens. Cd resistance genes are primarily located on plasmids in bacteria, which allow them to resist Cd stress competitively. According to some research, Cd resistance on R plasmids relates to multiple antibiotic resistance [247]. The R plasmid is often found in clinically isolated human infections, such as Pseudomonas aeruginosa, Klebsiella pneumoniae, Staphylococcus aureus, etc. Cd resistance gene loci may be found on plasmids or on chromosomes [248]. The idea of antibiosis, or biocontrol based on the synthesis of chemicals that destroy or impede the development of target pathogens, has been characterized over the last 20 years. Antibiotics are a broad category of organic, low-molecular-weight substances that stop bacteria and other microbes from growing or functioning metabolically. According to Kenawy et al. [249], six groups of antibiotic substances, such as phloroglucinols, phenazines, pyrrolnitrin, cyclic lipopeptides, pyoluteorin and HCN, are most effectively linked to the biocontrol of root diseases.
All these substances are diffusible except for HCN, which is volatile. Lipopeptide biosurfactants synthesized by Bacillus and Pseudomonas strains have recently been utilized as a biocontrol agent because of their probable positive impact on competitive interactions with fungi, protozoa, oomycetes, bacteria, nematodes and plants. Pseudomonas synthesize antibiotics, i.e., HCN, 2,4-diacetyl phloroglucinol (DAPG), amphisin, oomycin A, phenazine, tropolone, tensin, pyrrolnitrin, pyoluteorin and cyclic lipopeptides, and Bacillus, Streptomyces and Stenotrophomonas spp. produce xanthobaccin, kanosamine, oligomycin A and zwittermicin, which have been recognized as antibiotics having antifungal, antiviral, phytotoxic, cytotoxic, antitumor, and antioxidant properties. It has been shown that the antibiotic ‘pyrrolnitrin’, produced by the P. fluorescens BL915 strain, defends cotton plants against Rhizoctonia solani in Cd-contaminated soil [250]. There have been several studies into efflux processes. For the extrusion of biocides, antibiotics, toxic metals and toxic substances from inside the bacterium into the environment, efflux pumps are composed of integrated membrane proteins. More than 20 potential efflux pumps proteins, either encoded by plasmids or chromosomes, have been described in S. aureus up to this point due to developments in genome analysis and bioinformatics [248]. Cd resistance has also been observed in drug-resistant efflux pumps [251].

6.2.2. Production of Exopolysaccharides

Since the discovery of the exopolysaccharide (EPS) adsorption potential by bacteria, several studies have already been published on a distinct range of microbial strains and EPSs with the required potential, i.e., the remediation of metal-contaminated soils [252]. Because EPS has a polysaccharide backbone, it is possible to structurally alter it by changing the polymeric length or adding a variety of side chains, non-carbohydrate substituents, functional units, linkages and other bonds in a combinatorial manner. The kind and proportion of the carbon source; abiotic stress elements, including temperature, pH and HMs; and the growth phase of the rhizobacterium, during which synthesis takes place, are all variables that affect the composition of the polysaccharide. The utilization of negatively charged EPSs (EPSs with large anionic functional groups), a feasible biosorbent, must be highlighted in strategies for the remediation of metals using rhizobacterial EPSs [253]. The polymer has a general negative charge due to the presence of numerous ionizable and active non-carbohydrate side chains and functional groups, including structural polysaccharides (fungi); acetamido (chitin group), amine, sulfhydryl and carboxyl groups in proteins; and hydroxyl and phosphate groups and phosphodiester in polysaccharides [254]. Extracellular heteropolysaccharides are polyanionic, unlike homopolysaccharides, because some functional groups interact with polysaccharide backbones [255]. Complexation, ion exchange and precipitation are a few mechanisms that result in immobilization and sorption. Xanthomonas campestris, Streptococci sp., Pseudomonas aeruginosa, Pseudomonas oleovorans, Sphingomonas paucimobilis, Pasteurella multocida and Azotobacter vinelandii are some of the recorded commercial rhizobacterial EPS strains that are capable of anionicity [256]. The association between positively charged metal ions and negatively charged EPSs and cell surfaces affects EPS-facilitated biosorption. In addition to causing or triggering biofilm formation in response to Cd contamination, Herminiimonas arsenicoxydans, a rhizobacterium with a Gram-negative phenotype, has also been reported to use EPSs to scavenge Cd when exposed to a 5 mM concentration [257]. This work demonstrates that, although rhizobacterial EPS production may not be increased in response to HM stress, produced EPS can still adsorb the metal.
Similarly, remediation of trace amounts of Pb and Cd has been studied using EPSs synthesized by Marinobacter sp. proteins and polysaccharides containing charged groups, such as carboxyl, amino, amide and hydroxyl groups, which often make up EPS. A synthesizing enzyme, Urease, first converts urea into ammonium ions in a microbially induced carbonate precipitation process. Generated ammonium ions then cause the pH to rise, decomposing the substrate urea into CO32−, which triggers the precipitation of carbonates of HM ions or coprecipitation with CaCO3 [258]. The primary mechanism for Cd elimination is the synthesis of Cd carbonates induced by bacterial action. However, many safer rhizobacteria are dispersed in the environment and waiting to be uncovered for the remediation of HM-contaminated sites.

6.2.3. Hydrogen Cyanide

Rhizobacteria can produce volatile chemicals such as HCN, nitric oxide and hydrogen sulfide, as shown in Figure 3. Among these, HCN is a volatile substance that is crucial to the rhizosphere’s biology. According to Dimkić et al. [246], the biocontrol mechanisms of bacteria, such as those seen in certain Pseudomonas strains, often rely on secreted bioactive substances that target the pathogen, such as exoenzymes, antibiotics or HCN. Brahim and Ouhdouch [259] stated that phenazine-1-carboxamide, a phenazine formed by P. chlororaphis PCL1391, inhibited F. oxysporum from causing tomato root rot. Because of their speedy and vicious colonization of plant roots, fluorescent Pseudomonas has been deemed an effective biological control agent against soil-borne plant pathogens. They revealed that two processes were occurring: one included competition for nutrients in the rhizosphere, preferably at colonization sites, and the other involved the production of compounds such as siderophores, HCN and antibiotics. Rhizobacterial strains may produce HCN and affect the development of seedling roots in a range of plants [229]. In a collection of more than 2000 isolates, they found that 32% of bacteria were cyanogenic, with HCN levels ranging from trace to >30 nmol/mg cellular protein. Pseudomonads were the most vulnerable to cyanogenesis, facilitated by adding glycine to the culture medium. Previous studies have hypothesized that microbial HCN prevents pathogenic fungi from growing on their mycelia by preventing the production of ATP-mediated cytochrome oxidase. Microbial HCN promotes plant development and Cd mobilization in addition to biocontrol. The growth and Cd accumulation efficiency of Sinapis alba L. is significantly increased when the Brevibacterium casei MH8a strain produces HCN, ACCD and IAA [260]. The VOC produced by PGPR promotes plant growth and development, increasing shoot biomass and enhancing resistance to plants against Cd stress [261].

6.2.4. Lipo-Chito-Oligosaccharides

Plants release flavonoids as secondary metabolites via their roots. It is well known that flavonoids have chemo-attractive properties that cause the expression of bacterial nod genes and the creation of lipo-chitooligosaccharides (LCO), which are crucial for developing nodules in roots. Flavonoids have selectivity for plant bacteria. Various chemicals may attract different bacterial species, making it feasible for a particular visitor to colonize. Because N fixation may encourage plant development in Cd-polluted soil, researchers are looking at N-fixing bacteria found in root nodules to find novel species that can be exploited in bioremediation. Rhizobia association with legume roots may improve Cd, Cu and Pb phytostabilization. Plants synthesize flavonoids to reduce Cd stress and increase antioxidant activity, and their release in the soil is a plant defense strategy. In particular, flavonoids may counter ROS within plant cells and can chelate metals such as Cd, Fe, Cu, Ni and Zn. The photo-microbiome community also affects the behaviors of each via by the methods of signal compounds [262]. Such types of signals are the holobiont’s hormones. For instance, riboflavin and lumichrome may act as microbes to plant signaling chemicals to promote plant development. Both substances have the potential to alter plant development significantly; lumichrome can hasten leaf emergence and leaf expansion. It may also enhance the plant’s height and total leaf area, further improving its biomass. Numerous plant species, including monocots and dicots, are affected by this [263]. It has been shown that microbe-to-plant signaling substances, including lipo-chito-oligosaccharides and thuricin, promote growth and development in various species of plants, especially under Cd stress [264]. Inoculating tomato seeds under Cd stress with a PGPR strain significantly boosted plant production of flavonoids and other phenolic compounds, accelerating antioxidant activity and reducing CD toxicity [265]. However, various safer LCOs, dispersed in the environment, are waiting to be uncovered for the remediation of Cd-contaminated sites.

7. Factors Affecting PGPR Bioremediation

The effectiveness of bioremediation is primarily influenced by site features, which can be further affected by environmental parameters such as water content, temperature, pH, nutrient availability, moisture content and pollutant bioavailability [266,267]. In addition, the bioremediation procedure is a complex system regulated and adjusted by multiple variables [30]. The bioavailability and biodegradation of Cd are influenced by interactions between contaminants, bacteria, the availability of nutrients and environmental variables [268]. The site location and its characteristics are the most significant factors impacting bioremediation. The degree and type of contaminants prevailing at the site determine the effectiveness of remediation [269]. These issues can be addressed through site assessment and prior inquiry before remediation.
Temperature plays a significant role in influencing the survival and growth of microorganisms [270]. By interfering with microbial metabolism, growth rate and the soil matrix, it plays a crucial part in microbe-assisted remediation by changing Cd physical and chemical states in polluted areas [271]. High temperatures, according to research by Javanbakht et al. [272], disrupt the metabolic activity of bacterial cells and impact the bioaccumulation process. Moreover, the physiological characteristics of microbes are influenced, which can speed up or slow down the remediation process. Temperature also affects how Cd ions interact with fungal membrane binding sites and influences the structure and stability of the fungal membrane by ionizing chemical moieties [273]. According to Jin et al. [185], Simplicillium chinense QD10 showed maximum biosorption efficiency for Cd and Pb at 30 °C, with 60.4 and 38.3%, respectively, which reduced significantly as temperature increased to 45 °C. Therefore, temperature regulation is critical for the success of the bioremediation process [274].
The metabolic activity of bacteria is influenced by pH, which can either accelerate or decelerate the elimination process. Bioremediation can be feasible over a wide pH range. However, according to Abatenh et al. [269], a pH range of 6.5 to 8.5 has the most significant potential for remediating most terrestrial and aquatic systems. The pH level affects the biosorption process by separating functional groups on fungal membranes and impacting the solubility and mobility of Cd [275]. The Exiguobacterium sp. exhibits a Cd biosorption capability that improves with an increased pH up to 7.0 and remains neutral when the pH is more significant than 7.0 [276]. pH and ionic strength can also impact microbial adsorption [274].
Similarly, nutrient availability, concentration and type are crucial for microbial activity and growth during bioremediation. Essential elements, such as carbon (C), nitrogen (N) and phosphorus (P), aid in the bacteria’s ability to manufacture the enzymes required to remediate Cd [21]. Lower nutrient availability impacts plants and microorganisms, eventually influencing bioremediation rate and efficiency. Adjusting the bacterial C:N:P ratio can increase the bioremediation efficacy by balancing essential nutrients such as N and P [274]. Providing adequate nutrients in an optimum environment promotes the metabolic activity of microorganisms, increasing the remediation rate [277,278]. According to reports, too much N in a contaminated medium leads to microbial inhibition [279].
Moreover, soil moisture levels may have adverse effects on microorganisms. Moisture affects the number and type of soluble materials and the pH and osmotic pressure of terrestrial and aquatic sites, affecting the efficiency of Cd remediation [274]. For bacteria to grow and metabolize efficiently, they typically require water activity levels between 0.9 and 1.0, with most bacteria thriving at the highest water activity values [280]. Therefore, the water content of polluted areas is a crucial variable that may influence the bioremediation rate. Recent research by Khodaverdiloo et al. [268] highlights that water scarcity, sodicity and salinity have recently been emphasized as significant factors affecting bioremediation effectiveness.
Unsuitable microorganisms or insufficient suitable microorganisms in contaminated locations also affect the bioremediation efficiency [270]. Because bioaccumulation is metabolically dependent and requires cellular energy for metal uptake, microbial biophysical processes also affect bioaccumulation. According to Srinath et al. [281], Vijayaraghavan and Yun [282] and Issazadeh et al. [283], it depends on different microbial traits, such as biochemical characteristics, genetic and physiological abilities, internal structure, cell surface qualities (including charge shifts) and surrounding environmental variables. Razmi et al. [22] discovered that various biological and chemical factors affect the effectiveness of phytoremediation. For plant-based remediation, root systems, which may have tap or fibrous roots depending on the depth of contaminants; aboveground biomass, which should not be preferred for livestock consumption; survival; and adaptations of plants, as well as plant growth, are essential considerations for choosing suitable plants [284]. Nonetheless, plant type has been identified as the primary determinant in Cd, Pb, Ni and Zn phytoremediation. Additionally, most fungal strains exhibit the highest biosorption efficiency under their ideal growing conditions [285].
The limited bioavailability of Cd in polluted soil has a significant impact on the effectiveness of bioremediation [7]. Several physicochemical processes, such as sorption, diffusion, desorption and dissolution, regulate the bioavailability of pollutants [12]. To manage this issue, several surfactants and chelating compounds increase Cd’s bioavailability for microbial breakdown and plant uptake [5]. Recent developments include the use of a variety of organic and inorganic chelating agents, including [S, S]-ethylenediamine succinic acid (EDDS), ethylenediamine tetra-acetic acid (EDTA), ethylenediamine-di-hydroxyphenyl acetic acid (EDDHA), nhydroxy-ethylenediaminetriacetic acid (HEDTA) and diethylene-triaminepentaacetic acid (DTPA). According to Sarwar et al. [286], these chelating agents have successfully demonstrated their ability to form a complex with HMs and boost bioavailability.

8. Recent Advancements (Genetic and Metabolic Engineering), (Membrane and Enzyme Technology), (Metagenomics Approaches) and (Nanoparticle Technology)

8.1. Membrane and Enzyme Technology

The remediation of HMs, i.e., Cd, using microbial enzymes, is significantly more efficient than other bio-remediation techniques, as they are ecologically friendly, affordable and innovative [287]. Research is being conducted on the ability and affordability of enzymes, including oxidoreductases, nitro-reductases and dehalogenases, to detoxify the harmful effects of Cd in agricultural soils [288]. Exudates secreted by plants are used as carbon and energy sources by PGPR in conjunction with plant roots to produce the enzyme 1-aminocyclopropane-1-carboxylase deaminase and IAA, which are used to break down metal pollutants [289,290]. These enzymes must be utilized under ideal temperatures and pH conditions to degrade pollutants effectively. The harmful effect of bioactive metals such as Cd can be alleviated using physicochemical characteristics due to different enzymes [291]. The development of the enzyme phytochelatin synthase and the production of proline can quickly bind to HMs at toxic concentrations and act as a free radical scavenger, respectively [292,293]. Phytochelatins (PCs) are thiol-rich and short-chain repetitions of low-molecular-weight peptides produced by phytochelatin synthase from glutathione to activate plant defense against metal stress [294].
Microbially induced precipitation (MIP) is a critical aspect of the biogeochemical cycle, wherein ions or chemicals react with metabolic products released by microbes, leading to the deposition of mineral particles [294]. One of the crucial features is microbially induced carbonate precipitation (MICP), which reduces Cd’s mobility by utilizing a carbonate-biomineralization microbe’s metabolic activity. Microorganisms release urease, which produces carbonate by reacting with urea [295]. To build Cd carbonate crystals, first, Cd ions are coupled with cell binding sites such as carboxyl, phosphate, cyanide, imidazole, and amino binding sites [296,297] and then are catalyzed by carbonate to form Cd carbonate (CdCO3) crystals via bacterial urea hydrolysis [298]. Cd carbonate reaches a saturation state in solution due to its poor solubility. Burns et al. [299] showed that functional groups on a microbe’s surface mediate microbial cell adhesion to mineral surfaces, and Huang et al. [300] found that microbial discharge mediates the deposition of mineral particles.
Consequently, the removal percentage outperformed individual biosorption. Therefore, MICP based on biomineralization is an ideal method for removing Cd contamination at 10–50 mg L−1. A promising method for the microbial remediation of metals is oxidation–reduction, in which enzyme reactions produce fewer toxic species by altering the valence state of polyvalent metal ions [301]. The activity of soil enzymes is sensitive to many forms of HM contamination, as HM pollution can reduce the activities of different enzyme-contaminated soil [302]. Their activity can be upregulated by following a proper enzyme-based remediation process [303,304]. However, significant research is needed to develop innovative methods that might be more precise and effective than those that are currently available [305]. Moreover, a significant gap exists between laboratory-level study, bioreactor/scaling-up applications and field research.

8.2. Genetic and Metabolic Engineering

Microorganisms can remediate HMs such as Cd by metabolizing them into innocuous by-products through co-metabolism [306]. Making new pathways, altering existing gene sequences and introducing single genes or operons into microorganisms are the most frequently utilized approaches [306]. Enzymatic bioremediation is a straightforward, efficient and eco-friendly method for microbe-assisted removal and destruction of persistent xenobiotics [307]. Genetic engineering for bioremediation seeks to alter plants, microbes and enzymes to make them viable tools for breaking down hazardous materials [308]. The bioremediation of Cd can be facilitated via these approaches [309]. Applying genetically modified bacteria and plants during the bioremediation of Cd-contaminated soils and other organic pollutants has become a promising technique [310]. Most methods include locating and inserting metal-uptake-related genes into plants and competent bacterial cells. n-alkane-degrading microbes possess specific genes, including xylE, polycyclic aromatic hydrocarbons (alkB, alkB1, alkB2 and alkM) and aromatic hydrocarbons (ndoB and nidA), which are used as markers to identify microbial-assisted biodegradations. Metabolic engineering emphasizes microbial-based enzymes involved in various degradation processes, such as oxidases, esterases, phenoloxidases, monooxygenases and oxidoreductases [311]. Several enzymes, including mixed-function oxidases (MFO), laccase, glutathione-S-transferase and cytochrome P-450, participate in the biodegradation of contaminants [312]. Enzyme immobilization also considerably increases enzyme activity, half-life and stability [307]. The stability of microbes must be maintained before their field application to use GEMs for bioremediation because the stability of recombinant plasmids inserted into the organism is responsible primarily for the catabolic activity of released GEM [313]. According to Dixit et al. [88], biosensors can estimate concentrations of HMs such as Cd in contaminated areas. However, variations in reaction times, detection thresholds, sensitivity, signal relaxation lengths and stability can limit the application of biosensors [314]. Modern genetics and omics techniques have made it possible to study the catabolism of contaminants using different microorganisms, allowing scientists to better understand the ecology, physiology and biochemistry of microorganisms that remediate Cd [313]. The drawback of alternative culture-independent techniques has been overcome by DNA microarrays, a high-throughput method that can identify several genes in a single test. The most popular gene array method for examining gene function is the GeoChip array [314]. Recombinant DNA technology is crucial for bioremediation because it helps to analyze, monitor and evaluate the specific technique. Nevertheless, it must be used responsibly and follow biosafety guidelines.

8.3. Metagenomics Approaches

Metagenomics is a rapidly expanding and new field of study. It is an environment-friendly and practicable technique for analyzing genetic material extracted directly from environmental samples. Therefore, in a niche ecosystem, metagenomics provides information through sequence and function-based research techniques about the communities of microbes of non-cultivable organisms. It comprises methods such as shotgun metagenomics, high-throughput sequencing and bioinformatics that aid in identifying, characterizing and screening prospective species [315,316]. Metagenomics is important for detecting and monitoring microbial activities [317]. Using metagenomics, DNA can be studied from ambient samples without isolating and cultivating the microbes. This method was first applied to find new microbes and microbial products and to sample the microbial diversity from various environmental niches to analyze their ability to eliminate organic pollutants [318]. Metagenomics is a fast-growing and emerging area of study. Opportunities to discover new ecosystems are made possible by the function-focused metagenomic technique, which promotes the discovery of new genes and provides genetic analysis [319]. Sequence interpretation, regarded as essential for feature estimates, is the foundation of the sequence-based technique. The bioremediation technique also considers microbial diversity and particular genes identified by metagenomic research with the potential to act as pollution biomarkers with bioactive substances and enzymes determined using metagenomics techniques. Additionally, metagenomics approaches are useful in identifying the enzymes, metabolites and bioactive substances produced by bacteria, all of which play a vital role in water treatment. Metagenomics studies analyze strategies to uncover new genes and bacteria to mediate the detoxification of organic pollutants, including Cd [320,321].
In Cd-contaminated soils, the relative abundance of Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways increased with an enriched metabolism, biosynthesis and the degradation of different fatty acids and nucleotides, which was connected to microorganisms’ induced tolerance of Cd [316]. In addition to investigating the studied soil microbial population, their roles and genes for diverse applications, soil metagenomics can be envisioned as a technique for promoting plant growth in Cd-contaminated soils [322]. Metagenomic sequencing depends upon several strategies that may explore new ways to identify bioremediation contaminants. Pyrosequencing, single-molecule sequencing, ligation sequencing and reverse terminator sequencing are examples of next-generation sequencing techniques made possible by technical drift and enable high-throughput readings in a shorter time. Metagenomics is one of the best breakthroughs now available, and it has frequently been used in building better ecosystems through bioremediation. The Simple Metagenomics Analysis Shell for the community of microbes is the only pipeline based on metagenomic investigations to share the ideas and events of architecture with Smash-Cell [323]. This allows comparing multi-metagenome compositions, making functional graphical representations of these tests and estimating metagenome quantitative phylogenetic and functional compositions. The most common method is Meta Genome Analyzer (MEGAN), which uses sequence data for a practical analysis across various bioinformatics applications [324]. This approach is perfect for evaluating metagenomics and metaproteomics using interactive functional and taxonomical data. A sophisticated metagenome tool called the Metagenome Subsystems Technology Rapid Annotation (MGRAST) provides microbial communities with quantitative insights based on sequencing data [325]. The Integrated Microbial Genomes and Metagenomes (IMG/M) system supported the creation of genome and metagenome databases of different microbes.

8.4. Nanoparticle Technology

Nanotechnology has emerged as an attractive field to synthesize and modify innovative and nanostructured materials for several purposes, including the remediation of the environment from organic and inorganic pollutants, known as nano bioremediation. The removal of Cd pollution could be more effective with the available technology. However, novel physical and chemical characteristics of nanoparticles (NPs) can therefore be used to remediate Cd efficiently [326]. With their unique chemical, structural and multifunctional features, morphologies, various compositions and high mechanical strength, NP-based innovative materials and nanochemistry methods are promising for degradation, adsorption and catalysis applications in metal pollution [258,327,328]. Various bulk materials can be used to create NPs, and particles’ size, shape and chemical makeup all influence the behavior and composition of NPs [329]. They have paved the way for low-cost and effective ways to limit the toxic effects of environmental contaminants [330]. Nanotechnology usage has become a fiercely debated topic due to the need for more thorough investigations and knowledge of how NPs interact with other environmental elements. Silver NPs impact the microbial populations in the root zone, aiding in the removal of Cd [331]. The use of NPs improves the phytoremediation of HMs, including Cd, by upregulating the growth of plant roots and shoots [332]. Engineered nanoparticles (ENPs), a new class of environmental chemicals, have shown considerable effects on the fate and transport of coexisting ecological contaminants and the effectiveness of plant uptake [333].
Most studies investigating the effects of ENPs on plant metal uptake have focused on metallic ENPs. This makes sense, given that some metallic ENPs tend to dissolve in the root zone, leading to dissolved ions that may compete with metal ions for plant uptake, altering the course of HM uptake [334]. For example, the uptake of Cd was significantly reduced by the application of citrate-coated magnetic NPs in wheat, which ameliorated Cd toxicity. In addition, the uptake of Cd was reduced considerably by the application of TiO2NPs in rice [335]. The authors also examined how cerium oxide (CeO2) NPs affected the amount of co-occurring Cd that accumulated in soybeans (Glycine max) and found that CeO2NPs considerably altered the amount of Cd that accumulated, hindering its concentration in soybean shoots [336]. However, there is still limited research on how ENPs affect the uptake of coexisting Cd by plants. More mechanical knowledge about these interactions in the plant rhizosphere must be determined. Further research is necessary to better understand the complex chemical and biological processes that may affect plant Cd uptake in the presence of ENPs.

9. Challenges and Future Prospects

Various bioremediation techniques have been successfully in restoring contaminated locations exposed to Cd. However, when using bioremediation techniques, there are a few crucial considerations to remember. Before suggesting bioremediation, there is a requirement for the regular study and assessment of the amount of Cd and other pollutant concentrations in polluted locations. Choosing suitable microorganisms and plant species is difficult for sites with several metals such as Cd and other organic contaminants. Second, hazardous metals and metalloids such as Si, Hg and As at the site may volatilize into the atmosphere during plant-based bioremediation, posing a risk to living things. Third, edible plants used for bioremediation could be eaten by animals or insects, which could further contaminate the food chain and eventually reach people, posing a significant health concern. Therefore, phytoremediator plants that are neither edible nor palatable should be preferred. Sufficient precautions must be taken during the cultivation and harvesting of edible plants to prevent further issues. In situ, phytoremediation is more challenging when Cd is present more deeply in the soil, where plant roots cannot reach.
More research, assessments and inquiries are needed to improve our knowledge and comprehension of optimum management techniques for the practical bioremediation of Cd. Mechanisms, metabolites and new approaches/methods need to be clarified in a futuristic manner. Using hyperaccumulator plants to remove Cd from polluted soil effectively requires unique techniques for the continued development of plant-based bioremediation to be easy and effective. This can be performed in two different ways: first, by discovering and validating new species of diverse hyperaccumulator plants, and second, by creating hyperaccumulator plants through genetic engineering. Moreover, we can consider deep-rooted hyperaccumulator plants, such as woody plants or trees such as Populus canescens, Schima superba, Rinorea bengalensis and Pycnandra acuminata, which have higher growth rates, biomass and translocation rates, as well as more tolerant plant species.
Biotechnological interventions, such as genetic engineering, can increase a known metabolic pathway’s transfer and biodegradation rates and enhance the accumulation of Cd or the degradation of recalcitrant compounds by introducing a completely new metabolic pathway into the microbe. Moreover, it may be possible to remediate Cd from the soil by overexpressing foreign genes into a non-tolerant plant with a larger biomass. Modifying microbial niches that increase resistance against Cd pollution will be made easier with the help of the cutting-edge method of the holo-genomics analysis of plant microorganisms. There is a need to design suitable amendments for multi-metal-contaminated and multi-stress environmental situations to improve the survival of suitable plant species. Even though several organic and inorganic amendments and metal chelators are available, additional research is required to identify more effective and environmentally acceptable amendments that may be used to treat soil exposed to several metal contaminants and stresses. The effectiveness and dependability of bioremediation depend on the collaboration and contribution of researchers, scientists, policymakers, governments, industrial sectors and individuals.

10. Conclusions

High levels of Cd are being released into the environment due to human activity, directly and indirectly affecting all living things. Reports have indicated that contaminated soil contains numerous HMs simultaneously, and conventional detoxification techniques are less effective than the bioremediation procedure. It has been established that bioremediation procedures are significantly cheaper than other physicochemical remediation methods. Numerous bacterial and fungal strains have recently been isolated and characterized from metal-contaminated and mining-abandoned soils. Many strains of Bacillus spp., Pseudomonas spp., Aspergillus spp. and Penicillium spp. are present and exhibit excellent metal resistance in soil, especially against Cd. A variety of contaminated areas throughout the world are currently using bioremediation, with variable degrees of effectiveness. The main worry for the considerable yield of bioremediation is the inclusion of appropriate supplements and improving environmental conditions. To address the issue, adding organic matter and a group of microorganisms can increase microbial metabolic activity and enhance the potential for bioremediation. Moreover, more research is still needed to identify the best microbes and hyperaccumulator plants with a high tolerance for multi-metal-contaminated and multi-stress environmental locations and to accumulate several metals simultaneously. More emphasis will be paid to plant–microbe-based bioremediation strategies to find novel plant–microbe pairs with high metal removal effectiveness and to create an environment that is conducive to other microbial strains. This will indirectly improve the health of the soil. Additionally, more research is required on combining nanomaterials, biochar and microorganisms with bioremediation.

Author Contributions

Conceptualization, U.Z. and F.U.H.; methodology, M.F.M. and M.K.; software, W.M.-U.-D.; validation, M.S. and W.M.-U.-D.; formal analysis, M.A.; investigation, U.Z.; resources, F.U.H.; data curation, U.Z., F.U.H., M.F.M., M.S. and W.M.-U.-D.; writing—original draft preparation, U.Z., F.U.H., M.F.M., M.S. and W.M.-U.-D.; writing—review and editing, M.I. and Z.A.; visualization, U.Z., F.U.H. and B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zulfiqar, U.; Farooq, M.; Hussain, S.; Maqsood, M.; Hussain, M.; Ishfaq, M.; Ahmad, M.; Anjum, M.Z. Lead toxicity in plants: Impacts and remediation. J. Environ. Manag. 2019, 250, 109557. [Google Scholar] [CrossRef] [PubMed]
  2. Roy, A.; Sharma, A.; Yadav, S.; Jule, L.T.; Krishnaraj, R. Nanomaterials for remediation of environmental pollutants. Bioinorg. Chem. Appl. 2021, 2021, 1764647. [Google Scholar] [CrossRef] [PubMed]
  3. Fergusson, J.E. (Ed.) The Heavy Elements: Chemistry, Environmental Impact and Health Effects; Pergamon Press: Oxford, UK, 1990. [Google Scholar]
  4. Haider, F.U.; Wang, X.; Zulfiqar, U.; Farooq, M.; Hussain, S.; Mehmood, T.; Naveed, M.; Li, Y.; Liqun, C.; Saeed, Q.; et al. Biochar application for remediation of organic toxic pollutants in contaminated soils; An update. Ecotoxicol. Environ. Saf. 2022, 248, 114322. [Google Scholar] [CrossRef] [PubMed]
  5. Zulfiqar, U.; Haider, F.U.; Ahmad, M.; Hussain, D.; Maqsood, M.F.; Ishfaq, M.; Shahzad, B.; Waqas, M.M.; Ali, B.; Tayyab, M.N.; et al. Chromium Toxicity, Speciation, and Remediation Strategies in Soil-Plant Interface: A Critical Review. Front. Plant Sci. 2023, 13, 5468. [Google Scholar] [CrossRef]
  6. Zulfiqar, U.; Ayub, A.; Hussain, S.; Waraich, E.A.; El-Esawi, M.A.; Ishfaq, M.; Ahmad, M.; Ali, N.; Maqsood, M.F. Cadmium toxicity in plants: Recent progress on morpho-physiological effects and remediation strategies. J. Soil Sci. Plant Nutr. 2021, 22, 212–269. [Google Scholar] [CrossRef]
  7. Qianqian, M.; Haider, F.U.; Farooq, M.; Adeel, M.; Shakoor, N.; Jun, W.; Jiaying, X.; Wang, X.W.; Panjun, L.; Cai, L. Selenium treated Foliage and biochar treated soil for improved lettuce (Lactuca sativa L.) growth in Cd-polluted soil. J. Cleaner Prod. 2022, 335, 130267. [Google Scholar] [CrossRef]
  8. Rahimzadeh, M.R.; Rahimzadeh, M.R.; Kazemi, S.; Moghadamnia, A.A. Cadmium toxicity and treatment: An update. Casp. J. Intern. Med. 2017, 8, 135. [Google Scholar]
  9. Gul, I.; Manzoor, M.; Hashim, N.; Shah, G.M.; Waani, S.P.T.; Shahid, M.; Antoniadis, V.; Rinklebe, J.; Arshad, M. Challenges in microbially and chelate-assisted phytoextraction of cadmium and lead—A review. Environ. Pollut. 2021, 287, 117667. [Google Scholar] [CrossRef]
  10. Liu, L.; Li, J.W.; Yue, F.X.; Yan, X.W.; Wang, F.Y.; Bloszies, S.; Wang, Y. Effects of arbuscular mycorrhizal inoculation and biochar amendment on maize growth, cadmium uptake and soil cadmium speciation in Cd-contaminated soil. Chemosphere 2018, 194, 495–503. [Google Scholar] [CrossRef]
  11. Younas, F.; Younas, S.; Bibi, I.; Farooqi, Z.U.R.; Hameed, M.A.; Mohy-Ud-Din, W.; Shehzad, M.T.; Hussain, M.M.; Shakil, Q.; Shahid, M. A critical review on the separation of heavy metal (loid) s from the contaminated water using various agricultural wastes. Int. J. Phytoremed. 2023, 1–20. [Google Scholar] [CrossRef]
  12. Haider, F.U.; Liqun, C.; Coulter, J.A.; Cheema, S.A.; Wu, J.; Zhang, R.; Wenjun, M.; Farooq, M. Cadmium toxicity in plants: Impacts and remediation strategies. Ecotoxicol. Environ. Saf. 2021, 211, 111887. [Google Scholar] [CrossRef]
  13. Abbas, T.; Rizwan, M.; Ali, S.; Adrees, M.; Zia-ur-Rehman, M.; Qayyum, M.F.; Ok, Y.S.; Murtaza, G. Effect of Biochar on Alleviation of Cadmium Toxicity in Wheat (Triticum aestivum L.) Grown on Cd-Contaminated Saline Soil. Environ. Sci. Pollut. Res. 2018, 25, 25668–25680. [Google Scholar] [CrossRef]
  14. Chen, Q.; Lu, X.; Guo, X.; Pan, Y.; Yu, B.; Tang, Z.; Guo, Q. Differential Responses to Cd Stress Induced by Exogenous Application of Cu, Zn or Ca in the Medicinal Plant Catharanthus roseus. Ecotoxicol. Environ. Saf. 2018, 157, 266–275. [Google Scholar] [CrossRef] [PubMed]
  15. Zulfiqar, U.; Jiang, W.; Xiukang, W.; Hussain, S.; Ahmad, M.; Maqsood, M.F.; Ali, N.; Ishfaq, M.; Kaleem, M.; Haider, F.U.; et al. Cadmium Phytotoxicity, Tolerance, and Advanced Remediation Approaches in Agricultural Soils; A Comprehensive Review. Front. Plant Sci. 2022, 13, 773815. [Google Scholar] [CrossRef] [PubMed]
  16. Dermont, G.; Bergeron, M.; Mercier, G.; Richer-Laflèche, M. Soil washing for metal removal: A review of physical/chemical technologies and field applications. J. Hazard. Mater. 2008, 152, 1–31. [Google Scholar] [CrossRef] [PubMed]
  17. Xu, J.; Liu, C.; Hsu, P.C.; Zhao, J.; Wu, T.; Tang, J.; Liu, K.; Cui, Y. Remediation of heavy metal contaminated soil by asymmetrical alternating current electrochemistry. Nat. Commun. 2019, 10, 2440. [Google Scholar] [CrossRef]
  18. Kumar, A.; Subrahmanyam, G.; Mondal, R.; Cabral-Pinto, M.M.S.; Shabnam, A.A.; Jigyasu, D.K.; Malyan, S.K.; Kishor-Fagodiya, R.; Khan, S.A.; Kumar, A.; et al. Bio-remediation approaches for alleviation of cadmium contamination in natural resources. Chemosphere 2021, 268, 128855. [Google Scholar] [CrossRef]
  19. Younas, F.; Mustafa, A.; Farooqi, Z.U.R.; Wang, X.; Younas, S.; Mohy-Ud-Din, W.; Ashir Hameed, M.; Mohsin Abrar, M.; Maitlo, A.A.; Noreen, S. Current and emerging adsorbent technologies for wastewater treatment: Trends, limitations, and environmental implications. Water 2021, 13, 215. [Google Scholar] [CrossRef]
  20. Ahluwalia, S.S.; Goyal, D. Microbial and plant derived biomass for removal of heavy metals from wastewater. Bioresour. Technol. 2007, 98, 2243–2257. [Google Scholar] [CrossRef]
  21. Saha, L.; Tiwari, J.; Bauddh, K.; Ma, Y. Recent developments in microbe–plant-based bioremediation for tackling heavy metal-polluted soils. Front. Microbiol. 2021, 12, 731723. [Google Scholar] [CrossRef]
  22. Ortiz, A.; Sansinenea, E. The role of beneficial microorganisms in soil quality and plant health. Sustainability 2022, 14, 5358. [Google Scholar] [CrossRef]
  23. Mohy-Ud-Din, W.; Akhtar, M.J.; Bashir, S.; Asghar, H.N.; Nawaz, M.F.; Chen, F. Isolation of Glyphosate-Resistant Bacterial Strains to Improve the Growth of Maize and Degrade Glyphosate under Axenic Condition. Agriculture 2023, 13, 886. [Google Scholar] [CrossRef]
  24. Kuiper, I.; Bloemberg, G.V.; Lugtenberg, B.J.J. Selection of a plant-bacterium pair as a novel tool for rhizostimulation of polycyclic aromatic hydrocarbon degrading bacteria. Mol. Plant-Microbe Interact. 2001, 14, 1197–1205. [Google Scholar] [CrossRef]
  25. Radhakrishnan, R.; Hashem, A.; Abd-Allah, E.F. Bacillus: A Biological Tool for Crop Improvement through Bio-Molecular Changes in Adverse Environments. Front. Physiol. 2017, 8, 667. [Google Scholar] [CrossRef]
  26. Imam, S.S.A.; Rajpoot, I.K.; Gajjar, B.; Sachdeva, A. Comparative Study of Heavy Metal Bioremediation in Soil by Bacillus subtilis and Saccharomyces Cerevisiae. Int. J. Sci. Technol. 2016, 9, 106911. [Google Scholar] [CrossRef]
  27. Bakiyaraj, R.; Baskaran, L.; Chidambaram, A.L.; Mahakavi, T.; Santhoshkumar, M. Bioremediation of Chromium by Bacillus subtilis and Pseudomonas Aeruginosa. Int. J. Curr. Microbiol. Appl. Sci. 2014, 3, 715–719. [Google Scholar]
  28. Sabae, S.Z.; Hazaa, M.; Hallim, S.A.; Awny, N.M.; Daboor, S.M. Bioremediation of Zn+2, Cu+2 and Fe+2 Using Bacillus subtilis D215 and Pseudomonas Putida Biovar AD225. Biosci. Res. 2006, 3, 189–204. [Google Scholar]
  29. Sakthipriya, N.; Doble, M.; Sangwai, J.S. Bioremediation of Coastal and Marine Pollution due to Crude Oil Using a Microorganism Bacillus subtilis. Procedia Eng. 2015, 116, 213–220. [Google Scholar] [CrossRef]
  30. Leong, Y.K.; Chang, J.S. Bioremediation of Heavy Metals Using Microalgae: Recent Advances and Mechanisms. Bioresour. Technol. 2020, 303, 122886. [Google Scholar] [CrossRef]
  31. Saeed, M.U.; Hussain, N.; Sumrin, A.; Shahbaz, A.; Noor, S.; Bilal, M.; Aleya, L.; Iqbal, H.M. Microbial Bioremediation Strategies with Wastewater Treatment Potentialities—A Review. Sci. Total Environ. 2022, 818, 151754. [Google Scholar] [CrossRef]
  32. Cimboláková, I.; Uher, I.; Laktičová, K.V.; Vargová, M.; Kimáková, T.; Papajová, I. Heavy Metals and the Environment. In Environmental Factors Affecting Human Health; IntechOpen: London, UK, 2019; p. 29. [Google Scholar]
  33. Kubier, A.; Wilkin, R.T.; Pichler, T. Cadmium in Soils and Groundwater: A Review. Appl. Geochem. 2019, 108, 104388. [Google Scholar] [CrossRef] [PubMed]
  34. Khan, Z.; Elahi, A.; Bukhari, D.A.; Rehman, A. Cadmium Sources, Toxicity, Resistance and Removal by Microorganisms—A Potential Strategy for Cadmium Eradication. J. Saudi Chem. Soc. 2022, 26, 101569. [Google Scholar] [CrossRef]
  35. Hussain, M.M.; Mohy-Ud-Din, W.; Younas, F.; Niazi, N.K.; Bibi, I.; Yang, X.; Rasheed, F.; Farooqi, Z.U.R. Biochar: A game changer for sustainable agriculture. Sustain. Agric. Tech. Progress. Transit. 2022, 143–157. [Google Scholar]
  36. Huybrechts, M.; Cuypers, A.; Deckers, J.; Iven, V.; Vandionant, S.; Jozefczak, M.; Hendrix, S. Cadmium and Plant Development: An Agony from Seed to Seed. Int. J. Mol. Sci. 2019, 20, 3971. [Google Scholar] [CrossRef] [PubMed]
  37. Banerjee, A.; Roychoudhury, A. Gibberellic Acid-Priming Promotes Fluoride Tolerance in a Susceptible Indica Rice Cultivar by Regulating the Antioxidant and Phytohormone Homeostasis. J. Plant Growth Regul. 2020, 39, 1476–1487. [Google Scholar] [CrossRef]
  38. Reed, R.C.; Bradford, K.J.; Khanday, I. Seed Germination and Vigor: Ensuring Crop Sustainability in a Changing Climate. Heredity 2022, 128, 450–459. [Google Scholar] [CrossRef]
  39. Shu, K.; Liu, X.-D.; Xie, Q.; He, Z.-H. Two Faces of One Seed: Hormonal Regulation of Dormancy and Germination. Mol. Plant 2016, 9, 34–45. [Google Scholar] [CrossRef]
  40. Seneviratne, M.; Rajakaruna, N.; Rizwan, M.; Madawala, H.; Ok, Y.S.; Vithanage, M. Heavy Metal-Induced Oxidative Stress on Seed Germination and Seedling Development: A Critical Review. Environ. Geochem. Health 2019, 41, 1813–1831. [Google Scholar] [CrossRef]
  41. El Rasafi, T.; Oukarroum, A.; Haddioui, A.; Song, H.; Kwon, E.E.; Bolan, N.; Tack, F.M.G.; Sebastian, A.; Prasad, M.; Rinklebe, J. Cadmium Stress in Plants: A Critical Review of the Effects, Mechanisms, and Tolerance Strategies. Crit. Rev. Environ. Sci. Technol. 2022, 52, 675–726. [Google Scholar] [CrossRef]
  42. Aslam, M.M.; Okal, E.J.; Waseem, M. Cadmium Toxicity Impacts Plant Growth and Plant Remediation Strategies. Plants 2022, 11, 111887. [Google Scholar] [CrossRef]
  43. Farooqi, Z.U.R.; Khursheed, M.M.; Gulzar, N.; Akram, M.A.; Hussain, M.M.; Qadeer, A.; Mohy-Ud-Din, W.; Younas, S. The Risk of Inorganic Environmental Pollution to Humans. In Nanotechnology for Environmental Pollution Decontamination Tools, Methods, and Approaches for Detection and Remediation; Apple Academic Press: Burlington, ON, Canada, 2022; Volume 39. [Google Scholar]
  44. Seifikalhor, M.; Hassani, S.B.; Aliniaeifard, S. Seed Priming by Cyanobacteria (Spirulina platensis) and Salep Gum Enhances Tolerance of Maize Plant against Cadmium Toxicity. J. Plant Growth Regul. 2020, 39, 1009–1021. [Google Scholar] [CrossRef]
  45. Sardar, R.; Ahmed, S.; Shah, A.A.; Yasin, N.A. Selenium Nanoparticles Reduced Cadmium Uptake, Regulated Nutritional Homeostasis and Antioxidative System in Coriandrum sativum Grown in Cadmium Toxic Conditions. Chemosphere 2022, 287, 132332. [Google Scholar] [CrossRef] [PubMed]
  46. Taie, H.A.; Seif El-Yazal, M.A.; Ahmed, S.M.; Rady, M.M. Polyamines Modulate Growth, Antioxidant Activity, and Genomic DNA in Heavy Metal–Stressed Wheat Plant. Environ. Sci. Pollut. Res. 2019, 26, 22338–22350. [Google Scholar] [CrossRef]
  47. Bailly, C. The Signalling Role of ROS in the Regulation of Seed Germination and Dormancy. Biochem. J. 2019, 476, 3019–3032. [Google Scholar] [CrossRef] [PubMed]
  48. Larbi, S.; Ben Kaab, S.; Teixeira da Silva, J.; Bettaieb Ben Kaab, L. Cadmium and Copper Stresses Affect Germination and Enzymatic Activities in Chickpea (Cicer arietinum L.). Agrochimica 2020, 2, 191–203. [Google Scholar] [CrossRef]
  49. Zhu, T.; Li, L.; Duan, Q.; Liu, X.; Chen, M. Progress in Our Understanding of Plant Responses to the Stress of Heavy Metal Cadmium. Plant Signal. Behav. 2021, 16, 1836884. [Google Scholar] [CrossRef]
  50. Witkowska, D.; Słowik, J.; Chilicka, K. Heavy Metals and Human Health: Possible Exposure Pathways and the Competition for Protein Binding Sites. Molecules 2021, 26, 6060. [Google Scholar] [CrossRef]
  51. Shanmugaraj, B.M.; Malla, A.; Ramalingam, S. Cadmium Stress and Toxicity in Plants: An Overview. In Cadmium Toxicity and Tolerance in Plants; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–17. [Google Scholar]
  52. Kaur, H.; Hussain, S.J. Cadmium: Bioavailability in Soils and Phytotoxicity. In Sustainable Solutions for Elemental Deficiency and Excess in Crop Plants; Springer: Singapore, 2020; pp. 351–391. [Google Scholar]
  53. Riaz, U.; Aslam, A.; uz Zaman, Q.; Javeid, S.; Gul, R.; Iqbal, S.; Javid, S.; Murtaza, G.; Jamil, M. Cadmium Contamination, Bioavailability, Uptake Mechanism and Remediation Strategies in Soil-Plant-Environment System: A Critical Review. Curr. Anal. Chem. 2021, 17, 49–60. [Google Scholar] [CrossRef]
  54. Ismael, M.A.; Elyamine, A.M.; Moussa, M.G.; Cai, M.; Zhao, X.; Hu, C. Cadmium in Plants: Uptake, Toxicity, and its Interactions with Selenium Fertilizers. Metallomics 2019, 11, 255–277. [Google Scholar] [CrossRef]
  55. Wani, K.I.; Zehra, A.; Choudhary, S.; Naeem, M.; Aftab, T. Cadmium, a Nonessential Heavy Metal: Uptake, Translocation, Signaling, Detoxification, and Impact on Amino Acid Metabolism. In Plant Metal and Metalloid Transporters; Springer: Singapore, 2022. [Google Scholar]
  56. Zhu, G.; Xiao, H.; Guo, Q.; Zhang, Z.; Zhao, J.; Yang, D. Effects of Cadmium Stress on Growth and Amino Acid Metabolism in Two Compositae Plants. Ecotoxicol. Environ. Saf. 2018, 158, 300–308. [Google Scholar] [CrossRef]
  57. Sakouhi, L.; Kharbech, O.; Massoud, M.B.; Munemasa, S.; Murata, Y.; Chaoui, A. Oxalic Acid Mitigates Cadmium Toxicity in Cicer arietinum L. Germinating Seeds by Maintaining the Cellular Redox Homeostasis. J. Plant Growth Regul. 2022, 41, 697–709. [Google Scholar] [CrossRef]
  58. Kaleem, M.; Shabir, F.; Hussain, I.; Hameed, M.; Ahmad, M.S.A.; Mehmood, A.; Ashfaq, W.; Riaz, S.; Afzaal, Z.; Maqsood, M.F. Alleviation of Cadmium Toxicity in Zea mays L. through Up-Regulation of Growth, Antioxidant Defense System and Organic Osmolytes under Calcium Supplementation. PLoS ONE 2022, 17, e0269162. [Google Scholar] [CrossRef] [PubMed]
  59. Alzahrani, Y.; Rady, M.M. Compared to Antioxidants and Polyamines, the Role of Maize Grain-Derived Organic Biostimulants in Improving Cadmium Tolerance in Wheat Plants. Ecotoxicol. Environ. Saf. 2019, 182, 109378. [Google Scholar] [CrossRef]
  60. Kudoyarova, G.R.; Dodd, I.C.; Veselov, D.S.; Rothwell, S.A.; Veselov, Y. Common and Specific Responses to Availability of Mineral Nutrients and Water. J. Exp. Bot. 2015, 66, 2133–2144. [Google Scholar] [CrossRef]
  61. Kul, R.; Ekinci, M.; Turan, M.; Ors, S.; Yildirim, E. How Abiotic Stress Conditions Affects Plant Roots. In Plant Roots; IntechOpen: London, UK, 2020; pp. 6–10. [Google Scholar]
  62. Ghori, N.-H.; Ghori, T.; Hayat, M.; Imadi, S.; Gul, A.; Altay, V.; Ozturk, M. Heavy Metal Stress and Responses in Plants. Int. J. Environ. Sci. Technol. 2019, 16, 1807–1828. [Google Scholar] [CrossRef]
  63. Zhao, H.; Guan, J.; Liang, Q.; Zhang, X.; Hu, H.; Zhang, J. Effects of Cadmium Stress on Growth and Physiological Characteristics of Sassafras Seedlings. Sci. Rep. 2021, 11, 9913. [Google Scholar] [CrossRef]
  64. Carvalho, M.E.; Castro, P.R.; Kozak, M.; Azevedo, R.A. The Sweet Side of Misbalanced Nutrients in Cadmium–Stressed Plants. Ann. Appl. Biol. 2020, 176, 275–284. [Google Scholar] [CrossRef]
  65. Guo, J.; Chen, T.; Zheng, G.; Yang, J.; Qian, T.; Liu, X.; Meng, X.; Li, Y. Cadmium Accumulation Responses in Hylotelephium spectabile: The Role of Photosynthetic Characteristics under Different Nitrogen, Moisture, and Light Conditions. Chemosphere 2023, 319, 138019. [Google Scholar] [CrossRef]
  66. Zhu, Z.; Huang, Y.; Wu, X.; Liu, Z.; Zou, J.; Chen, Y.; Su, N.; Cui, J. Increased Antioxidative Capacity and Decreased Cadmium Uptake Contribute to Hemin-Induced Alleviation of Cadmium Toxicity in Chinese Cabbage Seedlings. Ecotoxicol. Environ. Saf. 2019, 177, 47–57. [Google Scholar] [CrossRef]
  67. Giannakoula, A.; Therios, I.; Chatzissavvidis, C. Effect of Lead and Copper on Photosynthetic Apparatus in Citrus (Citrus aurantium L.) Plants. The Role of Antioxidants in Oxidative Damage as a Response to Heavy Metal Stress. Plants 2021, 10, 155. [Google Scholar] [CrossRef]
  68. Moustakas, M.; Moustaka, J.; Sperdouli, I. Hormesis in Photosystem II: A Mechanistic Understanding. Curr. Opin. Toxicol. 2022, 29, 57–64. [Google Scholar] [CrossRef]
  69. Singha, K.T.; Sebastian, A.; Prasad, M.N.V. Iron Plaque Formation in the Roots of Pistia stratiotes L.: Importance in Phytoremediation of Cadmium. Int. J. Phytoremediat. 2019, 21, 120–128. [Google Scholar] [CrossRef] [PubMed]
  70. Sharma, A.; Kumar, V.; Shahzad, B.; Ramakrishnan, M.; Singh Sidhu, G.P.; Bali, A.S.; Handa, N.; Kapoor, D.; Yadav, P.; Khanna, K. Photosynthetic Response of Plants under Different Abiotic Stresses: A Review. J. Plant Growth Regul. 2020, 39, 509–531. [Google Scholar] [CrossRef]
  71. Rathi, S.; Mittal, N.; Kumar, D. Photosynthetic Response of Plants Against Heavy Metals. In Heavy Metals in Plants: Physiological to Molecular Approach; CRC Press: Boca Raton, FL, USA, 2022; pp. 1–23. [Google Scholar]
  72. Mir, A.R.; Pichtel, J.; Hayat, S. Copper: Uptake, Toxicity and Tolerance in Plants and Management of Cu-Contaminated Soil. Biometals 2021, 34, 737–759. [Google Scholar] [CrossRef] [PubMed]
  73. Jaouani, K.; Karmous, I.; Ostrowski, M.; El Ferjani, E.; Jakubowska, A.; Chaoui, A. Cadmium Effects on Embryo Growth of Pea Seeds during Germination: Investigation of the Mechanisms of Interference of the Heavy Metal with Protein Mobilization-Related Factors. J. Plant Physiol. 2018, 226, 64–76. [Google Scholar] [CrossRef]
  74. Fattahi, B.; Arzani, K.; Souri, M.K.; Barzegar, M. Effects of Cadmium and Lead on Seed Germination, Morphological Traits, and Essential Oil Composition of Sweet Basil (Ocimum basilicum L.). Ind. Crops Prod. 2019, 138, 111584. [Google Scholar] [CrossRef]
  75. Shah, A.A.; Khan, W.U.; Yasin, N.A.; Akram, W.; Ahmad, A.; Abbas, M.; Ali, A.; Safdar, M.N. Butanolide alleviated cadmium stress by improving plant growth, photosynthetic parameters and antioxidant defense system of Brassica oleracea. Chemosphere 2020, 261, 127728. [Google Scholar] [CrossRef]
  76. Chen, L.; Long, C.; Wang, D.; Yang, J. Phytoremediation of cadmium (Cd) and uranium (U) contaminated soils by Brassica juncea L. enhanced with exogenous application of plant growth regulators. Chemosphere 2020, 242, 125112. [Google Scholar] [CrossRef]
  77. Rizwan, M.; Ali, S.; ur Rehman, M.Z.; Adrees, M.; Arshad, M.; Qayyum, M.F.; Ali, L.; Hussain, A.; Chatha, S.A.S.; Imran, M. Alleviation of cadmium accumulation in maize (Zea mays L.) by foliar spray of zinc oxide nanoparticles and biochar to contaminated soil. Environ. Pollut. 2019, 248, 358–367. [Google Scholar] [CrossRef]
  78. Ullah, S.; Khan, J.; Hayat, K.; Abdelfattah Elateeq, A.; Salam, U.; Yu, B.; Ma, Y.; Wang, H.; Tang, Z.-H. Comparative study of growth, cadmium accumulation and tolerance of three chickpea (Cicer arietinum L.) cultivars. Plants 2020, 9, 310. [Google Scholar] [CrossRef]
  79. Tang, L.; Hamid, Y.; Liu, D.; Shohag, M.J.I.; Zehra, A.; He, Z.; Feng, Y.; Yang, X. Foliar application of zinc and selenium alleviates cadmium and lead toxicity of water spinach—Bioavailability/cytotoxicity study with human cell lines. Environ. Int. 2020, 145, 106122. [Google Scholar] [CrossRef] [PubMed]
  80. Hossain, M.S.; Abdelrahman, M.; Tran, C.D.; Nguyen, K.H.; Chu, H.D.; Watanabe, Y.; Fujita, M.; Tran, L.-S.P. Modulation of osmoprotection and antioxidant defense by exogenously applied acetate enhances cadmium stress tolerance in lentil seedlings. Environ. Pollut. 2022, 308, 119687. [Google Scholar] [CrossRef] [PubMed]
  81. Gutsch, A.; Keunen, E.; Guerriero, G.; Renaut, J.; Cuypers, A.; Hausman, J.-F.; Sergeant, K. Long-term cadmium exposure influences the abundance of proteins that impact the cell wall structure in Medicago sativa stems. Plant Biol. 2018, 20, 1023–1035. [Google Scholar] [CrossRef] [PubMed]
  82. Kaur, H.; Garg, N. Interactive effects of zinc-arbuscular mycorrhizal (AM) fungi on cadmium uptake, rubisco, osmolyte synthesis and yield in Cajanus cajan (L.) Millsp. Int. J. Sustain. Agric. Res. 2021, 8, 17–42. [Google Scholar] [CrossRef]
  83. Ahanger, M.A.; Aziz, U.; Alsahli, A.A.; Alyemeni, M.N.; Ahmad, P. Combined kinetin and spermidine treatments ameliorate growth and photosynthetic inhibition in Vigna angularis by up-regulating antioxidant and nitrogen metabolism under cadmium stress. Biomolecules 2020, 10, 147. [Google Scholar] [CrossRef]
  84. Hasan, N.; Choudhary, S.; Laskar, R.A.; Naaz, N.; Sharma, N. Comparative study of cadmium nitrate and lead nitrate [Cd(NO3)2 and Pb(NO3)2] stress in cyto-physiological parameters of Capsicum annuum L. Horticul. Environ. Biotechnol. 2022, 63, 627–641. [Google Scholar] [CrossRef]
  85. Zaid, A.; Mohammad, F. Methyl jasmonate and nitrogen interact to alleviate cadmium stress in Mentha arvensis by regulating physio-biochemical damages and ROS detoxification. J. Plant Growth Regul. 2018, 37, 1331–1348. [Google Scholar] [CrossRef]
  86. Alvarez, A.; Saez, J.M.; Costa, J.S.; Colin, V.L.; Fuentes, M.S.; Cuozzo, S.A.; Benimeli, C.S.; Polti, M.A.; Amoroso, M.J. Actinobacteria: Current research and perspectives for bioremediation of pesticides and heavy metals. Chemosphere 2017, 166, 41–62. [Google Scholar] [CrossRef]
  87. Mustafa, A.; Zulfiqar, U.; Mumtaz, M.Z.; Radziemska, M.; Haider, F.U.; Holatko, J.; Hammershmiedt, T.; Naveed, M.; Ali, H.; Kintl, A.; et al. Nickel (Ni) phytotoxicity and detoxification mechanisms: A review. Chemosphere 2023, 328, 138574. [Google Scholar] [CrossRef]
  88. Dixit, R.; Malaviya, D.; Pandiyan, K.; Singh, U.B.; Sahu, A.; Shukla, R.; Singh, B.P.; Rai, J.P.; Sharma, P.K.; Lade, H.; et al. Bioremediation of heavy metals from soil and aquatic environment: An overview of principles and criteria of fundamental processes. Sustainability 2015, 7, 2189–2212. [Google Scholar] [CrossRef]
  89. Ahemad, M. Remediation of metalliferous soils through the heavy metal resistant plant growth promoting bacteria: Paradigms and prospects. Arab. J. Chem. 2019, 12, 1365–1377. [Google Scholar] [CrossRef]
  90. Shabayev, V.P.; Bocharnikova, E.A.; Ostroumov, V.E. Remediation of cadmium-polluted soil using plant growth-promoting rhizobacteria and natural zeolite. Euras. Soil Sci. 2020, 53, 809–819. [Google Scholar] [CrossRef]
  91. He, X.; Xu, M.; Wei, Q.; Tang, M.; Guan, L.; Lou, L.; Xia, Y. Promotion of Growth and Phytoextraction of Cadmium and Lead in Solanum nigrum L. Mediated by Plant-Growth-Promoting Rhizobacteria. Ecotoxicol. Environ. Saf. 2020, 205, 111333. [Google Scholar] [CrossRef] [PubMed]
  92. Mitra, S.; Purkait, T.; Pramanik, K.; Maiti, T.K.; Dey, R.S. Three-Dimensional Graphene for Electrochemical Detection of Cadmium in Klebsiella michiganensis to Study the Influence of Cadmium Uptake in Rice Plant. Mater. Sci. Eng. C 2019, 103, 109802. [Google Scholar] [CrossRef] [PubMed]
  93. Li, X.; Yan, Z.; Gu, D.; Li, D.; Tao, Y.; Zhang, D.; Ao, Y. Characterization of Cadmium-Resistant Rhizobacteria and Their Promotion Effects on Brassica napus Growth and Cadmium Uptake. J. Basic Microbiol. 2019, 59, 579–590. [Google Scholar] [CrossRef] [PubMed]
  94. Li, Y.; Yu, X.; Cui, Y.; Tu, W.; Shen, T.; Yan, M.; Ma, M. The Potential of Cadmium Ion-Immobilized Rhizobium pusense KG2 to Prevent Soybean Root from Absorbing Cadmium in Cadmium-Contaminated Soil. J. Appl. Microbiol. 2019, 126, 919–930. [Google Scholar] [CrossRef]
  95. Yankey, R.; Karanja, J.K.; Okal, E.J.; Omoor, I.N.A.; Lin, H.; Bodjremou, D.M.; Lin, Z.X. A Consortium of Plant Growth-Promoting Rhizobacteria Strains Synergistically Assists Jujuncao (Pennisetum giganteum) to Remediate Cadmium Contaminated Soils. Appl. Ecol. Environ. Res. 2021, 19, 2425–2442. [Google Scholar] [CrossRef]
  96. Wu, B.; He, T.; Wang, Z.; Qiao, S.; Wang, Y.; Xu, F.; Xu, H. Insight into the Mechanisms of Plant Growth Promoting Strain SNB6 on Enhancing the Phytoextraction in Cadmium Contaminated Soil. J. Hazard. Mater. 2020, 385, 121587. [Google Scholar] [CrossRef]
  97. Abdelkrim, S.; Jebara, S.H.; Saadani, O.; Abid, G.; Taamalli, W.; Zemni, H.; Jebara, M. In Situ Effects of Lathyrus sativus-PGPR to Remediate and Restore Quality and Fertility of Pb and Cd Polluted Soils. Ecotoxicol. Environ. Saf. 2020, 192, 110260. [Google Scholar] [CrossRef]
  98. Li, Y.; Zeng, J.; Wang, S.; Lin, Q.; Ruan, D.; Chi, H.; Qiu, R.; Yang, Y. Effects of Cadmium-Resistant Plant Growth-Promoting Rhizobacteria and Funneliformis mosseae on the Cadmium Tolerance of Tomato (Lycopersicon esculentum L.). Int. J. Phytoremed. 2020, 22, 451–458. [Google Scholar] [CrossRef]
  99. Kotoky, R.; Nath, S.; Kumar Maheshwari, D.; Pandey, P. Cadmium Resistant Plant Growth Promoting Rhizobacteria Serratia marcescens S2I7 Associated with the Growth Promotion of Rice Plant. Environ. Sustain. 2019, 2, 135–144. [Google Scholar] [CrossRef]
  100. Liaquat, F.; Munis, M.F.H.; Arif, S.; Haroon, U.; Shengquan, C.; Qunlu, L. Cd-Tolerant SY-2 Strain of Stenotrophomonas maltophilia: A Potential PGPR, Isolated from the Nanjing Mining Area in China. 3 Biotech 2020, 10, 519. [Google Scholar] [CrossRef] [PubMed]
  101. Belimov, A.A.; Shaposhnikov, A.I.; Azarova, T.S.; Makarova, N.M.; Safronova, V.I.; Litvinskiy, V.A.; Nosikov, V.N.; Zavalin, A.A.; Tikhonovich, I.A. Microbial Consortium of PGPR, Rhizobia and Arbuscular Mycorrhizal Fungus Makes Pea Mutant SGECdt Comparable with Indian Mustard in Cadmium Tolerance and Accumulation. Plants 2020, 9, 975. [Google Scholar] [CrossRef] [PubMed]
  102. You, M.; Wang, L.; Zhou, G.; Wang, Y.; Wang, K.; Zou, R.; Cao, W.; Fan, H. Effects of Microbial Agents on Cadmium Uptake in Solanum nigrum L. and Rhizosphere Microbial Communities in Cadmium-Contaminated Soil. Front. Microbiol. 2022, 13, 1106254. [Google Scholar] [CrossRef] [PubMed]
  103. Liu, A.; Wang, W.; Zheng, X.; Chen, X.; Fu, W.; Wang, G.; Ji, J.; Jin, C.; Guan, C. Improvement of the Cd and Zn Phytoremediation Efficiency of Rice (Oryza sativa) through the Inoculation of a Metal-Resistant PGPR Strain. Chemosphere 2022, 302, 134900. [Google Scholar] [CrossRef] [PubMed]
  104. Sumranwanich, T.; Leartsiwawinyu, W.; Meeinkuirt, W.; Chayapan, P. Application of Plant Growth-Promoting Rhizobacteria (PGPR) Associated with Energy Plant, Pennisetum purpurenum, in Cadmium and Zinc Contaminated Soil. Res. Sq. 2022. [Google Scholar] [CrossRef]
  105. Xu, Z.; Wang, D.; Tang, W.; Wang, L.; Li, Q.; Lu, Z.; Guo, S. Phytoremediation of Cadmium-Polluted Soil Assisted by D-gluconate-Enhanced Enterobacter cloacae Colonization in the Solanum nigrum L. Rhizosphere. Sci. Total Environ. 2020, 732, 139265. [Google Scholar] [CrossRef]
  106. Ali, Q.; Ayaz, M.; Yu, C.; Wang, Y.; Gu, Q.; Wu, H.; Gao, X. Cadmium Tolerant Microbial Strains Possess Different Mechanisms for Cadmium Biosorption and Immobilization in Rice Seedlings. Chemosphere 2022, 303, 135206. [Google Scholar] [CrossRef]
  107. Khanna, K.; Jamwal, V.L.; Sharma, A.; Gandhi, S.G.; Ohri, P.; Bhardwaj, R.; Ahmad, P. Supplementation with Plant Growth Promoting Rhizobacteria (PGPR) Alleviates Cadmium Toxicity in Solanum lycopersicum by Modulating the Expression of Secondary Metabolites. Chemosphere 2019, 230, 628–639. [Google Scholar] [CrossRef]
  108. Ghosh, A.; Pramanik, K.; Bhattacharya, S.; Mondal, S.; Ghosh, S.K.; Maiti, T.K. A Potent Cadmium Bioaccumulating Enterobacter cloacae Strain Displays Phytobeneficial Property in Cd-Exposed Rice Seedlings. Curr. Res. Microb. Sci. 2022, 3, 100101. [Google Scholar] [CrossRef]
  109. Patel, M.; Patel, K.; Al-Keridis, L.A.; Alshammari, N.; Badraoui, R.; Elasbali, A.M.; Adnan, M. Cadmium-Tolerant Plant Growth-Promoting Bacteria Curtobacterium oceanosedimentum Improves Growth Attributes and Strengthens Antioxidant System in Chili (Capsicum frutescens). Sustainability 2022, 14, 4335. [Google Scholar] [CrossRef]
  110. Cheng, X.; Cao, X.; Tan, C.; Liu, L.; Bai, J.; Liang, Y.; Cai, R. Effects of Four Endophytic Bacteria on Cadmium Speciation and Remediation Efficiency of Sedum plumbizincicola in Farmland Soil. Environ. Sci. Pollut. Res. 2022, 29, 89557–89569. [Google Scholar] [CrossRef] [PubMed]
  111. Chellaiah, E.R. Cadmium (Heavy Metals) Bioremediation by Pseudomonas aeruginosa: A Mini review. Appl. Water Sci. 2018, 8, 154. [Google Scholar] [CrossRef]
  112. Smieljan, A.; Wilkinson, K.J.; Rossier, C. Cd Bioaccumulation by a Freshwater Bacterium, Rhodospirillum rubrum. Environ. Sci. Technol. 2003, 37, 701–706. [Google Scholar] [CrossRef] [PubMed]
  113. Arivalagan, P.; Singaraj, D.; Haridass, V.; Kaliannan, T. Removal of Cadmium from Aqueous Solution by Batch Studies Using Bacillus cereus. Ecol. Eng. 2014, 71, 728–735. [Google Scholar] [CrossRef]
  114. Bagot, D.; Lebeau, T.; Jezequel, K. Microorganisms for Remediation of Cadmium-Contaminated Soils. Environ. Chem. Lett. 2006, 4, 207–211. [Google Scholar] [CrossRef]
  115. Ahmad, I.; Akhtar, M.J.; Zahir, Z.A.; Naveed, M.; Mitter, B.; Sessitsch, A. Cadmium-Tolerant Bacteria Induce Metal Stress Tolerance in Cereals. Environ. Sci. Pollut. Res. 2014, 21, 11054–11065. [Google Scholar] [CrossRef]
  116. Brunetti, G.; Farrag, K.; Soler-Rovira, P.; Ferrara, M.; Nigro, F.; Senesi, N. The Effect of Compost and Bacillus licheniformis on the Phytoextraction of Cr, Cu, Pb and Zn by Three Brassicaceae Species from Contaminated Soils in the Apulia Region, Southern Italy. Geoderma 2012, 170, 322–330. [Google Scholar] [CrossRef]
  117. Awan, S.A.; Ilyas, N.; Khan, I.; Raza, M.A.; Rehman, A.U.; Rizwan, M.; Rastogi, A.; Tariq, R.; Brestic, M. Bacillus siamensis Reduces Cadmium Accumulation and Improves Growth and Antioxidant Defense System in Two Wheat (Triticum aestivum L.) Varieties. Plants 2020, 9, 878. [Google Scholar] [CrossRef]
  118. Belimov, A.A.; Hontzeas, N.; Safronova, V.I.; Demchinskaya, S.V.; Piluzza, G.; Bullitta, S.; Glick, B.R. Cadmium-Tolerant Plant Growth-Promoting Bacteria Associated with the Roots of Indian Mustard (Brassica juncea L. Czern.). Soil Biol. Biochem. 2005, 37, 241–250. [Google Scholar] [CrossRef]
  119. Dell’Amico, E.; Cavalca, L.; Andreoni, V. Analysis of Rhizobacterial Communities in Perennial Graminaceae from Polluted Water Meadow Soil, and Screening of Metal-Resistant, Potentially Plant Growth-Promoting Bacteria. FEMS Microbiol. Ecol. 2005, 52, 153–162. [Google Scholar] [CrossRef] [PubMed]
  120. Ganesan, V. Rhizoremediation of Cadmium Soil Using a Cadmium-Resistant Plant Growth-Promoting Rhizopseudomonad. Curr. Microbiol. 2008, 56, 403–407. [Google Scholar] [CrossRef] [PubMed]
  121. Sinha, S.; Mukherjee, S.K. Pseudomonas aeruginosa KUCd1, a Possible Candidate for Cadmium Bioremediation. Braz. J. Microbiol. 2009, 40, 655–662. [Google Scholar] [CrossRef]
  122. Dary, M.; Chamber-Pérez, M.A.; Palomares, A.J.; Pajuelo, E. In Situ Phytostabilisation of Heavy Metal Polluted Soils Using Lupinus luteus Inoculated with Metal Resistant Plant-Growth Promoting Rhizobacteria. J. Hazard. Mater. 2010, 177, 323–330. [Google Scholar] [CrossRef]
  123. Jeong, S.; Moon, H.S.; Nam, K.; Kim, J.Y.; Kim, T.S. Application of Phosphate-Solubilizing Bacteria for Enhancing Bioavailability and Phytoextraction of Cadmium (Cd) from Polluted Soil. Chemosphere 2012, 88, 204–210. [Google Scholar] [CrossRef] [PubMed]
  124. Ahemad, M.; Kibret, M. Mechanisms and Applications of Plant Growth Promoting Rhizobacteria: Current Perspective. J. King Saud Univ. Sci. 2014, 26, 1–20. [Google Scholar] [CrossRef]
  125. Antoniadis, V.; Levizou, E.; Shaheen, S.M.; Ok, Y.S.; Sebastian, A.; Baum, C.; Prasad, M.N.V.; Wenzel, W.W.; Rinklebe, J. Trace Elements in the Soil-Plant Interface: Phytoavailability, Translocation, and Phytoremediation—A Review. Earth-Science Rev. 2017, 171, 621–645. [Google Scholar] [CrossRef]
  126. Satyanarayana, T.; Deshmukh, S.K.; Deshpande, M.V. Advancing Frontiers in Mycology & Mycotechnology: Basic and Applied Aspects of Fungi, 1st ed.; Springer: Singapore, 2019; pp. 1–675. [Google Scholar]
  127. Ali, A.; Guo, D.; Mahar, A.; Wang, P.; Shen, F.; Li, R.; Zhang, Z. Mycoremediation of Potentially Toxic Trace Elements—A Biological Tool for Soil Cleanup: A Review. Pedosphere 2017, 27, 205–222. [Google Scholar] [CrossRef]
  128. Harms, H.; Schlosser, D.; Wick, L.Y. Untapped Potential: Exploiting Fungi in Bioremediation of Hazardous Chemicals. Nat. Rev. Microbiol. 2011, 9, 177–192. [Google Scholar] [CrossRef]
  129. Jacob, J.M.; Karthik, C.; Saratale, R.G.; Kumar, S.S.; Prabakar, D.; Kadirvelu, K.; Pugazhendhi, A. Biological Approaches to Tackle Heavy Metal Pollution: A Survey of Literature. J. Environ. Manag. 2018, 217, 56–70. [Google Scholar] [CrossRef]
  130. Igiri, B.E.; Okoduwa, S.I.; Idoko, G.O.; Akabuogu, E.P.; Adeyi, A.O.; Ejiogu, I.K. Toxicity and bioremediation of heavy metals contaminated ecosystem from tannery wastewater: A review. J. Toxicol. 2018, 2018, 2568038. [Google Scholar] [CrossRef] [PubMed]
  131. Boswell, G.P.; Jacobs, H.; Ritz, K.; Gadd, G.M.; Davidson, F.A. The development of fungal networks in complex environments. Bull. Math. Biol. 2007, 69, 605–634. [Google Scholar] [CrossRef] [PubMed]
  132. Mani, D.; Kumar, C. Biotechnological advances in bioremediation of heavy metals contaminated ecosystems: An overview with special reference to phytoremediation. Int. J. Environ. Sci. Technol. 2014, 11, 843–872. [Google Scholar] [CrossRef]
  133. Kumar, S.; Prasad, S.; Yadav, K.K.; Shrivastava, M.; Gupta, N.; Nagar, S.; Bach, Q.V.; Kamyab, H.; Khan, S.A.; Yadav, S.; et al. Hazardous heavy metals contamination of vegetables and food chain: Role of sustainable remediation approaches—A review. Environ. Res. 2019, 179, 108792. [Google Scholar] [CrossRef] [PubMed]
  134. Yang, Y.; Liang, Y.; Han, X.; Chiu, T.Y.; Ghosh, A.; Chen, H.; Tang, M. The roles of arbuscular mycorrhizal fungi (AMF) in phytoremediation and tree-herb interactions in Pb contaminated soil. Sci. Rep. 2016, 6, 20469. [Google Scholar] [CrossRef]
  135. Begum, N.; Qin, C.; Ahanger, M.A.; Raza, S.; Khan, M.I.; Ashraf, M.; Ahmed, N.; Zhang, L. Role of Arbuscular Mycorrhizal Fungi in Plant Growth Regulation: Implications in Abiotic Stress Tolerance. Front. Plant Sci. 2019, 10, 1068. [Google Scholar] [CrossRef]
  136. Khan, A.G. Producing Mycorrhizal Inoculum for Pytoremedation. In Phytoremediation. Methods in Biotechnology; Humana Press: Totowa, NJ, USA, 2005; pp. 89–97. [Google Scholar]
  137. Xiao, X.; Luo, S.; Zeng, G.; Wei, W.; Wan, Y.; Chen, L.; Guo, H.; Cao, Z.; Yang, L.; Chen, J.; et al. Biosorption of cadmium by endophytic fungus (EF) Microsphaeropsis sp. LSE10 isolated from cadmium hyperaccumulator Solanum nigrum L. Bioresour. Technol. 2010, 101, 1668–1674. [Google Scholar] [CrossRef]
  138. Cabral, L.; Soares, C.R.; Giachini, A.J.; Siqueira, J.O. Arbuscular mycorrhizal fungi in phytoremediation of contaminated areas by trace elements: Mechanisms and major benefits of their applications. World J. Microbiol. Biotechnol. 2015, 31, 1655–1664. [Google Scholar] [CrossRef]
  139. Zhang, F.; Liu, M.; Li, Y.; Che, Y.; Xiao, Y. Effects of arbuscular mycorrhizal fungi, biochar and cadmium on the yield and element uptake of Medicago sativa. Sci. Total Environ. 2019, 655, 1150–1158. [Google Scholar] [CrossRef]
  140. Wang, F.; Liu, X.; Shi, Z.; Tong, R.; Adams, C.A.; Shi, X. Arbuscular mycorrhizae alleviate negative effects of zinc oxide nanoparticle and zinc accumulation in maize plants—A soil microcosm experiment. Chemosphere 2016, 147, 88–97. [Google Scholar] [CrossRef]
  141. Garg, N.; Bhandari, P. Cadmium toxicity in crop plants and its alleviation by arbuscular mycorrhizal (AM) fungi: An overview. Plant Biosyst. 2014, 148, 609–621. [Google Scholar] [CrossRef]
  142. Varma, A.; Sherameti, I.; Tripathi, S.; Prasad, R.; Das, A.; Sharma, M.; Bakshi, M.; Johnson, J.M.; Bhardwaj, S.; Arora, M.; et al. The symbiotic fungus Piriformosporaindica: Review. In Fungal Associations; Springer: Berlin/Heidelberg, Germany, 2012; pp. 231–254. [Google Scholar]
  143. Sahu, A.; Mandal, A.; Thakur, J.; Manna, M.; Rao, A.S. Exploring bioaccumulation efficacy of Trichoderma viride: An alternative bioremediation of cadmium and lead. Natl. Acad. Sci. Lett. 2012, 35, 299–302. [Google Scholar] [CrossRef]
  144. Yaghoubian, Y.; Siadat, S.A.; Telavat, M.R.M.; Pirdashti, H.; Yaghoubia, I. Bioremoval of cadmium from aqueous solutions by filamentous fungi: Trichoderma spp. and Piriformospora indica. Environ. Sci. Pollut. Res. 2019, 26, 7863–7872. [Google Scholar] [CrossRef]
  145. Barros, L.; Macedo, G.; Duarte, M.; Silva, E.; Lobato, A.K.C.L. Biosorption of cadmium using the fungus Aspergillus niger. Braz. J. Chem. Eng. 2003, 20, 229–239. [Google Scholar] [CrossRef]
  146. Tripathi, P.; Singh, P.C.; Mishra, A.; Chauhan, P.S.; Dwivedi, S.; Bais, R.T.; Tripathi, R.D. Trichoderma: A Potential Bioremediator for Environmental Clean up. Clean Technol. Environ. Policy 2013, 15, 541–550. [Google Scholar] [CrossRef]
  147. Wang, B.; Liu, L.; Gao, Y.; Chen, J. Improved phytoremediation of oilseed rape (Brassica napus) by Trichoderma mutant constructed by restriction enzyme-mediated integration (REMI) in cadmium polluted soil. Chemosphere 2009, 74, 1400–1403. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, M.; Zhou, Q. Single and Joint Toxicity of Chlorimuron-Ethyl, Cadmium, and Copper Acting on Wheat Triticum aestivum. Ecotoxicol. Environ. Saf. 2005, 60, 169–175. [Google Scholar] [CrossRef]
  149. Aksu, Z. Equilibrium and Kinetic Modeling of Cadmium (II) Biosorption by C. vulgaris in a Batch System. Purify Technol. 2002, 21, 285–294. [Google Scholar] [CrossRef]
  150. Mohsenzadeh, F.; Shahrokhi, F. Biological Removing of Cadmium from Contaminated Media by Fungal Biomass of Trichoderma Species. J. Environ. Health Sci. Eng. 2014, 12, 102. [Google Scholar] [CrossRef]
  151. Olguín, E.J.; Sánchez-Galván, G. Heavy Metal Removal in Phytofiltration and Phycoremediation: The Need to Differentiate between Bioadsorption and Bioaccumulation. New Biotechnol. 2012, 30, 3–8. [Google Scholar] [CrossRef]
  152. Sattayawat, P.; Yunus, I.S.; Noirungsee, N.; Mukjang, N.; Pathom-Aree, W.; Pekkoh, J.; Pumas, C. Synthetic Biology-Based Approaches for Microalgal Bio-removal of Heavy Metals from Wastewater Effluents. Front. Environ. Sci. 2021, 9, 562. [Google Scholar] [CrossRef]
  153. Kumar, S.S.; Kadier, A.; Malyan, S.K.; Ahmad, A.; Bishnoi, N.R. Phytoremediation and Rhizoremediation: Uptake, Mobilization and Sequestration of Heavy Metals by Plants, Plant-Microbe Interactions in Agro-Ecological Perspectives; Springer: Singapore, 2017; pp. 309–332. [Google Scholar]
  154. Chen, H.; Arocena, J.M.; Li, J.; Thring, R.W.; Zhou, J. Assessments of Chromium (and Other Metals) in Vegetables and Potential Bio-accumulations in Humans Living in Areas Affected by Tannery Wastes. Chemosphere 2014, 112, 412–419. [Google Scholar] [CrossRef] [PubMed]
  155. Davis, T.A.; Volesky, B.; Mucci, A. A Review of the Biochemistry of Heavy Metal Biosorption by Brown Algae. Water Res. 2003, 37, 4311–4330. [Google Scholar] [CrossRef] [PubMed]
  156. Rybak, A.; Messyasz, B.; Łeska, B. Freshwater Ulva (Chlorophyta) as a Bioaccumulator of Selected Heavy Metals (Cd, Ni and Pb) and Alkaline Earth Metals (Ca and Mg). Chemosphere 2012, 89, 1066–1076. [Google Scholar] [CrossRef]
  157. Soeprobowati, T.R.; Hariyati, R. Bioaccumulation of Pb, Cd, Cu, and Cr by Porphyridium cruentum (S.F. Gray) Nägeli. Int. J. Mar. Sci. 2013, 3, 212–218. [Google Scholar] [CrossRef]
  158. Saunders, R.J.; Paul, N.A.; Hu, Y.; de Nys, R. Sustainable sources of biomass for bioremediation of heavy metals in waste water derived from coal-fired power generation. PLoS ONE. 2012, 7, e36470. [Google Scholar] [CrossRef] [PubMed]
  159. Kumar, J.I.; Oommen, C. Removal of heavy metals by biosorption using freshwater alga Spirogyra hyalina. J. Environ. Biol. 2012, 33, 27–31. [Google Scholar]
  160. Kotrba, P.; Mackova, M.; Macek, T. Microbial Biosorption of Metals—General Introduction. In Microbial Biosorption of Metals; Springer: Dordrecht, The Netherlands, 2011; pp. 1–329. [Google Scholar]
  161. Tuzen, M.; Sari, A. Biosorption of Selenium from Aqueous Solution by Green Algae (Cladophora hutchinsiae) Biomass: Equilibrium, Thermodynamic and Kinetic Studies. Chem. Eng. J. 2010, 158, 200–206. [Google Scholar] [CrossRef]
  162. Sarin, C.; Sarin, S. Removal of Cadmium and Zinc from Soil Using Immobilized Cell of Biosurfactant Producing Bacteria. Environ. Asia 2010, 3, 49–53. [Google Scholar]
  163. Zhu, Z.; Yang, X.; Wang, K.; Huang, H.; Zhang, X.; Fang, H.; He, Z. Bioremediation of Cd-DDT Co-Contaminated Soil Using the Cd-Hyperaccumulator Sedum alfredii and DDT-Degrading Microbes. J. Hazard. Mater. 2012, 235–236, 144–151. [Google Scholar] [CrossRef]
  164. Haq, F.; Butt, M.; Ali, H.; Chaudhary, H.J. Biosorption of Cadmium and Chromium from Water by Endophytic Kocuria rhizophila: Equilibrium and Kinetic Studies. Desalination Water Treat. 2015, 57, 19946–19958. [Google Scholar] [CrossRef]
  165. Mitra, S.; Pramanik, K.; Ghosh, P.K.; Soren, T.; Sarkar, A.; Dey, R.S.; Pandey, S.; Maiti, T.K. Characterization of Cd-Resistant Klebsiella michiganensis MCC3089 and Its Potential for Rice Seedling Growth Promotion under Cd Stress. Microbiol. Res. 2018, 210, 12–25. [Google Scholar] [CrossRef] [PubMed]
  166. Mitra, S.; Pramanik, K.; Sarkar, A.; Ghosh, P.K.; Soren, T.; Maiti, T.K. Bioaccumulation of Cadmium by Enterobacter sp. and Enhancement of Rice Seedling Growth under Cadmium Stress. Ecotoxicol. Environ. Saf. 2018, 156, 183–196. [Google Scholar] [CrossRef] [PubMed]
  167. Peng, W.; Li, X.; Song, J.; Jiang, W.; Liu, Y.; Fan, W. Bioremediation of cadmium-and zinc-contaminated soil using Rhodobacter sphaeroides. Chemosphere 2018, 197, 33–41. [Google Scholar] [CrossRef]
  168. Khadim, H.J.; Ammar, S.H.; Ebrahim, S.E. Biomineralization based remediation of cadmium and nickel contaminated wastewater by ureolytic bacteria isolated from barn horses soil. Environ. Technol. Innov. 2019, 14, 100315. [Google Scholar] [CrossRef]
  169. Zhang, C.; Tao, Y.; Li, S.; Ke, T.; Wang, P.; Wei, S.; Chen, L. Bioremediation of cadmium-trichlorfon co-contaminated soil by Indian mustard (Brassica juncea) associated with the trichlorfon-degrading microbe Aspergillus sydowii: Related physiological responses and soil enzyme activities. Ecotoxicol. Environ. Saf. 2019, 179, 109756. [Google Scholar] [CrossRef]
  170. Zeng, X.; Xu, H.; Lu, J.; Chen, Q.; Li, W.; Wu, L.; Tang, J.; Ma, L. The Immobilization of Soil Cadmium by the Combined Amendment of Bacteria and Hydroxyapatite. Sci. Rep. 2020, 10, 20473. [Google Scholar] [CrossRef]
  171. Wang, Y.; Luo, Y.; Zeng, G.; Wu, X.; Wu, B.; Li, X.; Xu, H. Characteristics and in situ remediation effects of heavy metal immobilizing bacteria on cadmium and nickel co-contaminated soil. Ecotoxicol. Environ. Saf. 2020, 192, 110294. [Google Scholar] [CrossRef]
  172. Ma, H.; Wei, M.; Wang, Z.; Hou, S.; Li, X.; Heng, X. Bioremediation of cadmium polluted soil using a novel cadmium immobilizing plant growth promotion strain Bacillus sp. TZ5 loaded on biochar. J. Hazard. Mater. 2020, 388, 124065. [Google Scholar] [CrossRef]
  173. Xu, M.; Liu, Y.; Deng, Y.; Zhang, S.; Hao, X.; Zhu, P.; Zhou, J.; Yin, H.; Liang, Y.; Jiang, H.; et al. Bioremediation of cadmium-contaminated paddy soil using an autotrophic and heterotrophic mixture. RSC Adv. 2020, 10, 26090–26101. [Google Scholar] [CrossRef]
  174. Minari, G.D.; Saran, L.M.; Lima Constancio, M.T.; Correia da Silva, R.; Rosalen, D.L.; José de Melo, W.; Carareto Alves, L.M. Bioremediation potential of new cadmium, chromium, and nickel-resistant bacteria isolated from tropical agricultural soil. Ecotoxicol. Environ. Saf. 2020, 204, 111038. [Google Scholar] [CrossRef] [PubMed]
  175. Oziegbe, O.; Oluduro, A.O.; Oziegbe, E.J.; Ahuekwe, E.F.; Olorunsola, S.J. Assessment of heavy metal bioremediation potential of bacterial isolates from landfill soils. Saudi J. Biol. Sci. 2021, 28, 3948–3956. [Google Scholar] [CrossRef] [PubMed]
  176. Haider, F.U.; Coulter, J.A.; Cheema, S.A.; Farooq, M.; Wu, J.; Zhang, R.; Shuaijie, G.; Liqun, C. Co-application of Biochar and Microorganisms Improves Soybean Performance and Remediate Cadmium-Contaminated Soil. Ecotoxicol. Environ. Saf. 2021, 214, 112112. [Google Scholar] [CrossRef] [PubMed]
  177. Yu, X.; Zhao, J.; Liu, X.; Sun, L.; Tian, J.; Wu, N. Cadmium Pollution Impact on the Bacterial Community Structure of Arable Soil and the Isolation of the Cadmium Resistant Bacteria. Front. Microbiol. 2021, 12, 698834. [Google Scholar] [CrossRef] [PubMed]
  178. Mo, T.; Jiang, D.; Shi, D.; Xu, S.; Huang, X.; Huang, Z. Remediation mechanism of “double-resistant” bacteria—Sedum alfredii Hance on Pb- and Cd-contaminated soil. Ecol. Process 2022, 11, 20. [Google Scholar] [CrossRef]
  179. Haider, F.U.; Farooq, M.; Naveed, M.; Cheema, S.A.; ul Ain, N.; Salim, M.A.; Liqun, C.; Mustafa, A. Influence of Biochar and Microorganism Co-Application on Stabilization of Cadmium (Cd) and Improved Maize Growth in Cd-Contaminated Soil. Front. Plant Sci. 2022, 13, 983830. [Google Scholar] [CrossRef]
  180. Yuan, B.; Huang, L.; Liu, X.; Bai, L.; Liu, H.; Jiang, H.; Zhu, P.; Xiao, Y.; Geng, J.; Liu, Q.; et al. Application of Mixotrophic Acidophiles for the Bioremediation of Cadmium-Contaminated Soils Elevates Cadmium Removal, Soil Nutrient Availability, and Rice Growth. Ecotoxicol. Environ. Saf. 2022, 236, 113499. [Google Scholar] [CrossRef]
  181. Lu, H.; Xia, C.; Chinnathambi, A.; Nasif, O.; Narayanan, M.; Shanmugam, S.; Chi, N.T.L.; Pugazhendhi, A.; On-Uma, R.; Jutamas, K.; et al. Evaluation of Cadmium Tolerance and Remediated Efficacy of Wild and Mutated Enterobacter Species Isolated from Potassium Nitrate (KNO₃) Processing Unit Contaminated Soil. Chemosphere 2023, 311, 136899. [Google Scholar] [CrossRef]
  182. Xu, T.; Xi, J.; Ke, J.; Wang, Y.; Chen, X.; Zhang, Z.; Lin, Y. Deciphering soil amendments and actinomycetes for remediation of cadmium (Cd) contaminated farmland. Ecotoxicol. Environ. Saf. 2023, 249, 114388. [Google Scholar] [CrossRef]
  183. Peng, C.; Zhao, X.; Ji, X.; Wu, J.; Liang, W.; Song, H.; Zhang, W.; Wang, X. Mixed Bacteria Passivation for the Remediation of Arsenic, Lead, and Cadmium: Medium Optimization and Mechanisms. Process Saf. Environ. Prot. 2023, 170, 720–727. [Google Scholar] [CrossRef]
  184. Khan, I.; Aftab, M.; Shakir, S.U.; Ali, M.; Qayyum, S.; Rehman, M.U.; Haleem, K.S.; Touseef, I. Mycoremediation of heavy metal (Cd and Cr)–polluted soil through indigenous metallotolerant fungal isolates. Environ. Monit. Assess. 2019, 191, 585. [Google Scholar] [CrossRef] [PubMed]
  185. Jin, Z.; Deng, S.; Wen, Y.; Jin, Y.; Pan, L.; Zhang, Y.; Black, T.; Jones, K.C.; Zhang, H.; Zhang, D. Application of Simplicillium chinense for Cd and Pb biosorption and enhancing heavy metal phytoremediation of soils. Sci. Total Environ. 2019, 697, 134148. [Google Scholar] [CrossRef] [PubMed]
  186. Bano, A.; Hussain, J.; Akbar, A.; Mehmood, K.; Anwar, M.; Hasni, M.S.; Ullah, S.; Sajid, S.; Ali, I. Biosorption of heavy metals by obligate halophilic fungi. Chemosphere 2018, 199, 218–222. [Google Scholar] [CrossRef] [PubMed]
  187. Noormohamadi, H.R.; Fat’hi, M.R.; Ghaedi, M.; Ghezelbash, G.R. Potentiality of white-rot fungi in biosorption of nickel and cadmium: Modeling optimization and kinetics study. Chemosphere 2019, 216, 124–130. [Google Scholar] [CrossRef]
  188. Lamrood, P.Y.; Ralegankar, S.D. Biosorption of Cu, Zn, Fe, Cd, Pb and Ni by non-treated biomass of some edible mushrooms. Asian J. Exp. Biol. Sci. 2013, 4, 190195. [Google Scholar]
  189. Nagy, B.; Maicaneanu, A.; Indolean, C.; Manzatu, C.; Silaghi-Dumitrescu, L.; Majdik, C. Comparative study of Cd(II) biosorption on cultivated Agaricus bisporus and wild Lactarius piperatus based biocomposites. Linear and nonlinear equilibrium modelling and kinetics. J. Taiwan Inst. Chem. Eng. 2014, 45, 921–929. [Google Scholar] [CrossRef]
  190. Hussain, B.; Ashraf, M.N.; ur Rahman, S.; Abbas, A.; Li, J.; Farooq, M. Cadmium stress in paddy fields: Effects of soil conditions and remediation strategies. Sci. Total Environ. 2021, 754, 142188. [Google Scholar] [CrossRef]
  191. Romera, E.; González, F.; Ballester, A.; Blázquez, M.L.; Munoz, J.A. Comparative study of biosorption of heavy metals using different types of algae. Bioresour. Technol. 2007, 98, 3344–3353. [Google Scholar] [CrossRef]
  192. Sjahrul, M.; Arifin, D. Phytoremediation of Cd2+ by marine phytoplanktons, Tetracelmis chuii and Chaetoceros calcitrans. Int. J. Chem. 2012, 4, 69–74. [Google Scholar] [CrossRef]
  193. Ibrahim, W.M.; Hassan, A.F.; Azab, Y.A. Biosorption of toxic heavy metals from aqueous solution by Ulva lactuca activated carbon. Egypt. J. Basic Appl. Sci. 2016, 3, 241–249. [Google Scholar] [CrossRef]
  194. Sooksawat, N.; Meetam, M.; Kruatrachue, M.; Pokethitiyook, P.; Inthorn, D. Equilibrium and kinetic studies on biosorption potential of charophyte biomass to remove heavy metals from synthetic metal solution and municipal wastewater. Bioremed. J. 2016, 20, 240–251. [Google Scholar] [CrossRef]
  195. Chandrashekharaiah, P.; Debanjan, S.; Santanu, D.; Avishek, B. Cadmium biosorption and biomass production by two freshwater microalgae Scenedesmus acutus and Chlorella pyrenoidosa: An integrated approach. Chemosphere 2021, 269, 128755. [Google Scholar]
  196. Ahanger, M.A.; Qin, C.; Begum, N.; Maodong, Q.; Dong, X.X.; El-Esawi, M.; El-Sheikh, M.-A.; Alatar, A.A.; Zhang, L. Nitrogen availability prevents oxidative effects of salinity on wheat growth and photosynthesis by up-regulating the antioxidants and osmolytes metabolism, and secondary metabolite accumulation. BMC Plant Biol. 2019, 19, 479. [Google Scholar] [CrossRef] [PubMed]
  197. Ruiz-González, C.; Rodríguez-Pie, L.; Maister, O.; Rodellas, V.; Alorda-Keinglass, A.; Diego-Feliu, M.; Folch, A.; Garcia-Orellana, J.; Gasol, J.M. High spatial heterogeneity and low connectivity of bacterial communities along a Mediterranean subterranean estuary. Mol. Ecol. 2022, 31, 5745–5764. [Google Scholar] [CrossRef] [PubMed]
  198. Plett, D.C.; Ranathunge, K.; Melino, V.J.; Kuya, N.; Uga, Y.; Kronzucker, H.J. The intersection of nitrogen nutrition and water use in plants: New paths toward improved crop productivity. J. Exp. Bot. 2020, 71, 4452–4468. [Google Scholar] [CrossRef] [PubMed]
  199. Raza, A.; Zahra, N.; Hafeez, M.B.; Ahmad, M.; Iqbal, S.; Shaukat, K.; Ahmad, G. Nitrogen fixation of legumes: Biology and Physiology. In The Plant Family Fabaceae: Biology and Physiological Responses to Environmental Stresses; Springer: Singapore, 2020; pp. 43–74. [Google Scholar]
  200. Dutta, S.; Bhattacharjya, D.; Sinha, S.; Mandal, A.K. Salt-tolerant and plant growth-promoting Rhizobacteria: A new-fangled approach for improving crop yield. In Harsh Environment and Plant Resilience: Molecular and Functional Aspects; Springer: Singapore, 2021; pp. 367–385. [Google Scholar]
  201. Hussain, M.M.; Farooqi, Z.U.R.; Rasheed, F.; Din, W.M.U. Role of microorganisms as climate engineers: Mitigation and adaptations to climate change. In Climate Change and Microbes: Impacts and Vulnerability; Springer: Singapore, 2021; pp. 1–18. [Google Scholar]
  202. Singh, R.K.; Singh, P.; Li, H.-B.; Song, Q.-Q.; Guo, D.-J.; Solanki, M.K.; Verma, K.K.; Malviya, M.K.; Song, X.-P.; Lakshmanan, P.; et al. Diversity of nitrogen-fixing rhizobacteria associated with sugarcane: A comprehensive study of plant-microbe interactions for growth enhancement in Saccharum spp. BMC Plant Biol. 2020, 20, 220. [Google Scholar] [CrossRef]
  203. Leakey, R.R. The Role of Trees in Agroecology. In Routledge Handbook of Agricultural Biodiversity; Routledge: London, UK, 2017; pp. 238–252. [Google Scholar]
  204. Gómez-Sagasti, M.T.; Marino, D. PGPRs and Nitrogen-Fixing Legumes: A Perfect Team for Efficient Cd Phytoremediation? Front. Plant Sci. 2015, 6, 81. [Google Scholar]
  205. Cao, Z.; Kühn, P.; He, J.-S.; Bauhus, J.; Guan, Z.-H.; Scholten, T. Calibration of Near-Infrared Spectra for Phosphorus Fractions in Grassland Soils on the Tibetan Plateau. Agronomy 2022, 12, 783. [Google Scholar] [CrossRef]
  206. Li, Y.; Wei, S.; Chen, X.; Dong, Y.; Zeng, M.; Yan, C.; Hou, L.; Jiao, R. Isolation of cadmium-resistance and siderophore-producing endophytic bacteria and their potential use for soil cadmium remediation. Heliyon 2023, 9, e17661. [Google Scholar] [CrossRef]
  207. Kalayu, G. Phosphate solubilizing microorganisms: Promising approach as biofertilizers. Int. J. Agron. 2019, 2019, 4917256. [Google Scholar] [CrossRef]
  208. Gerke, J. Phytate (inositol hexakisphosphate) in soil and phosphate acquisition from inositol phosphates by higher plants. A review. Plants 2015, 4, 253–266. [Google Scholar] [CrossRef] [PubMed]
  209. Gurdeep, K.; Reddy, M.S. Effects of phosphate-solubilizing bacteria, rock phosphate and chemical fertilizers on maize-wheat cropping cycle and economics. Pedosphere 2015, 25, 428–437. [Google Scholar]
  210. Yang, P.; Zhou, X.F.; Wang, L.L.; Li, Q.S.; Zhou, T.; Chen, Y.K.; Zhao, Z.Y.; He, B.Y. Effect of phosphate-solubilizing bacteria on the mobility of insoluble cadmium and metabolic analysis. Int. J. Environ. Rese. Public Health 2018, 15, 1330. [Google Scholar] [CrossRef] [PubMed]
  211. Chibuike, G.U.; Obiora, S.C. Heavy metal polluted soils: Effect on plants and bioremediation methods. Appl. Environ. Soil Sci. 2014, 2014, 752708. [Google Scholar] [CrossRef]
  212. Ren, Z.; Cheng, R.; Chen, P.; Xue, Y.; Xu, H.; Yin, Y.; Zhang, L. Plant-associated microbe system in treatment of heavy metals–contaminated soil: Mechanisms and applications. Water Air Soil Pollut. 2023, 234, 39. [Google Scholar] [CrossRef]
  213. Khan, A. Promises and potential of in situ nano-phytoremediation strategy to mycorrhizo-remediate heavy metal contaminated soils using non-food bioenergy crops (Vetiver zizinoides & Cannabis sativa). Int. J. Phytoremed. 2020, 22, 900–915. [Google Scholar]
  214. Alsamhary, K. Vermi-cyanobacterial remediation of cadmium-contaminated soil with rice husk biochar: An eco-friendly approach. Chemosphere 2023, 311 Pt 1, 136931. [Google Scholar] [CrossRef]
  215. Fahad, S.; Hussain, S.; Bano, A.; Saud, S.; Hassan, S.; Shan, D.; Wu, C. Potential role of phytohormones and plant growth-promoting rhizobacteria in abiotic stresses: Consequences for changing environment. Environ. Sci. Pollut. Res. 2015, 22, 4907–4921. [Google Scholar] [CrossRef]
  216. Farooqi, Z.U.R.; Din, W.M.U.; Hussain, M.M. Microbial Responses under Climate Change Scenarios: Adaptation and Mitigations Climate Change and Microbial Diversity; Apple Academic Press: Burlington, ON, Canada, 2023; pp. 1–20. [Google Scholar]
  217. Pathania, P.; Bhatia, R.; Khatri, M. Cross-competence and affectivity of maize rhizosphere bacteria Bacillus sp. MT7 in tomato rhizosphere. Sci. Horticul. 2020, 272, 109480. [Google Scholar] [CrossRef]
  218. Pan, W.; Lu, Q.; Xu, Q.R.; Zhang, R.R.; Li, H.Y.; Yang, Y.H.; Liu, H.J.; Du, S.T. Abscisic acid-generating bacteria can reduce Cd concentration in pakchoi grown in Cd-contaminated soil. Ecotoxicol. Environ. Saf. 2019, 177, 100–107. [Google Scholar] [CrossRef]
  219. Morales-Cedeño, L.R.; del Carmen Orozco-Mosqueda, M.; Loeza-Lara, P.D.; Parra-Cota, F.I.; de Los Santos-Villalobos, S.; Santoyo, G. Plant growth-promoting bacterial endophytes as biocontrol agents of pre-and post-harvest diseases: Fundamentals, methods of application and future perspectives. Microbiol. Res. 2021, 242, 126612. [Google Scholar] [CrossRef] [PubMed]
  220. Chandra, H.; Kumari, P.; Bisht, R.; Prasad, R.; Yadav, S. Plant growth promoting Pseudomonas aeruginosa from Valeriana wallichii displays antagonistic potential against three phytopathogenic fungi. Mol. Biol. Rep. 2020, 47, 6015–6026. [Google Scholar] [CrossRef] [PubMed]
  221. Kareem, A.; Farooqi, Z.U.R.; Kalsom, A.; Mohy-Ud-Din, W.; Hussain, M.M.; Raza, M.; Khursheed, M.M. Organic farming for sustainable soil use, management, food production and climate change mitigation. In Sustainable Agriculture: Technical Progressions and Transitions; Springer: Cham, Switzerland, 2022; pp. 39–59. [Google Scholar]
  222. Sivasakthi, S.; Usharani, G.; Saranraj, P. Biocontrol potentiality of plant growth promoting bacteria (PGPR)-Pseudomonas fluorescens and Bacillus subtilis: A review. Afr. J. Agric. Res. 2014, 9, 1265–1277. [Google Scholar]
  223. Mahmud, K.; Missaoui, A.; Lee, K.; Ghimire, B.; Presley, H.W.; Makaju, S. Rhizosphere microbiome manipulation for sustainable crop production. Curr. Plant Biol. 2021, 27, 100210. [Google Scholar] [CrossRef]
  224. Thathana, M.G.; Murage, H.; Abia, A.L.K.; Pillay, M. Morphological Characterization and Determination of Aflatoxin-Production Potentials of Aspergillus flavus Isolated from Maize and Soil in Kenya. Agriculture 2017, 7, 80. [Google Scholar] [CrossRef]
  225. Vurukonda, S.S.K.P.; Giovanardi, D.; Stefani, E. Plant Growth Promoting and Biocontrol Activity of Streptomyces spp. as Endophytes. Int. J. Mol. Sci. 2018, 19, 952. [Google Scholar] [CrossRef] [PubMed]
  226. Lopez-Lima, D.; Mtz-Enriquez, A.I.; Carrión, G.; Basurto-Cereceda, S.; Pariona, N. The Bifunctional Role of Copper Nanoparticles in Tomato: Effective Treatment for Fusarium Wilt and Plant Growth Promoter. Sci. Horticul. 2021, 277, 109810. [Google Scholar] [CrossRef]
  227. Ramadan, E.M.; AbdelHafez, A.A.; Hassan, E.A.; Saber, F.M. Plant growth promoting rhizobacteria and their potential for biocontrol of phytopathogens. Afr. J. Microbiol. Res. 2016, 10, 486–504. [Google Scholar]
  228. Paul, A.; Bhakta, J.N. Biosorption-driven green technology for the treatment of heavy metal (loids)-contaminated effluents. In Intelligent Environmental Data Monitoring for Pollution Management; Elsevier: Amsterdam, The Netherlands, 2021; pp. 71–91. [Google Scholar]
  229. Tarfeen, N.; Nisa, K.U.; Hamid, B.; Bashir, Z.; Yatoo, A.M.; Dar, M.A.; Mohiddin, F.A.; Amin, Z.; Ahmad, R.A.A.; Sayyed, R.Z. Microbial remediation: A promising tool for reclamation of contaminated sites with special emphasis on heavy metal and pesticide pollution: A review. Processes 2022, 10, 1358. [Google Scholar] [CrossRef]
  230. Hou, D.; O’Connor, D.; Igalavithana, A.D.; Alessi, D.S.; Luo, J.; Tsang, D.C.; Cundy, A.B.; Rinklebe, J.; Bolan, N.S.; Ok, Y.S. Metal Contamination and Bioremediation of Agricultural Soils for Food Safety and Sustainability. Nat. Rev. Earth Environ. 2020, 1, 366–381. [Google Scholar] [CrossRef]
  231. Oliva-Arancibia, B.; Órdenes-Aenishanslins, N.; Bruna, N.; Ibarra, P.S.; Zacconi, F.C.; Pérez-Donoso, J.M.; Poblete-Castro, I. Co-synthesis of Medium-Chain-Length Polyhydroxyalkanoates and CdS Quantum Dots Nanoparticles in Pseudomonas putida KT2440. J. Biotechnol. 2017, 264, 29–37. [Google Scholar] [CrossRef] [PubMed]
  232. Gupta, S.; Singh, D. Role of Genetically Modified Microorganisms in Heavy Metal Bioremediation. In Advances in Environmental Biotechnology; Springer: Singapore, 2017; pp. 197–214. [Google Scholar]
  233. Etesami, H. Bacterial mediated alleviation of heavy metal stress and decreased accumulation of metals in plant tissues: Mechanisms and future prospects. Ecotoxicol. Environ. Saf. 2018, 147, 175–191. [Google Scholar] [CrossRef]
  234. Gazitúa, M.C.; Morgante, V.; Poupin, M.J.; Ledger, T.; Rodríguez-Valdecantos, G.; Herrera, C.; Jorquera, M.A.; González, B. The Microbial Community from the Early-Plant Colonizer (Baccharis linearis) Is Required for Plant Establishment on Copper Mine Tailings. Sci. Rep. 2021, 11, 10448. [Google Scholar] [CrossRef] [PubMed]
  235. Kalaivanan, D.; Ganeshamurthy, A.N. Mechanisms of Heavy Metal Toxicity in Plants. In Abiotic Stress Physiology of Horticultural Crops; Springer: Singapore, 2016; pp. 85–102. [Google Scholar]
  236. Araki, R.; Murata, J.; Murata, Y. A Novel Barley Yellow Stripe 1-Like Transporter (HvYSL2) Localized to the Root Endodermis Transports Metal–Phytosiderophore Complexes. Plant Cell Physiol. 2011, 52, 1931–1940. [Google Scholar] [CrossRef] [PubMed]
  237. Złoch, M.; Thiem, D.; Gadzała-Kopciuch, R.; Hrynkiewicz, K. Synthesis of Siderophores by Plant-Associated Metallotolerant Bacteria under Exposure to Cd2+. Chemosphere 2016, 156, 312–325. [Google Scholar] [CrossRef]
  238. Wang, W.; Qiu, Z.; Tan, H.; Cao, L. Siderophore Production by Actinobacteria. Biometals 2014, 27, 623–631. [Google Scholar] [CrossRef]
  239. Kang, P.; Wu, P.; Jin, Y.; Shi, S.; Gao, D.; Chen, G.; Li, Q. Formation and Emissions of Volatile Organic Compounds from Homo-PP and Co-PP Resins during Manufacturing Process and Accelerated Photoaging Degradation. Molecules 2020, 25, 2761. [Google Scholar] [CrossRef]
  240. Mhlongo, M.I.; Piater, L.A.; Madala, N.E.; Labuschagne, N.; Dubery, I.A. The Chemistry of Plant–Microbe Interactions in the Rhizosphere and the Potential for Metabolomics to Reveal Signaling Related to Defense Priming and Induced Systemic Resistance. Front. Plant Sci. 2018, 9, 112. [Google Scholar] [CrossRef]
  241. Montejano-Ramírez, V.; García-Pineda, E.; Valencia-Cantero, E. Bacterial Compound N, N-Dimethylhexadecylamine Modulates Expression of Iron Deficiency and Defense Response Genes in Medicago truncatula Independently of the Jasmonic Acid Pathway. Plants 2020, 9, 624. [Google Scholar] [CrossRef]
  242. Montejano-Ramírez, V.; Martínez-Cámara, R.; García-Pineda, E.; Valencia-Cantero, E. Rhizobacterium Arthrobacter agilis UMCV2 Increases Organ-Specific Expression of FRO Genes in Conjunction with Genes Associated with the Systemic Resistance Pathways of Medicago truncatula. Acta Physiol. Plant. 2018, 40, 138. [Google Scholar] [CrossRef]
  243. Kong, H.G.; Shin, T.S.; Kim, T.H.; Ryu, C.M. Stereoisomers of the Bacterial Volatile Compound 2, 3-Butanediol Differently Elicit Systemic Defense Responses of Pepper against Multiple Viruses in the Field. Front. Plant Sci. 2018, 9, 90. [Google Scholar] [CrossRef]
  244. Rojas-Solis, D.; Vences-Guzmán, M.A.; Sohlenkamp, C.; Santoyo, G. Bacillus toyonensis COPE52 modifies lipid and fatty acid composition, exhibits antifungal activity, and stimulates growth of tomato plants under saline conditions. Curr. Microbiol. 2020, 77, 2735–2744. [Google Scholar] [CrossRef] [PubMed]
  245. Sharma, P. Role and significance of biofilm-forming microbes in phytoremediation—A review. Environ. Technol. Innov. 2022, 25, 102182. [Google Scholar] [CrossRef]
  246. Dimkić, I.; Janakiev, T.; Petrović, M.; Degrassi, G.; Fira, D. Plant-associated Bacillus and Pseudomonas antimicrobial activities in plant disease suppression via biological control mechanisms—A review. Physiol. Mol. Plant Pathol. 2022, 117, 101754. [Google Scholar] [CrossRef]
  247. Imran, M.; Das, K.R.; Naik, M.M. Co-selection of multi-antibiotic resistance in bacterial pathogens in metal and microplastic contaminated environments: An emerging health threat. Chemosphere 2019, 215, 846–857. [Google Scholar] [CrossRef]
  248. Gelbicova, T.; Florianova, M.; Hluchanova, L.; Kalova, A.; Korena, K.; Strakova, N.; Karpiskova, R. Comparative analysis of genetic determinants encoding cadmium, arsenic, and benzalkonium chloride resistance in Listeria monocytogenes of human, food, and environmental origin. Front. Microbiol. 2021, 11, 599882. [Google Scholar] [CrossRef]
  249. Kenawy, A.; Dailin, D.J.; Abo-Zaid, G.A.; Malek, R.A.; Ambehabati, K.K.; Zakaria, K.H.N.; El Enshasy, H.A. Biosynthesis of antibiotics by PGPR and their roles in biocontrol of plant diseases. In Plant Growth Promoting Rhizobacteria for Sustainable Stress Management: Volume 2: Rhizobacteria in Biotic Stress Management; Springer: Singapore, 2019; pp. 1–35. [Google Scholar]
  250. Aydin, M.H. Rhizoctonia Solani and Its Biological Control. Türkiye Tarımsal Araştırmalar Derg. 2022, 9, 118–135. [Google Scholar] [CrossRef]
  251. Katara, S.; Devki, V.G.; Neelam, D.; Kant, R. Role of bacteria and fungi in antibiotic production. Antibiotics and Antimicrobial Resistance Genes in the Environment. Pharma Innov. J. 2021, 1, 31–42. [Google Scholar]
  252. Mitra, A.; Chatterjee, S.; Kataki, S.; Rastogi, R.P.; Gupta, D.K. Bacterial tolerance strategies against lead toxicity and their relevance in bioremediation application. Environ. Sci. Pollut. Res. 2021, 28, 14271–14284. [Google Scholar] [CrossRef]
  253. Mukherjee, P.; Mitra, A.; Roy, M. Halomonas rhizobacteria of Avicennia marina of Indian Sundarbans promote rice growth under saline and heavy metal stresses through exopolysaccharide production. Front. Microbiol. 2019, 10, 1207. [Google Scholar] [CrossRef]
  254. Pietri, G.P.; Tontini, M.; Brogioni, B.; Oldrini, D.; Robakiewicz, S.; Henriques, P.; Malić, S. Elucidating the structural and minimal protective epitope of the serogroup X meningococcal capsular polysaccharide. Front. Mol. Biosci. 2021, 8, 745360. [Google Scholar] [CrossRef] [PubMed]
  255. Gupta, P.; Diwan, B. Bacterial exopolysaccharide mediated heavy metal removal: A review on biosynthesis, mechanism and remediation strategies. Biotechnol. Rep. 2017, 13, 58–71. [Google Scholar] [CrossRef] [PubMed]
  256. Srivastava, S.; Sharma, S. Insight into Exopolysaccharide-Mediated Stress Tolerance in Plants: A Feasible Approach towards the Development of Next-Generation Bioformulations. J. Soil Sci. Plant Nutr. 2023, 23, 22–33. [Google Scholar] [CrossRef]
  257. Majumdar, A.; Afsal, F.; Pathak, S.; Upadhayay, M.K.; Roychowdhury, T.; Srivastava, S. Molecular Aspects of Arsenic Responsive Microbes in Soil-Plant-Aqueous Triphasic Systems. In Global Arsenic Hazard: Ecotoxicology and Remediation; Springer: Singapore, 2022; pp. 291–312. [Google Scholar]
  258. Liu, Y.; Ali, A.; Su, J.-F.; Li, K.; Hu, R.-Z.; Wang, Z. Microbial-Induced Calcium Carbonate Precipitation: Influencing Factors, Nucleation Pathways, and Application in Waste Water Remediation. Sci. Total Environ. 2022, 860, 160439. [Google Scholar] [CrossRef] [PubMed]
  259. Brahim, B.G.; Ouhdouch, Y. Management of Tomato Foot and Root Rot (TFRR) by Biocontrol Agents with Emphasis on Factors Affecting its Effectiveness. In Probiotics and Plant Health; Springer: Singapore, 2017; pp. 1–19. [Google Scholar]
  260. Asaf, S.; Khan, A.L.; Khan, M.A.; Imran, Q.M.; Yun, B.-W.; Lee, I.-J. Osmoprotective Functions Conferred to Soybean Plants via Inoculation with Sphingomonas sp. LK11 and Exogenous Trehalose. Microbiol. Res. 2017, 205, 135–145. [Google Scholar] [CrossRef] [PubMed]
  261. Ilangumaran, G.; Smith, D.L. Plant Growth Promoting Rhizobacteria in Amelioration of Salinity Stress: A Systems Biology Perspective. Front. Plant Sci. 2017, 8, 1768. [Google Scholar] [CrossRef] [PubMed]
  262. John, C.J.; Kumar, S.; Ge, M. Probiotic Prospects of PGPR for Green and Sustainable Agriculture. Arch. Phytopathol. Plant Protec. 2020, 53, 899–914. [Google Scholar] [CrossRef]
  263. Dakora, F.D.; Matiru, V.N.; Kanu, A.S. Rhizosphere Ecology of Lumichrome and Riboflavin, Two Bacterial Signal Molecules Eliciting Developmental Changes in Plants. Front. Plant Sci. 2015, 6, 700. [Google Scholar] [CrossRef]
  264. Ayuso-Calles, M.; Flores-Félix, J.D.; Rivas, R. Overview of the Role of Rhizobacteria in Plant Salt Stress Tolerance. Agronomy 2021, 11, 1759. [Google Scholar] [CrossRef]
  265. Asaf, S.; Jan, R.; Khan, M.A.; Khan, A.L.; Asif, S.; Bilal, S.; Ahmad, W.; Waqas, M.; Kim, K.M.; Ahmed, A.H.; et al. Unraveling the mutualistic interaction between endophytic Curvularia lunata CSL1 and tomato to mitigate cadmium (Cd) toxicity via transcriptomic insights. Sci. Total Environ. 2023, 861, 160542. [Google Scholar] [CrossRef]
  266. Freitas, E.V.; Nascimento, C.W.; Souza, A.; Silva, F.B. Citric acid-assisted phytoextraction of lead: A field experiment. Chemosphere 2013, 92, 213–217. [Google Scholar] [CrossRef]
  267. Azubuike, C.C.; Chikere, C.B.; Okpokwasili, G.C. Bioremediation techniques-classification based on site of application: Principles, advantages, limitations and prospects. World J. Microbiol. Biotechnol. 2016, 32, 180. [Google Scholar] [CrossRef] [PubMed]
  268. Khodaverdiloo, H.; Han, F.X.; Hamzenejad Taghlidabad, R.; Karimi, A.; Moradi, N.; Kazery, J.A. Potentially toxic element contamination of arid and semi-arid soils and its phytoremediation. Arid Land Res. Manag. 2020, 34, 361–391. [Google Scholar] [CrossRef]
  269. Abatenh, E.; Gizaw, B.; Tsegaye, Z.; Wassie, M. The role of microorganisms in bioremediation—A review. Open J. Environ. Biol. 2017, 2, 38–46. [Google Scholar] [CrossRef]
  270. Yang, S.Z.; Jin, H.J.; Wei, Z.; He, R.X.; Ji, Y.J.; Lim, X.M.; Shao-Peng, Y.U. Bioremediation of oil spills in cold environments: A review. Pedosphere 2009, 19, 371–381. [Google Scholar] [CrossRef]
  271. Megharaj, M.; Ramakrishnan, B.; Venkateswarlu, K.; Sethunathan, N.; Naidu, R. Bioremediation Approaches for Organic Pollutants: A Critical Perspective. Environ. Int. 2011, 37, 1362–1375. [Google Scholar] [CrossRef]
  272. Javanbakht, V.; Alavi, S.A.; Zilouei, H. Mechanisms of Heavy Metal Removal Using Microorganisms as Biosorbent. Water Sci. Technol. 2014, 69, 1775–1787. [Google Scholar] [CrossRef]
  273. Oka, T.; Sameshima, Y.; Koga, T.; Kim, H.; Goto, M.; Furukawa, K. Protein Omannosyltransferase a of Aspergillus awamori Is Involved in O-Mannosylation of Glucoamylase I. Microbiology 2005, 151, 3657–3667. [Google Scholar] [CrossRef]
  274. Timková, I.; Sedláková-Kaduková, J.; Pristaš, P. Biosorption and Bioaccumulation Abilities of Actinomycetes/Streptomycetes Isolated from Metal Contaminated Sites. Separations 2018, 5, 54. [Google Scholar] [CrossRef]
  275. Wang, J.; Li, Q.; Li, M.M.; Chen, T.H.; Zhou, Y.F.; Yue, Z.B. Competitive Adsorption of Heavy Metal by Extracellular Polymeric Substances (EPS) Extracted from Sulfate Reducing Bacteria. Bioresour. Technol. 2014, 163, 374–376. [Google Scholar] [CrossRef]
  276. Park, J.H.; Chon, H.T. Characterization of Cadmium Biosorption by Exiguobacterium sp. Isolated from Farmland Soil near Cu-Pb-Zn Mine. Environ. Sci. Pollut. Res. 2016, 23, 11814–11822. [Google Scholar] [CrossRef] [PubMed]
  277. Phulia, V.; Jamwal, A.; Saxena, N.; Chadha, N.K.; Muralidhar, A.P.; Prusty, A.K. Technologies in Aquatic Bioremediation. In Freshwater Ecosystem and Xenobiotics; Discovery Publishing House PVT. Ltd.: New Delhi, India, 2013; pp. 65–91. [Google Scholar]
  278. Couto, N.; Fritt-Rasmussen, J.; Jensen, P.E.; Højrup, M.; Rodrigo, A.P.; Ribeiro, A.B. Suitability of Oil Bioremediation in an Arctic Soil Using Surplus Heating from an Incineration Facility. Environ. Sci. Pollut. Res. 2014, 21, 6221–6227. [Google Scholar] [CrossRef] [PubMed]
  279. Varjani, S.J.; Upasani, V.N. A New Look on Factors Affecting Microbial Degradation of Petroleum Hydrocarbon Pollutants. Int. Biodeterior. Biodegrad. 2017, 120, 71–83. [Google Scholar] [CrossRef]
  280. Sharma, J. Advantages and limitations of in situ methods of bioremediation. Recent Adv. Biol. Med. 2019, 5, 955923. [Google Scholar] [CrossRef]
  281. Srinath, T.; Verma, T.; Ramteke, P.W.; Garg, S.K. Chromium (VI) biosorption and bioaccumulation by chromate resistant bacteria. Chemosphere 2002, 48, 427–435. [Google Scholar] [CrossRef]
  282. Vijayaraghavan, K.; Yun, Y.S. Bacterial biosorbents and biosorption. Biotechnol. Adv. 2008, 26, 266–291. [Google Scholar] [CrossRef]
  283. Issazadeh, K.; Jahanpour, N.; Pourghorbanali, F.; Raeisi, G.; Faekhondeh, J. Heavy Metals Resistance by Bacterial Strains. Ann. Biol. Res. 2013, 4, 60–63. [Google Scholar]
  284. Babu, S.O.; Hossain, M.B.; Rahman, M.S.; Rahman, M.; Ahmed, A.S.; Hasan, M.M.; Rakib, A.; Emran, T.B.; Xiao, J.; Simal-Gandara, J. Phytoremediation of toxic metals: A sustainable green solution for clean environment. Appl. Sci. 2021, 11, 10348. [Google Scholar] [CrossRef]
  285. Iram, S.; Shabbir, R.; Zafar, H.; Javaid, M. Biosorption and Bioaccumulation of Copper and Lead by Heavy Metal-Resistant Fungal Isolates. Arab. J. Sci. Eng. 2015, 40, 1867–1873. [Google Scholar] [CrossRef]
  286. Sarwar, N.; Imran, M.; Shaheen, M.R.; Ishaque, W.; Kamran, M.A.; Matloob, A.; Rehim, A.; Hussain, S. Phytoremediation Strategies for Soils Contaminated with Heavy Metals: Modifications and Future Perspectives. Chemosphere 2017, 171, 710–721. [Google Scholar] [CrossRef]
  287. Iwamoto, T.; Nasu, M. Current bioremediation practice and perspective. J. Biosci. Bioeng. 2001, 92, 1–8. [Google Scholar] [CrossRef] [PubMed]
  288. Favas, P.J.C.; Pratas, J.; Varun, M.; D’Souza, R.; Paul, M.S. Phytoremediation of Soils Contaminated with Metals and Metalloids at Mining Areas: Potential of Native Flora. Environments 2014, 3, 485–516. [Google Scholar]
  289. Tak, H.I.; Ahmad, F.; Babalola, O.O. Advances in the application of plant growth-promoting rhizobacteria in phytoremediation of heavy metals. In Reviews of Environmental Contamination and Toxicology; Whitacre, D.M., Ed.; Springer: New York, NY, USA, 2013; Volume 223, pp. 33–52. [Google Scholar]
  290. Nadeem, S.M.; Zahir, Z.A.; Naveed, M.; Nawaz, S. Mitigation of salinity-induced negative impact on the growth and yield of wheat by plant growth-promoting rhizobacteria in naturally saline conditions. Ann. Microbiol. 2013, 63, 225–232. [Google Scholar] [CrossRef]
  291. Bücker-Neto, L.; Paiva, A.L.S.; Machado, R.D.; Arenhart, R.A.; Margis-Pinheiro, M. Interactions between plant hormones and heavy metals responses. Genet. Mol. Biol. 2017, 40 (Suppl. S1), 373–386. [Google Scholar] [CrossRef] [PubMed]
  292. Solanki, R.; Dhankhar, R. Biochemical changes and adaptive strategies of plants under heavy metal stress. Biologia 2011, 66, 195–204. [Google Scholar] [CrossRef]
  293. Jan, S.; Parray, J.A. Approaches to Heavy Metal Tolerance in Plants; Springer: New Delhi, India, 2016. [Google Scholar]
  294. Emamverdian, A.; Ding, Y.; Mokhberdoran, F.; Xie, Y. Heavy metal stress and some mechanisms of plant defense response. Sci. World J. 2015, 2015, 756120. [Google Scholar] [CrossRef]
  295. Zhao, Y.; Yao, J.; Yuan, Z.; Wang, T.; Zhang, Y.; Wang, F. Bioremediation of Cd by Strain GZ-22 Isolated from Mine Soil Based on Biosorption and Microbially Induced Carbonate Precipitation. Environ. Sci. Pollut. Res. 2017, 24, 372–380. [Google Scholar] [CrossRef]
  296. Gong, X. Kinetic and Equilibrium Studies on the Adsorption of Pb (II), Cd (II) and Cu (II) by Rape Straw. Adsorpt. Sci. Technol. 2013, 31, 559–571. [Google Scholar] [CrossRef]
  297. Wang, X.S. Cd (II) Removal by Marine Arthrobacter Protophormiae Biomass: Mechanism Characterization and Adsorption Performance. Desalin. Water Treat. 2013, 51, 7710–7720. [Google Scholar] [CrossRef]
  298. Li, L.; Qian, C.; Cheng, L.; Wang, R. A Laboratory Investigation of Microbe-Inducing CdCO3 Precipitate Treatment in Cd2+ Contaminated Soil. J. Soils Sediments 2010, 10, 248–254. [Google Scholar] [CrossRef]
  299. Burns, J.L.; Ginn, B.R.; Bates, D.J.; Dublin, S.N.; Taylor, J.V.; Apkarian, R.P.; Amaro-Garcia, S.; Neal, A.L.; Dichristina, T.J. Outer Membrane-Associated Serine Protease Involved in Adhesion of Shewanella oneidensis to Fe(III) Oxides. Environ. Sci. Technol. 2010, 44, 68–73. [Google Scholar] [CrossRef] [PubMed]
  300. Huang, J.; Elzinga, E.J.; Brechbühl, Y.; Voegelin, A.; Kretzschmar, R. Impacts of Shewanella Putrefaciens Strain CN-32 Cells and Extracellular Polymeric Substances on the Sorption of as (V) and As (III) on Fe (III)-(Hydr)oxides. Environ. Sci. Technol. 2011, 45, 2804–2810. [Google Scholar] [CrossRef] [PubMed]
  301. Rascio, N.; Navari-Izzo, F. Heavy Metal Hyperaccumulating Plants: How and Why Do They Do It? And What Makes Them So Interesting? Plants 2011, 180, 169–181. [Google Scholar] [CrossRef] [PubMed]
  302. Chaperon, S.; Sauvé, S. Toxicity Interactions of Cadmium, Copper, and Lead on Soil Urease and Dehydrogenase Activity in Relation to Chemical Speciation. Ecotoxicol. Environ. Saf. 2008, 70, 1–9. [Google Scholar] [CrossRef] [PubMed]
  303. Jusselme, M.D.; Miambi, E.; Mora, P.; Diouf, M.; Rouland-Lefèvre, C. Increased Lead Availability and Enzyme Activities in Root-Adhering Soil of Lantana camara during Phytoextraction in the Presence of Earthworms. Sci. Total Environ. 2013, 445, 101–109. [Google Scholar] [CrossRef]
  304. Ciarkowska, K.; Sołek-Podwika, K.; Wieczorek, J. Enzyme Activity as an Indicator of Soil-Rehabilitation Processes at a Zinc and Lead Ore Mining and Processing Area. J. Environ. Manag. 2014, 132, 250–256. [Google Scholar] [CrossRef]
  305. Al-Maqdi, K.A.; Elmerhi, N.; Athamneh, K.; Bilal, M.; Alzamly, A.; Ashraf, S.S.; Shah, I. Challenges and Recent Advances in Enzyme-Mediated Wastewater Remediation—A Review. Nanomaterials 2021, 11, 3124. [Google Scholar] [CrossRef]
  306. Wasilkowski, D.; Śwędzioł, Z.; Mrozik, A. Przydatność Genetycznie Modyfikowanych Mikroorganizmów do Bioremediacji Zanieczyszczonych Środowisk. Chemik 2012, 66, 817–826. (In Polish) [Google Scholar]
  307. Sharma, B.; Dangi, A.K.; Shukla, P. Contemporary Enzyme Based Technologies for Bioremediation: A Review. J. Environ. Manag. 2018, 210, 10–22. [Google Scholar] [CrossRef]
  308. Wolejko, E.; Wydro, U.; Loboda, T. The Ways to Increase Efficiency of Soil Bioremediation. Ecol. Chem. Eng. 2016, 23, 155. [Google Scholar] [CrossRef]
  309. Liang, Y.; Jiao, S.; Wang, M.; Yu, H.; Shen, Z. A CRISPR/Cas9-Based Genome Editing System for Rhodococcus ruber TH. Metab. Eng. 2020, 57, 13–22. [Google Scholar] [CrossRef] [PubMed]
  310. Azad, M.A.K.; Amin, L.; Sidik, N.M. Genetically engineered organisms for bioremediation of pollutants in contaminated sites. Chin. Sci. Bull. 2014, 59, 703–714. [Google Scholar] [CrossRef]
  311. Mujawar, S.Y.; Shamim, K.; Vaigankar, D.C.; Dubey, S.K. Arsenite Biotransformation and Bioaccumulation by Klebsiella pneumoniae Strain SSSW7 Possessing Arsenite Oxidase (aioA) Gene. Biometals 2019, 32, 65–76. [Google Scholar] [CrossRef] [PubMed]
  312. Singh, N.; Kumar, A.; Sharma, B. Role of Fungal Enzymes for Bioremediation of Hazardous Chemicals. In Recent Advancement in White Biotechnology through Fungi: Volume 3: Perspective for Sustainable Environments; Gupta, V.K., Sharma, G.D., Tuohy, M., Eds.; Springer: Cham, Switzerland, 2019; pp. 237–256. [Google Scholar]
  313. Ghosal, D.; Ghosh, S.; Dutta, T.K.; Ahn, Y. Current State of Knowledge in Microbial Degradation of Polycyclic Aromatic Hydrocarbons (PAHS): A Review. Front. Microbiol. 2016, 7, 1369. [Google Scholar] [CrossRef]
  314. Van Nostrand, J.D.; He, Z.; Zhou, J. Use of Functional Gene Arrays for Elucidating In Situ Biodegradation. Front. Microbiol. 2012, 3, 339. [Google Scholar] [CrossRef] [PubMed]
  315. Chen, Z.; Zheng, Y.; Ding, C.; Ren, X.; Yuan, J.; Sun, F.; Li, Y. Integrated Metagenomics and Molecular Ecological Network Analysis of Bacterial Community Composition during the Phytoremediation of Cadmium-Contaminated Soils by Bioenergy Crops. Ecotoxicol. Environ. Saf. 2017, 145, 111–118. [Google Scholar] [CrossRef]
  316. Feng, G.; Xie, T.; Wang, X.; Bai, J.; Tang, L.; Zhao, H.; Wei, W.; Wang, M.; Zhao, Y. Metagenomic Analysis of Microbial Community and Function Involved in Cd-Contaminated Soil. BMC Microbiol. 2018, 18, 13. [Google Scholar] [CrossRef]
  317. Datta, S.; Rajnish, K.N.; Samuel, M.S.; Pugazlendhi, A.; Selvarajan, E. Metagenomic Applications in Microbial Diversity, Bioremediation, Pollution Monitoring, Enzyme, and Drug Discovery: A Review. Environ. Chem. Lett. 2020, 18, 1229–1241. [Google Scholar] [CrossRef]
  318. Malla, M.A.; Dubey, A.; Yadav, S.; Kumar, A.; Hashem, A.; Abd_Allah, E.F. Understanding and Designing the Strategies for the Microbe-Mediated Remediation of Environmental Contaminants Using Omics Approaches. Front. Microbiol. 2018, 9, 1132. [Google Scholar] [CrossRef]
  319. Tripathi, M.; Singh, D.N.; Vikram, S.; Singh, V.S.; Kumar, S. Metagenomic approach towards bioprospection of novel biomolecule(s) and environmental bioremediation, Annu. Res. Rev. Biol. 2018, 22, 1–12. [Google Scholar]
  320. Jackson, S.A.; Borchert, E.; OGara, F.; Dobson, A.D. Metagenomics for the Discovery of Novel Biosurfactants of Environmental Interest from Marine Ecosystems. Curr. Opin. Biotechnol. 2015, 33, 176–182. [Google Scholar] [CrossRef] [PubMed]
  321. Williams, W.; Trindade, M. Metagenomics for the discovery of novel biosurfactants. In Functional Metagenomics: Tools and Applications; Springer: Cham, Switzerland, 2017; pp. 95–117. [Google Scholar]
  322. Awasthi, M.K.; Guo, D.; Awasthi, S.K.; Wang, Q.; Chen, H.; Liu, T.; Duan, Y.; Soundari, P.G.; Zhang, Z. Recent Advances in Phytoremediation of Toxic Metals from Contaminated Sites: A Road Map to a Safer Environment. In Bioremediation of Industrial Waste for Environmental Safety; Bharagava, R.N., Saxena, G., Eds.; Springer: Singapore, 2020; pp. 77–112. [Google Scholar]
  323. Arumugam, M.; Harrington, E.D.; Foerstner, K.U.; Raes, J.; Bork, P. SmashCommunity: A Metagenomic Annotation and Analysis Tool. Bioinformatics 2010, 26, 2977–2978. [Google Scholar] [CrossRef] [PubMed]
  324. El-Hadidi, M.; Ruscheweyh, H.-J.; Huson, D.H. Improved Metagenome Analysis Using MEGAN5. In Proceedings of the Joint 21st Annual International Conference on Intelligent Systems for Molecular Biology (ISMB) and 12th European Conference on Computational Biology (ECCB), Berlin, Germany, 21–23 July 2013. [Google Scholar]
  325. Markowitz, V.M.; Ivanova, N.N.; Szeto, E.; Palaniappan, K.; Chu, K.; Dalevi, D.; Chen, I.-M.A.; Grechkin, Y.; Dubchak, I.; Anderson, I.; et al. IMG/M: A Data Management and Analysis System for Metagenomes. Nucleic Acids Res. 2007, 36, D534–D538. [Google Scholar] [CrossRef] [PubMed]
  326. Rizwan, M.; Singh, M.; Mitra, C.K.; Morve, R.K. Ecofriendly Application of Nanomaterials: Nanobioremediation. J. Nanopart. Res. 2014, 2014, 431787. [Google Scholar] [CrossRef]
  327. Karimi-Maleh, H.; Karaman, C.; Karaman, O.; Karimi, F.; Vasseghian, Y.; Fu, L.; Mirabi, A. Nanochemistry Approach for the Fabrication of Fe and N Co-Decorated Biomass-Derived Activated Carbon Frameworks: A Promising Oxygen Reduction Reaction Electrocatalyst in Neutral Media. J. Nanostruct. Chem. 2022, 12, 429–439. [Google Scholar] [CrossRef]
  328. Rafeeq, H.; Hussain, A.; Ambreen, A.; Waqas, M.; Bilal, M.; Iqbal, H. Functionalized nanoparticles and their environmental remediation potential: A review. J. Nanostruct. Chem. 2022, 12, 1007–1031. [Google Scholar] [CrossRef]
  329. Brunner, T.J.; Wick, P.; Manser, P.; Spohn, P.; Grass, R.N.; Limbach, L.K.; Bruinink, A.; Stark, W.J. In vitro cytotoxicity of oxide nanoparticles: Comparison to asbestos, silica, and the effect of particle solubility. Environ. Sci. Technol. 2006, 40, 4374–4381. [Google Scholar] [CrossRef]
  330. Kumar, A.; Choudhary, P.; Kumar, A.; Camargo, P.H.; Krishnan, V. Recent advances in plasmonic photocatalysis based on TiO2 and noble metal nanoparticles for energy conversion, environmental remediation, and organic synthesis. Small 2022, 18, 2101638. [Google Scholar] [CrossRef]
  331. Fernandes, J.P.; Mucha, A.P.; Francisco, T.; Gomes, C.R.; Almeida, C.M.R. Silver nanoparticles uptake by salt marsh plants—Implications for phytoremediation processes and effects in microbial community dynamics. Mar. Pollut. Bull. 2017, 119, 176–183. [Google Scholar] [CrossRef]
  332. Yadav, K.K.; Singh, J.K.; Gupta, N.; Kumar, V. A review of nanobioremediation technologies for environmental cleanup: A novel biological approach. J. Mater. Environ. Sci. 2017, 8, 740–757. [Google Scholar]
  333. Ma, X.; Geisler-Lee, J.; Deng, Y.; Kolmakov, A. Interactions between engineered nanoparticles (ENPs) and plants: Phytotoxicity, uptake and accumulation. Sci. Total Environ. 2010, 408, 3053–3061. [Google Scholar] [CrossRef] [PubMed]
  334. Rossi, L.; Zhang, W.; Schwab, A.P.; Ma, X. Uptake, accumulation, and in planta distribution of coexisting cerium oxide nanoparticles and cadmium in Glycine max (L.) Merr. Environ. Sci. Technol. 2017, 51, 12815–12824. [Google Scholar] [CrossRef] [PubMed]
  335. López-Luna, J.; Silva-Silva, M.J.; Martinez-Vargas, S.; Mijangos-Ricardez, O.F.; González-Chávez, M.C.; Solís-Domínguez, F.A.; Cuevas-Díaz, M.C. Magnetite nanoparticle (NP) uptake by wheat plants and its effect on cadmium and chromium toxicological behavior. Sci. Total Environ. 2016, 565, 941–950. [Google Scholar] [CrossRef] [PubMed]
  336. Rossi, L.; Sharifan, H.; Zhang, W.; Schwab, A.P.; Ma, X. Mutual effects and in planta accumulation of co-existing cerium oxide nanoparticles and cadmium in hydroponically grown soybean (Glycine max (L.) Merr.). Environ. Sci. Nano 2018, 5, 150. [Google Scholar] [CrossRef]
Figure 1. Factors affecting cadmium speciation in soil, and its toxic impacts on plant physiology, morphology and metabolism.
Figure 1. Factors affecting cadmium speciation in soil, and its toxic impacts on plant physiology, morphology and metabolism.
Plants 12 03147 g001
Figure 2. Microbially mediated direct mechanisms for contaminant detoxification.
Figure 2. Microbially mediated direct mechanisms for contaminant detoxification.
Plants 12 03147 g002
Figure 3. Microbially mediated indirect mechanisms for contaminant detoxification.
Figure 3. Microbially mediated indirect mechanisms for contaminant detoxification.
Plants 12 03147 g003
Table 1. Impact of Cd toxicity on different plant species.
Table 1. Impact of Cd toxicity on different plant species.
Plant SpeciesLevel of CdChanges/DamagesReferences
Seed germination and seedling growth
  Pisum sativum L.20–500 µMInhibition of proteolytic enzymes and restriction of starch metabolism, leading to the failure of protein mobilization in seeds. [73]
  Ocimum basilicum L.20 mg kg−1Alterations in the embryo and reductions in the oil contents of seeds. [74]
  Brassica oleracea L.5 mg L−1Decreased seed germination with an increase in MDA contents, electrolyte leakage and H2O2 contents.[75]
  Sassafras tzumu Hemsl.100 mg kg−1Restricted seedling growth and germination, and impairment of photosynthesis at higher doses.[63]
  Brassica juncea L.15 mg kg−1Disintegration occured in roots and shoots, and levels of ROS increased in plant shoots.[76]
  Oryza sativa L.50 μMLower seed germination rate due to the hyperaccumulation of Cd.[53]
  Zea mays L.100 mg kg−1Reduced seedling growth and activity of cellular antioxidants.[77]
Growth and development
  Cicer arietinum L.50 μMReduction in growth and appearance of symptoms of necrosis and chlorosis in leaves. [78]
  Ipomoea aquatica Forsk Reduced growth and development of root and shoots.[79]
  Lens culinaris50 µg g−1Increased electrolyte leakage and ROS production, resulting in lower plant growth. [80]
  Medicago sativa L.10 mg kg−1Higher concentrations damaged proteins, changed cell wall infrastructure and metabolism, and limited growth.[81]
Osmolytes and photosynthesis
  Cajanus cajan L.10 mg kg−1Lower organic osmolytes ultimately caused a disturbance in osmotic adjustments.[82]
  Vigna angularis64 mg L−1Cellular antioxidants decreased at higher concentrations, resulting in the lower production of low-molecular-weight osmolytes. [83]
  Zea mays L.150 μMReduction in photosynthetic pigments and gas exchange traits. [58]
  Coriandrum sativum L.20 µM L−1Inhibited gas exchange traits and biochemical processes.[45]
  Capsicum annuum L.500 ppmInduction of stomatal closure, resulting in decreased photosynthetic pigments, a smaller stomatal size and reduced transpiration.[84]
  Mentha arvensis150 mg Kg−1Reductions in mineral assimilation, photosynthetic attributes and photosynthetic pigments occurred. [85]
Table 2. Microbe-assisted phytoremediation of cadmium from soil under different experimental settings.
Table 2. Microbe-assisted phytoremediation of cadmium from soil under different experimental settings.
ExperimentContamination LevelMicroorganisms PlantResultsReferences
Pot10 mg kg−1Pseudomonas fluorescensHordeum vulgarePhyto-stabilization of Cd due to PGPR activity, increased uptake of essential plant nutrients and enhanced plant growth attributes.[90]
Greenhouse pot10.7 mg kg−1 Cd Bacillus spp.Solanum nigrumIncreased plant growth attributes under Cd stress, enhanced absorption of P and Fe as well as increased Cd contents in aerial plant parts. [91]
Incubation study200 μg/mLKlebsiella michiganensisOryza sativaCd bioaccumulation by tolerant bacteria with a concurrent decline in its uptake by plants.[92]
Pot50 mg kg−1Cupriavidus necator, Sphingomonas and Curtobacterium spp.Brassica napusIncreased plant biomass and growth traits under Cd contamination in inoculated treatments along with enhanced Cd uptake by aerial plant parts.[93]
Pot0, 50, and 100 mg L−1Rhizobium pusenseGlycine maxDecreased soybean root Cd contents by 45.9 and 35.3%, respectively, at contamination levels of 50 and 100 mg L−1.[94]
Pot(0, 25, 50, 75, 100, 150 and 200 mg kg−1)Enterobacter cloacae, Klebsiella pneumonia and Klebsiella spp. Pennisetum giganteumCombined application of rhizobacteria increased the bioaccumulation factor of Cd for plants. [95]
Pot0, 5, 10, 15 and 20 mg kg−1Serratia marcescensChrysopogon zizanioidesIncreased phytoaccumulation of Cd, soil biological health, as well as antioxidative potential of plants under bacterial inoculation. Maximum Cd phytoextraction in roots (289.47 mg kg−1), leaves (59.38 mg kg−1) and stem (88.33 mg kg−1) with a concomitant increase in plant biomass (9.68–45.99%). [96]
Field2.2 mg kg−1Rhizobium leguminosarum, Bacillus simplex, Luteibacter sp. + Variovorax sp., Pseudomonas fluorescensLathyrus sativusIncreased growth attributes as well as nodule number, and plant nutrient uptake, and phytoaccumulation along with reduced rhizosphere concentration of Cd (61%).[97]
Pot50 and 100 mg kg−1Fungi “Funneliformis mosseae” and bacteria Enterobacter sp. and Enterobacter ludwigiiLycopersicon esculentumIncreased dry weights of shoots (119–154%) and roots (91–173%) under combined inoculation. Furthermore, decreased Cd concentrations in shoots as well as translocation factors under inoculated treatments were observed. [98]
Pot0, 0.25, 0.5, 0.75 and 1 M CdSO4Serratia marcescensOryza sativaIncreased Cd removal from soil (66 mg kg−1 after 20 days).[99]
Incubation study0, 0.25, and 0.5 mM CdStenotrophomonas maltophiliaCapsicum annuumUnder Cd stress, increased root lengths (1.46 times) in the inoculated treatment compared to the control. [100]
Pot15 mg kg−1Variovorax paradoxus, Rhizobium leguminosarum and fungus Glomus spp. Pisum sativum and Brassica junceaMore prominent positive effect of consortium inoculation on Pisum sativum rather than Brassica juncea, in terms of growth, nutrient uptake and increased seed Cd concentration. [101]
Greenhouse pot2.12 mg kg−1Bacillus megaterium, Glomus mosseae, and Piriformospora indicaSolanum nigrumCd accumulation (104%) observed under the combined application of Bacillus megaterium and Glomus mosseae in addition to increased soil biological health under contaminated conditions.[102]
Pot100 mg kg−1Bacillus sp.Oryza sativaReduced bioavailable Cd concentration (39.3%), increased phytoextraction efficiency of rice for Cd (48.2%) and increased rice growth and yield traits under inoculation compared to the control. [103]
Greenhouse 0.064 mg L−1Klebsiella huaxiensis and Pantoea cypripediiPennisetum purpurenumEnhanced Cd phytoaccumulation in all variants of the tested plant (28.43–38.07 mg kg−1). [104]
Pot30 μmol L−1Enterobacter cloacaeSolanum nigrumIncreased soil Cd phytoextraction by plants along with increased plant growth under Cd stress.[105]
Pot0, 0.25, and 0.50 mg kg−1Bacillus spp. Oryza sativaIncreased Cd immobilization in soil by its surface adsorption concomitant with increased plant growth.[106]
Incubation0.4 mM CdCl2Pseudomonas aeruginosa and Burkholderia gladioliSolanum lycopersicumAlleviation of Cd toxicity in plants was evident by an increase in phenolic compounds, osmolytes and low-molecular-weight organic acids.[107]
Growth room trial0–400 μg/mLEnterobacter cloacaeOryza sativaIncreased Cd removal efficiency (72.11%) against a contamination level of 400 μg/mL[108]
Pot2 gCurtobacterium oceanosedimentumCapsicum frutescensIncreased root (58%) and shoot (60%) lengths, enhanced accumulation of Cd in roots compared to shoots under bacterial inoculation.[109]
Pot1.68 mg kg−1Buttiauxella, Pedobacter, Aeromonas eucrenophila, and Ralstonia pickettiiSedum plumbizincicolaInoculation led to reduced reducible and residual Cd and increased Cd availability coefficients by 1.15–6.41 units. Cd contents in shoots (29.63–46.01%) and roots (11.42–84.47%), bioconcentration factor (2.13–2.72) and Cd removal rate (48.25%) compared to the control treatment [110]
Table 3. Microbial biosorption by different microbes.
Table 3. Microbial biosorption by different microbes.
Species nameInitial Cd ConcentrationExperimental MediumCd Remediation EfficiencyReferences
Bacterial species
Pseudomonas fluorescens and Bacillus subtilis150 mg L−1Soil16.7[162]
Pseudomonas sp. DDT-10.9 mg kg−1Soil40.3[163]
Kocuria rhizophila150 mg L−1Aqueous9.07 mg g−1[164]
Klebsiella michiganensis1000 µg ml−1Soil97%[165]
Enterobacter sp3500 µg ml−1Soil95%[166]
Rhodobacter sphaeroides65.33 mg kg−1Soil30.7[167]
Bacillus aryabhattai and Bacillus amyloliquefaciens250 mg L−1Soil96%[168]
Aspergillus sydowii50 mg kg−1Soil10.44%[169]
Cupriavidus sp.13.82 mg kg−1Soil58.2%[170]
Paenibacillus sp. and Bacillus sp.20 mg L−1Soil128.50%[171]
Bacillus sp. TZ5150 mg L−1Soil48.49%[172]
Acidithiobacillus caldus DX, Acidithiobacillus thiooxidans DX, Acidithiobacillus thiooxidans ZJ, Acidithiobacillus thiooxidans AO1, Ferroplasma acidiphilum DX, Acidithiobacillus caldus S1 and Leptospirillum ferriphilum DX9.09 mg kg−1Soil32.09%[173]
Cupriavidus sp. (KU168590), Ensifer sp. (KU168586), Burkholderia sp.
(KU168588), and Paenibacillus sp.
(KU168587)
0.21 mg kg−1Soil33.0%[174]
Enterobacter cloacae, Pseudomonas aeruginosa, and Klebsiella
Edwardsii
50 mg L−1Soil58.80%[175]
Bacillus subtilis147.75 mg kg−1Soil35.17%[176]
Burkholderia sp. and Bacillus sp.5 mMSoil84.17%[177]
Bacillus sp. 49 mg kg−1Soil43.53%[178]
Bacillus subtilis147.75 mg kg−1Soil18.56%[179]
Firmicutes sp. and Proteobacteria sp.14.9 mg kg−1Soil40.0[180]
Enterobacter hormaechei SFC3100 µg ml−1Soil90.21%[181]
Streptomyces pactum Act12 and Streptomyces Roche D741.62 mg kg−1Soil56.39%[182]
Bacillus velezensis-Soil1.65 μg g−1[183]
Fungi
Aspergillus niger, Aspergillus fumigatus, and Penicillium rubens0.6 mg L−1Soil79%[184]
Simplicillium chinense400 mg L−1Soil88%[185]
Aspergillus flflavus, Aspergillus gracilis, Aspergillus penicillioides, Aspergillus restrictus, and Sterigmatomyces halophilus1000 mg L−1Soil95%[186]
Phanerochaete chrysosporium25 mg L−1Soil96%[187]
Agaricus bisporus, Pleurotus platypus, and Calocybe indica--98.97[188]
Lactarius piperatus and Agaricus bisporus265 mg L−1Aqueous95%[189]
Rhizophagus intraradices-Soil38%[190]
Trichoderma harzianum147.75 mg kg−1Soil47.69%[176]
Algae/cyanobacteria
Asparagopsis armata150 mg L−1Aqueous10.6%[191]
Chaetoceros calcitrans and Tetracelmis chuii-Aqueous-[192]
Ulva lactuca80 mg L−1Aqueous41.0%[193]
Chara aculeolate-Aqueous23 mg g−1[194]
Chlorella pyrenoidosa1.5 mg L−1Aqueous45.45[195]
Scenedesmus acutus1.5 mg L−1Aqueous57.14[195]
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

Zulfiqar, U.; Haider, F.U.; Maqsood, M.F.; Mohy-Ud-Din, W.; Shabaan, M.; Ahmad, M.; Kaleem, M.; Ishfaq, M.; Aslam, Z.; Shahzad, B. Recent Advances in Microbial-Assisted Remediation of Cadmium-Contaminated Soil. Plants 2023, 12, 3147. https://doi.org/10.3390/plants12173147

AMA Style

Zulfiqar U, Haider FU, Maqsood MF, Mohy-Ud-Din W, Shabaan M, Ahmad M, Kaleem M, Ishfaq M, Aslam Z, Shahzad B. Recent Advances in Microbial-Assisted Remediation of Cadmium-Contaminated Soil. Plants. 2023; 12(17):3147. https://doi.org/10.3390/plants12173147

Chicago/Turabian Style

Zulfiqar, Usman, Fasih Ullah Haider, Muhammad Faisal Maqsood, Waqas Mohy-Ud-Din, Muhammad Shabaan, Muhammad Ahmad, Muhammad Kaleem, Muhammad Ishfaq, Zoya Aslam, and Babar Shahzad. 2023. "Recent Advances in Microbial-Assisted Remediation of Cadmium-Contaminated Soil" Plants 12, no. 17: 3147. https://doi.org/10.3390/plants12173147

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